2,731
views
1
recommends
+1 Recommend
2 collections
    1
    shares

      Interested in becoming an AMM published author?

      • Platinum Open Access with no APCs.
      • Fast peer review/Fast publication online after article acceptance.

      Check out the call for papers on our website https://amm-journal.org/index.php/2023/04/26/acta-materia-medica-call-for-papers-2/

      scite_
       
      • Record: found
      • Abstract: found
      • Article: found
      Is Open Access

      Recent advances in nano-targeting drug delivery systems for rheumatoid arthritis treatment

      review-article
      Bookmark

            Abstract

            Rheumatoid arthritis is a systemic inflammatory disease that can lead to articular cartilage destruction and periarticular bone erosion, thus ultimately compromising joint integrity and function. Anti-inflammatory drugs and biological agents are commonly used to treat rheumatoid arthritis, but they cannot selectively target inflamed joints, because of their systemic mechanisms, short half-lives and low bioavailability. Consequently, these agents must be used at high doses and delivered frequently, thereby increasing costs and the risk of adverse effects. Drug delivery systems, such as nanoparticles, liposomes and micelles, can significantly prolong drug half-life in the body and enable targeted delivery into the joints. In this review, we comprehensively describe the pathogenesis and clinical diagnosis of rheumatoid arthritis, and summarize recent advances in targeted therapeutic strategies, particularly nano-targeting systems for rheumatoid arthritis.

            Main article text

            1. INTRODUCTION

            Rheumatoid arthritis (RA) is a chronic systemic inflammatory disease that is typically characterized by persistent synovitis [1] and that manifests as symmetric polyarthritis of large and small joints, which may lead to joint and periarticular structural damage [2]. Although the etiology of RA is unclear, genetic and environmental factors appear to play key roles in the disease [3]. RA affects 0.5–1.0% of the population worldwide [4]; it affects women more often than men, and its prevalence is highest among individuals 35–50 years old [5].

            Anti-inflammatory drugs and biological agents are commonly used in RA treatment, but they cannot selectively target inflamed joints, because of their systemic mechanisms, short half-lives and low bioavailability. To circumvent these drawbacks, these agents must be delivered frequently at high doses, thereby increasing the risk of serious adverse effects in healthy joint tissues [6]. Recent research has focused on the development of nanoparticle drug delivery systems that actively or passively target inflamed joints [7] while also improving the release of insoluble drugs, thereby maximizing bioavailability and therapeutic efficacy. Such drug delivery systems can significantly prolong drug half-life in the body and promote drug accumulation in the joints [8]. In this review, we focus on the pathogenesis, pathophysiology and clinical diagnosis of RA, as well as on nanocarriers and other targeted methods that are already used, or that may be used in the future, to treat the disease.

            2. THE PATHOGENESIS AND PHYSIOLOGICAL CHARACTERISTICS OF RA

            Although the pathogenesis of RA remains unclear, studies have suggested that interactions between genetic and environmental factors lead to innate and adaptive immune responses that promote the development of the disease [9]. External stimuli may also activate innate immune responses by binding Toll-like receptors (TLRs) [10], which activate macrophages, dendritic cells, mast cells, neutrophils, T cells, B cells and fibroblasts [11]. Macrophages drive RA by stimulating neovascularization, clearing apoptotic immune cells, and promoting fibroblast proliferation and protease secretion. Activated macrophages also promote inflammation in RA by releasing pro-inflammatory cytokines, such as tumor necrosis factor (TNF)-α, interleukin (IL)-1 and IL-6, as well as reactive oxygen intermediates and prostaglandins [12]. Strong activation of macrophages also upregulates TLR2, TLR3, TLR4 and TLR7, as well as enzymes, cytokines and other inflammatory factors that promote synovial inflammation and cartilage destruction ( Figure 1 ) [13, 14].

            Figure 1 |

            Comparison of normal joints and rheumatoid arthritis joints.

            IL, interleukin; TNF, tumor necrosis factor.

            Activation of innate immunity causes dendritic cells to promote RA pathogenesis by taking up self-antigens and presenting them to T cells, thus leading to their activation or inhibition [15]. Neutrophils interact with fibroblast-like synoviocytes in the synovium, and consequently promote inflammatory and antigen-presenting phenotypes [16], whereas protein-protein interactions between the receptor activator of nuclear factor (NF)-κB ligand (RANKL) and its receptor, RANK, contribute to osteoclast differentiation in bone remodeling [17]. The RANKL/RANK pathway also promotes inflammation by activating transcription factors and signaling molecules such as NF-κB, JNK, AKT/PKB, ERK, Src kinase and p38 mitogen-activated protein kinase [18]. TNF receptor-associated factors (TRAFs) 1–7 also induce hyperimmune responses that contribute to RA; these factors bind the TNF receptor, IL-1 receptors and TLRs [19]. TRAF6 activates NF-κB, which in turn upregulates expression of genes encoding various inflammatory factors that drive synovitis and the destruction of cartilage and bone [20].

            Similarly to tumor tissues, RA tissues are affected by extravasation through leaky vasculature and subsequent inflammatory cell-mediated sequestration (ELVIS), angiogenesis, hypoxia and acidosis. The leaky vasculature of inflamed joints enables the penetration of nanodrugs, which are subsequently internalized by activated synovial cells [21]. The high metabolic demand and rapid growth of synovial membranes in inflamed synovial tissues lead to hypoxia in inflamed joints, thus causing hypoxia-inducible and vascular endothelial growth factors to induce angiogenesis and promote cell growth [22]. Angiogenesis is essential for synovial tissue proliferation and pannus formation. Joint synovial tissues show upregulation of angiogenesis-stimulating factors and downregulation of angiogenesis inhibitors, thus leading to synovial micro-angiogenesis [23]. Gal-9 induces angiogenesis via signaling pathways involving JNK, Erk1/2 and p38 [24]. Survivin, a member of a family of apoptosis inhibitors that block caspase activity [25], upregulates angiogenesis-associated proteins and activates the NOTCH pathway, thereby suppressing apoptosis [26]. Endothelial cells and synovial vessels in inflamed joints contain elevated concentrations of semaphorins, which may contribute to angiogenesis [27]. The pH of interstitial tissues in inflamed joints ranges between 6.0 and 7.0, and acidic pH enhances the inflammatory response of neutrophils [28]. These characteristics of RA can be exploited to achieved targeted treatment of the disease.

            3. DIAGNOSIS OF RA

            Common diagnostic methods for RA include magnetic resonance imaging (MRI), ultrasonography and assays for autoantibodies. MRI can effectively detect changes in inflamed soft tissues, such as synovitis, tenosynovitis and bone marrow edema, through multiplanar tomographic imaging of bone and soft tissue structures in inflamed joints. MRI can also be used to evaluate cartilage damage, bone erosion and tendon tears of peripheral joints. However, the technique is expensive and time-consuming, and each MRI scan can cover only a limited area of tissue. Ultrasonography, in contrast, enables real-time, relatively low-cost imaging of synovial proliferation and bone erosion in inflamed joints. Ultrasound has similar sensitivity and accuracy to MRI in the diagnosis of synovitis and tenosynovitis in patients with early rheumatoid arthritis [29]. However, ultrasonography cannot detect bone marrow edema. RA can be diagnosed on the basis of the presence of autoantibodies, which are produced in response to abnormal cellular and humoral immune responses. These autoantibodies include rheumatoid factor (RF), which comprises IgG, IgA and IgM, all of which are also present in healthy people but are elevated in individuals with RA; and anti-citrullinated protein antibody (ACPA), which promotes bone loss by activating macrophages or by binding citrullinated vimentin in cell membranes [30]. The concentration of RF and the diversity of its epitopes increase with the levels of pro-inflammatory cytokines. Although RF and ACPA are diagnostic markers of RA, approximately one-third of patients with RA are negative for RF and ACPA. Therefore, simultaneous detection of antibody to RA33, RF and ACPA can improve the diagnostic sensitivity of serological tests [31]. In addition, some researchers have reported that four additional biomarkers—angiotensinogen, serum amylase A-4 protein, vitamin D-binding protein and retinol-binding protein-4—can avoid false-negative findings and improve the accuracy of RA diagnosis [32].

            4. DRUGS USED TO TREAT RA

            Drugs used to treat RA include non-steroidal anti-inflammatory drugs (NSAIDs), disease-modifying anti-rheumatic drugs (DMARDs), glucocorticoids (GCs), biological agents and RNAs acting through RNA interference (RNAi) ( Table 1 ). These drugs can relieve pain, reduce injury, or efficiently slow the progress of disease [33, 34].

            Table 1 |

            Drugs commonly used to treat rheumatoid arthritis.

            Drug classDrugsAdverse effects
            NSAIDsAcetylsalicylates, naproxen, diclofenac, ibuprofen, etodolacNausea, abdominal pain, ulcers, gastrointestinal bleeding, heart failure, high blood pressure
            DMARDsMethotrexate, leflunomide, hydroxychloroquine, sulfasalazine, minocyclineGastrointestinal intolerance (nausea, stomatitis or diarrhea), hepatic toxicity, post-treatment fatigue, headache, dizziness, rheumatoid nodule formation
            GCsDexamethasone, prednisone, prednisolone, betamethasoneGastrointestinal bleeding, hyperglycemia/diabetes, osteoporosis, infection
            Biological agentsInfliximab, etanercept, adalimumab, golimumab, certolizumab pegol, tocilizumab, sarilumabInfection, malignancy, cardiovascular risk, immunogenicity, risk of exposure in pregnancy
            RNAimiRNA, siRNARapid enzymatic degradation in the blood, short half-life in serum, low cellular uptake

            DMARDs, disease-modifying anti-rheumatic drugs; GCs, glucocorticoids; miRNA, microRNA; NSAIDs, non-steroidal anti-inflammatory drugs; RNAi, RNA interference; siRNA, small interfering RNA.

            NSAIDs are suitable for treatment in early stages of RA; although they can rapidly alleviate symptoms and improve patients’ quality of life, they cannot prevent further joint damage [35]. NSAIDs also inhibit the production of prostaglandins by blocking the ability of cyclooxygenase-2 to transfer arachidonic acid to the endoperoxide pathway, thereby decreasing inflammation associated with RA [36]. Patients’ responses to NSAIDs greatly vary, particularly among older patients. Older patients with RA have an elevated risk of adverse reactions, because they tend to take other drugs in addition to NSAIDs. In addition, older patients tend to have poor compliance with NSAID regimens, as a result of physical dysfunction (such as visual impairment, arthritis, dementia or depression), thus resulting in poor curative effects [37]. Commonly used NSAIDs include acetylsalicylate (aspirin), naproxen, diclofenac, ibuprofen and etodolac. Their long-term use is associated with high risk of nausea, abdominal pain, ulcers, gastrointestinal bleeding, heart failure and high blood pressure [38].

            DMARDs, such as methotrexate (MTX), sulfasalazine, hydroxychloroquine, leflunomide and minocycline, can decrease joint swelling, pain and systemic inflammation [35]. MTX is a folic acid analogue that inhibits nucleotide synthesis and purine metabolism by inhibiting the activity of dihydrofolate reductase. This drug was originally used to treat hematological malignancies and has been shown to be effective at low doses in the treatment inflammatory arthritis [34, 39]. Because of its low cost and long-term safety, MTX is currently the most preferred DMARD for RA treatment [40]. However, it has been associated with gastrointestinal intolerance (nausea, stomatitis or diarrhea), hepatotoxicity, post-treatment fatigue, headache, dizziness and rheumatoid nodule formation [41].

            GCs, such as dexamethasone, prednisone, prednisolone and betamethasone, are strong anti-inflammatory and immunosuppressive drugs widely used to treat RA [42], because of their ability to control pain, stiffness and swelling. However, GCs are less effective in preventing disease progression [43], and their long-term use increases the risk of cardiovascular disease, gastrointestinal bleeding, hyperglycemia/diabetes, osteoporosis and infection [44, 45].

            Biological agents have recently emerged as a novel therapeutic approach for RA, and have been found to effectively alleviate symptoms, slow disease progression and prevent joint injury. Currently, 12 biological agents are used in clinical practice, including five TNF-α inhibitors (infliximab, etanercept, adalimumab, golimumab and certolizumab pegol); inhibitors of IL-6 and its receptor (tocilizumab and sarilumab); a CD80/86-CD inhibitor (abatacept); and an anti-CD20 antibody (rituximab) [34, 46]. However, the use of biological agents has been associated with elevated risk of infection, malignancy, cardiovascular injury, immunogenicity and other adverse events [47].

            MicroRNAs (miRNAs) and small interfering RNAs (siRNAs) are small RNA molecules that downregulate protein expression by triggering the degradation of messenger RNAs before their translation into proteins, through a process known as RNAi [48, 49]. In rat models of arthritis, miR-449 has been found to inhibit the production of the inflammatory factor IL-6, whereas miR-708-5p blocks inflammatory cell infiltration, synovial hyperplasia and cartilage destruction [50]. Although siRNA may show promise for therapeutic RNAi, siRNAs are rapidly degraded in the blood and are inefficiently internalized by cells, thus limiting their application in clinical settings [51, 52].

            The efficacy of these drugs may be improved through use in combination [53, 54]. Nevertheless, the systemic mechanisms of these drugs and their poor accumulation in inflamed joints require them to be administered often and at high doses, thus increasing the risk of adverse effects [55].

            5. TARGETED DRUG DELIVERY SYSTEMS

            5.1. Drug delivery systems used to treat RA in animal models

            To overcome the disadvantages of conventional RA drugs, several targeting nanodelivery systems are being developed [56, 57] to control drug release and prolong drug circulation in the blood [58] while decreasing systemic toxicity. Nanoparticles, liposomes and micelles are commonly used in medical applications for drug delivery, diagnostics and imaging [59], because of their good biodegradability and sustainability [60] ( Table 2 and Figure 2 ). These materials can stabilize drugs, control drug release and enhance drug accumulation at inflamed sites [61, 85].

            Table 2 |

            Drug-loaded nanoparticle drug delivery vehicles with the potential to treat RA.

            DrugCarrier typeBrief descriptionRoute of administrationAnimal model
            DexLiposomesProlonged half-lifeIntravenous injectionAIA rats [61]
            DexLiposomesEnhanced targeting effectIntravenous injectionAIA rats [62]
            Tofacitinib citrateLiposomesEnhanced distribution in inflamed sitesIntravenous injectionAIA rats [63]
            BerberineLiposomesDiminished inflammatory responseIntravenous injectionAIA rats [64]
            MTXGold NPsImproved efficacy and diminished toxicityIntravenous injectionAIA rats [65]
            IndomethacinMicellesEnhanced anti-inflammatory activityIntravenous injectionAIA rats [66]
            Dex/palmitateMicellesEnhanced targeting effectIntravenous injectionAIA rats [67]
            TacrolimusMicellesEnhanced targeting effectIntravenous injectionAIA rats [68]
            microRNA124/MTXMicellesEnhanced synergyIntravenous injectionAIA ats [69]
            Dex/p65 siRNAMicellesCo-delivery of siRNA and Dex into macrophagesIntravenous injectionCIA mice [70]
            Dex/palmitatePLGA NPsImproved pharmacokinetics and decreased cell damageIntravenous injectionCIA mice [71]
            siRNAPLGA NPsRetarded progression of inflammationIntravenous injectionCIA mice [72]
            β-SitosterolSolid lipid NPsInhibition of the NF-κB signaling pathwayIntravenous injectionAIA rats [73]
            PrednisoloneSolid lipid NPsSpecific binding to CD44 on inflammatory cellsIntravenous injectionCIA mice [74]
            EmbelinChitosan NPsProlonged retention time and improved targetingIntravenous injectionAIA rats [75]
            EugenolChitosan NPsEnhanced anti-arthritic activityIntravenous injectionCIA mice [76]
            Zinc gluconateChitosan NPsEnhanced targeting effectsIntravenous injectionCIA mice [77]
            CarvacrolAlbumin NPsPotent suppressive effectsIntravenous injectionAIA rats [78]
            CelastrolAlbumin NPsDecreased celastrol dose with effective therapyIntravenous injectionAIA rats [79]
            MethotrexateAlbumin NPsEnhanced retention and decreased carvacrol doseIntravenous injectionCIA mice [80]
            Prednisolone/curcuminAlbumin NPsCo-delivery of prednisolone and curcumin to inflamed sitesIntravenous injectionAIA rats [81]
            HydroxychloroquineBiomimetic NPsIncreased half-life and targeting to inflamed jointsIntravenous injectionCIA mice [82]
            DexInjectable hydrogelEnhanced therapeutic efficacyIntra-articular injectionCIA mice [83]
            Indomethacin/MTX/MMP-9 siRNAInjectable hydrogelSynergistic treatment with multiple drugsIntra-articular injectionAIA rats [84]

            AIA, adjuvant-induced arthritis; CIA, collagen-induced arthritis; NPs, nanoparticles; Dex, dexamethasone; MTX, methotrexate; PLGA, poly (lactic-co-glycolic acid).

            Figure 2 |

            Nanoparticles commonly used in targeted delivery systems.

            PLGA, poly(lactic-co-glycolic acid).

            5.1.1 Liposomes

            Liposomes consist of phospholipids and cholesterol, which form a lipid bilayer with an aqueous core. Their particle sizes are usually in the range of 25 nm to 2.5 μm [86]. Liposomes can encapsulate both hydrophobic and hydrophilic drugs, and they have good biocompatibility and biodegradability. However, the ability of liposomes to encapsulate hydrophobic drugs is not ideal, and drugs can easily leak out [87]. Although traditional liposomes are rapidly cleared by the reticuloendothelial system, modifying liposomes with polyethylene glycol (PEG) effectively decreases the adsorption of plasma proteins and subsequent clearance by the reticuloendothelial system, thus prolonging the circulation of drugs in the blood and improving their distribution in inflamed joints [88]. Intravenous administration of dexamethasone-loaded polymerized stealth liposomes to arthritic rats has been found to significantly prolong drug circulation in the blood and to enhance drug accumulation in inflamed joints [61]. This treatment significantly decreases the levels of TNF-α and IL-1β at lesion sites as well as the degree of joint swelling, thus indicating inhibition of RA progression.

            The peptide ART-2 (CKPFDRALC) shows preferential homing to arthritic joints of rats and strong binding to endothelial cells. To improve the targeting efficiency of liposomes to inflamed joints, one study has designed liposomes modified with ART-2 peptide. These dexamethasone-loaded liposomes with ART-2 modification accumulate in inflamed joints to a greater extent than dexamethasone-loaded liposomes without ART-2 modification, and relieve RA more efficiently [62]. Beyond hydrophobic drugs, liposomes can also be used to efficiently encapsulate hydrophilic drugs. Hydrophilic drugs can be encapsulated within the aqueous core of liposomes. Liposomes loaded with tofacitinib citrate, a water-soluble anti-inflammatory drug, have been found to be selectively internalized by inflammatory cells in a rat model of arthritis and to accumulate in arthritic paws [63]. This method of tofacitinib citrate delivery significantly improves its therapeutic efficacy, downregulates inflammatory cytokines in joint tissues and relieves RA symptoms. In another study, PEGylated liposomes loaded with water-soluble berberine have been found to accumulate selectively in inflamed joints of rats with adjuvant-induced arthritis (AIA). Berberine potently activates miR-23a, thus downregulating inflammatory kinases such as ASK1 and GSK-3β, as well as mediators of Wnt1 signaling, and ultimately mitigating bone erosion [64]. The safety and efficacy of liposomal bupivacaine has been confirmed in surgery; however, its pharmacokinetic parameters and safety in a Chinese population have not been evaluated. A phase I study has confirmed that liposomal bupivacain is well tolerated and safe among individuals of Chinese descent [89].

            5.1.2 Gold nanoparticles

            Gold nanoparticles (AuNPs), with particle sizes ranging from 1 to 100 nm, are widely used in diagnostics, therapy and biological imaging [90]. AuNPs have excellent stability and biocompatibility, customizable shapes and dimensions, easily functionalized surfaces, high drug loading capacity and low toxicity. However, AuNPs tend to accumulate in the kidneys, liver and spleen after entering the body, thus potentially leading to incomplete metabolism in the body [91]. AuNPs have strong affinity to thiol and amine groups, and therefore can bind targeting agents possessing these groups. In addition, AuNPs have good binding ability toward vascular endothelial growth factor and show natural antiangiogenic effects in inflamed synovium. Intra-articular injection of AuNPs of various dimensions into CIA mice has indicated clear antioxidant action, by significantly increasing catalase activity without causing any adverse effects on hematological indices. AuNPs of 50 nm, compared with 13 nm, have shown superior effects in inhibiting synovial angiogenesis, and have achieved better antioxidant and therapeutic effects in early stages of arthritis [92]. In another study, an MTX delivery system consisting of gold nanorods with a mesoporous silica shell (FAGMs) has been used for controlled release of MTX. The release rate of MTX from FAGMs in vitro markedly increases under 808 nm laser irradiation, thus achieving superior effects in inhibiting RA progression in AIA rats while decreasing the systemic toxicity of MTX, compared with free MTX [65]. These findings suggest that FAGMs may hold promise in the treatment of RA. In addition, clinical trials have tested the coupling of human proinsulin peptide (C19-A3) with AuNPs (C19-A3-AuNPs) for the treatment of type 1 diabetes. In a phase I clinical trial, the safety of intradermal administration of C19-A3-AuNP through microneedles has been explored [93]. Patients with type 1 diabetes have shown good tolerance to C19-A3-AuNPs with no signs of systemic allergy.

            5.1.3 Polymeric micelles

            Polymeric micelles are core-shell structures with a particle size of 10–100 nm that form through self-assembly of amphiphilic block copolymers in aqueous solutions [94]. Polymeric micelles can effectively avoid clearance by the kidneys and reticuloendothelial system, and they can target inflamed tissues through the ELVIS effect [95]. Micelles form when the polymer concentration in the solution exceeds the critical micellar concentration, and they dissociate into monomers when the polymer concentration is below the critical micellar concentration. Micelles with lower critical micellar concentrations are therefore more stable in circulation [96].

            Copolymeric micelles loaded with indomethacin, a non-steroidal anti-inflammatory drug that can effectively control inflammation [66], have been found to significantly relieve inflammatory symptoms and decrease the arthritic index and paw diameter in AIA rats. Similarly, micelles loaded with dexamethasone palmitate have been found to accumulate at inflamed sites and to decrease joint inflammation [67]. In another study, maltodextrin-α-tocopherol nano-micelles loaded with tacrolimus (TAC@MD-α-TOC) have been prepared for RA therapy. The micelles have been found to show stronger anti-rheumatic effects than the free drug. In vitro and in vivo experiments have demonstrated that TAC@MD-α-TOC is more effective than the free drug in promoting the viability of Vero cells and decreasing the levels of IL-6 and TNF-α in the serum and synovial fluid [68]. Monotherapy with MTX usually leads to irreversible joint injury because of its slow onset and long duration. MicroRNA-124 (miR-124) has shown direct bone protective effects against RA. Hybrid micelles containing both MTX conjugated polymer and miR-124 have been found to target and accumulate in the inflamed joints of AIA rats, and effectively enhance the synergistic effects of MTX and miR-124 [69]. Wang et al. have used polymeric micelles to co-deliver Dex and siRNA against p65, a member of the NF-κB family, to arthritis sites to treat RA [70]. These co-loaded micelles have been found to efficiently inhibit NF-κB signaling in macrophages in CIA mice and to cause activated macrophages in arthritic synovial membranes to revert to an anti-inflammatory state. In the treatment of RA, micelles co-loaded with Dex and p65 siRNA have shown better efficiency than free Dex or naked siRNA. In addition, a phase II study on the efficacy and safety of docetaxel-PM for treating recurrent or metastatic squamous cell carcinoma of the head and neck was conducted in 2015, and this docetaxel-PM is now on the market.

            5.1.4 PLGA nanoparticles

            Poly(lactic-co-glycolic acid) (PLGA) is a non-toxic, biodegradable polymer often used as a drug delivery carrier. PLGA nanoparticles can regulate drug release, and their surfaces can be easily modified. They are commonly used to protect biological agents, such as proteins and nucleic acids, against rapid metabolism and clearance in vivo [97]. The drug release from PLGA particles depends on the molecular weight of the polymer. Polymers with higher molecular weights have longer polymer chains, thus resulting in longer degradation times and slower drug release rates [98]. For example, loading of apremilast, a small-molecule drug designed for oral delivery, into PLGA nanoparticles significantly prolongs its half-life and mean residence time in vivo [99]. Luteinium-177, a radionuclide with a half-life of 6.71 d, can be used for radiation treatment of joints with advanced arthritis, owing to its β maximum emission energy (0.497 MeV, 78%). Hyaluronic acid (HA) is a polymer that specifically binds CD44 overexpressed on inflamed synovial cells. In one study, PLGA modified with Luteinium-177 and HA has been used for encapsulating MTX (177Lu-DOTA-HA-PLGA(MTX)). The 177Lu-DOTA-HA-PLGA(MTX) nanoparticles strongly bind, and are efficiently internalized by, inflamed synovial cells [100]. In addition, encapsulating dexamethasone palmitate in PLGA-PEG nanoparticles has been found to improve the drug’s pharmacokinetic profile, and decrease its tendency to damage cells and aggregate in the liver, kidneys and lungs [71]. Bruton’s tyrosine kinase (BTK) in macrophages and B cells is an important target in RA therapy. However, high dosages of BTK inhibitors are needed for effective BTK inhibition, thus limiting their clinical application. Zhao et al. have developed cationic lipid-assisted PEG-b-PLGA nanoparticles (CLAN) loaded with BTK siRNA (CLANsiBTK) [72]. In the CIA mouse model, CLANsiBTK has been found to significantly alleviate arthritis symptoms, downregulate the expression of inflammatory cytokines (TNF-α, IL-1β and IFN-γ) and decrease damage to the paw joints.

            5.1.5 Solid lipid nanoparticles (SLNs)

            SLNs, a lipid-based drug delivery system with a particle size of 10–1000 nm, specifically target inflamed tissues and enable controlled drug release [101]. SLNs show low immunogenicity in the human body and can easily infiltrate biological tissues while carrying large amounts of lipophilic compounds. However, the ability of SLNs to encapsulate hydrophilic drugs is poor. Moreover, changes in the drug release curve, polymorphic transformation and particle aggregation occur during storage [102]. Zhang et al. have loaded β-sitosterol into SLNs to improve its water solubility and bioavailability in AIA rats. The developed nanoparticles (β-sitosterol-SLNs) have been found to exhibit good anti-arthritic effects by inhibiting NF-κB and activating the heme oxygenase-1/NF-erythroid 2-associated factor 2 pathway [73]. To improve the targeting of inflamed tissue, prednisolone-loaded SLNs coated with HA have been prepared; these SLNs have been found to accumulate in the inflamed joints of CIA mice and to decrease joint swelling, bone erosion and levels of inflammatory cytokines in the serum [74].

            5.1.6 Chitosan nanoparticles

            Chitosan, a polysaccharide that arises through the deacetylation of chitin, is widely used in the preparation of microparticles and nanoparticles [103]. Chitosan nanoparticles have good biodegradability, low immunogenicity and high cell permeability, and therefore are suitable nanocarriers for targeted drug delivery [104]. For example, chitosan nanoparticles can improve the efficacy of the anti-inflammatory drug embelin, which is poorly absorbed in the body, and is rapidly metabolized and cleared. Loading embelin into chitosan nanoparticles downregulates malondialdehyde and nitroxide, as well as TNF-α, IL-6 and IL-1β in the serum in AIA rats, while decreasing oxidative stress [75]. In a study in CIA rats, eugenol loading into chitosan nanoparticles has been found to significantly improve the drug’s ability to decrease the expression of monocyte chemoattractant protein-1 and transforming growth factor-β, and to alleviate joint synovial hyperplasia and cartilage injury [76]. Similarly, loading zinc gluconate into chitosan nanoparticles has been found to improve the compound’s ability to inhibit the infiltration of inflammatory cells in ankle joints; downregulate TNF-α, IL-6 and inducible nitric oxide synthase; and upregulate SOD1 [77].

            5.1.7 Albumin nanoparticles

            Albumin is the most abundant protein in the plasma, accounting for approximately 60% of the total protein in the blood [105]. It is considered an ideal candidate for drug delivery because of its strong ability to bind hydrophobic and hydrophilic drugs, relatively long half-life in the blood (19 days), biodegradability and lack of immunogenicity [106]. Albumin, which has a molecular weight of 66.5 kDa, can be obtained from various sources, such as egg white (ovalbumin), bovine serum (BSA) or human serum (HSA) [107]. The nanoparticles can prolong the circulation in the blood by adsorbing albumin [108]. For example, coating albumin on the surfaces of liposomes or embedding it directly in phospholipids of liposomes enables the nanostructures to evade phagocytosis, thus prolonging their circulation [109]. In a study in AIA rats, loading carvacrol into BSA nanoparticles has been found to significantly improve the anti-inflammatory agent’s ability to mitigate swelling and decrease release of the inflammatory cytokines NO and IL-17 in arthritic rats [78]. To improve the targeting of BSA nanoparticles, Gong et al. have prepared palmitic acid modified BSA nanoparticles (PAB NPs) and loaded them with celastrol. The PAB NPs efficiently bind scavenger receptor-A (SR-A) and elicit 9–10 times more macrophages than normal BSA NPs. PAB NPs have been found to deliver the anti-inflammatory drug celastrol to inflamed tissues more effectively than BSA NPs, and to alleviate RA symptoms at lower doses [79]. Loading MTX into HSA nanoparticles labeled with chlorin e6 has been found to increase the drug’s accumulation and retention in inflamed joints of CIA rats; the encapsulated drug slows RA progression as effectively as a 50% higher dose of free drug [80]. In a study in AIA rats, treatment with prednisolone and curcumin loaded into HSA nanoparticles has led to lower levels of pro-inflammatory cytokines in activated macrophages, higher levels of the anti-inflammatory cytokine IL-10, greater drug accumulation in inflamed joints and stronger therapeutic efficacy than nanoparticles loaded with single drugs or free drugs alone or a mixture of two drugs [81]. Abraxane (paclitaxel combined with albumin) was approved by the FDA in 2005. In 2021, the FDA approved Fyarro (sirolimus albumin-bound nanoparticles, nab-sirolimus, ABI-009) to be marketed for intravenous infusion in the treatment of locally advanced unresectable or metastatic malignant perivascular epithelioid cell tumors.

            All the above studies used exogenous albumin to prepare nanoparticles in vitro. However, after nanoparticles enter the body, they can cause immunogenic reactions. Therefore, some researchers have attempted to use the endogenous albumin in the body to prepare albumin-bound nanoparticles for targeting therapy. The albumin-binding domain (ABD, 46 amino acids) has been reported to have strong affinity for serum albumin. The ABD035 variant has shown superior affinity toward albumin from various sources (such as human, mouse, rat, cow and monkey albumin). Zhang et al. have designed a redox-responsive paclitaxel micelle system modified by the ABD035 peptide [110]. After intravenous injection, the micellar system has been found to combine with endogenous albumin in blood, then deliver paclitaxel to tumor cells via gp60 and SPARC receptors. Inflammatory tissue has similar characteristics to tumor tissue, such as abundant albumin receptors and incomplete vascular structure. Therefore, developing nanoparticles that bind endogenous albumin may serve as a favorable strategy for targeting RA.

            5.1.8 Biomimetic nanoparticles

            Biomimetic nanoparticles have attracted particular attention in recent years because of their ability to evade clearance by the reticuloendothelial system [111]. Biomembranes are usually extracted through the following method. Cells are placed in hypotonic liquid for cell disruption and lysis, then purified and collected by discontinuous gradient centrifugation at 4 °C. Protease inhibitor is added throughout the entire extraction process to protect protein activity [112]. Nanoplatforms are prepared by encapsulating nanoparticles into cell membranes extracted from red blood cells, macrophages, neutrophils or platelets [113] through co-extrusion, extrusion/sonication, freeze-thaw/sonication, extrusion/sonication and other methods [114, 115]. These nanoplatforms have shown excellent biocompatibility, effective drug delivery, prolonged circulation times in the blood and minimal adverse immune responses [116, 117]. For example, spherical, prolate-spheroidal and oblate-spheroidal PLGA nanoparticles coated with erythrocyte membrane effectively evade clearance by macrophages [118]. Encapsulation of hydroxychloroquine-loaded nanoparticles in membranes from umbilical vein endothelial cells expressing TNF-associated apoptosis-inducing ligand has generated nanoparticles for the delivery of hydroxychloroquine to inflamed joints in CIA mice while also inducing M1 macrophage apoptosis by upregulating death receptor-5 [82], thus effectively inhibiting RA progression. In CIA mice, coating nanoparticles with neutrophil membranes and then administering these nanostructures together with immunoregulatory molecules has been found to promote tissue repair, downregulate pro-inflammatory cytokines and inhibit synovitis [119].

            5.1.9 Injectable hydrogels

            Hydrogels are highly hydrated three-dimensional networks of hydrophilic polymers, whose bionic structure is similar to that of the extracellular matrix of natural biological tissues and has good biocompatibility. Injectable hydrogels are more comfortable, and elicit less pain and fewer adverse effects, than non-injectable hydrogels. Favorable injectable hydrogels cannot easily be obtained by using a single material. Incorporating nanofiller into a polymer matrix can achieve desirable injectable gels with high biocompatibility and biodegradability, more easily modifiable properties, and better ability to deliver hydrophilic or hydrophobic macromolecules in a sustained manner. Injectable hydrogels with loosely interconnected polymer chains have shown higher burst release and more rapid drug clearance [120]. Wang et al. have developed a temperature-sensitive hydrogel (DLTH) based on chitosan-glycerol-borax for intra-articular delivery of dexamethasone [83]. In the CIA mouse model, intra-articular injection of DLTH loaded with dexamethasone has shown good anti-inflammatory and analgesic effects. Similarly, in situ hydrogels have also been designed to co-deliver indomethacin, methotrexate and MMP-9 siRNA for synergistic and comprehensive treatment of RA [84]. These in situ hydrogels significantly down-regulate the expression of inflammatory factors (TNF-α, IL-6) and MMP-9 in plasma and knee joint fluid after intra-articular injection.

            5.2. Targeted RA treatment in animal models
            5.2.1 Passive targeting

            Synovial thickening during the development of RA induces hypoxia and angiogenesis, thus leading to vascular leakage at sites of inflammation. Because the ELVIS effect in RA is similar to the EPR effect in tumor tissues [21], as exploited by many cancer therapies, passive targeting strategies against RA have been developed on the basis of the ELVIS effect [121]. Nanoparticles of 20–200 nm can penetrate the synovial tissue through the intercellular space between endothelial cells and accumulate at inflamed sites, thus significantly limiting drug adverse effects ( Figure 3 ). In contrast, nanoparticles smaller than 10 nm are easily cleared by renal filtration, whereas those larger than 200 nm are removed by phagocytes in the reticuloendothelial system [122, 123].

            Figure 3 |

            Schematic illustration of the extravasation through leaky vasculature and subsequent inflammatory cell-mediated sequestration (ELVIS) effect.

            To improve the passive targeting ability of nanoparticles of suitable size, their surfaces have been modified with PEG, which shields surface charge, inhibits the adsorption of serum proteins and helps evade detection by circulating macrophages [124, 125]. For example, allowing hydrophilic short-chain methoxy PEG to self-assemble with gambogic acid, a natural drug that inhibits inflammation by downregulating IL-1β and TNF-α, has been found to increase the drug’s ability to ameliorate paw inflammation in CIA mice [126]. The free drug, in contrast, shows low water solubility, poor pharmacokinetics and hemolytic toxicity. Similarly, modifying liposomes 100 nm in diameter with slightly negative charge with 10% PEG5000 has been found to increase their in vivo circulation time and ability to target inflamed joints in CIA mice, thus resulting in much better effects of encapsulated dexamethasone than the free drug [127]. Similarly, liposomes prepared with hydrogenated phosphatidylinositol can avoid clearance by the reticuloendothelial system and prolong the blood circulation time. Doxorubicin loaded in hydrogenated phosphatidylinositol liposomes has been detected within 72 hours after injection. The circulation time of hydrogenated phosphatidylinositol liposomes is significantly longer than that of ordinary liposomes [128]. In addition, nanoparticles coated with poly(ethylene oxide)-block-poly (γ-methylprednisolone) have nearly neutral surface charges, can avoid the non-specific binding of macromolecules to nanoparticles, and additionally can escape clearance by reticuloendothelial systems [129].

            5.2.2. Active targeting

            Active targeting primarily involves modifying the nanocarrier surface with appropriate ligands that bind receptors expressed on the surfaces of target cells at inflamed sites ( Table 3 ) [140]. RA involves vascular regeneration and inflammation, and these processes involve B cells, T cells, macrophages and other immune cells [11]. These processes are driven by growth factors, pro-inflammatory cytokines, chemokines, cell adhesion molecules, proteases and the hypoxia-vascular endothelial growth factor angiopoietin. Activated macrophages are abundant at sites of inflammation, where they produce large amounts of pro-inflammatory IL-6, IL-1β and TNF-α [141]. The surfaces of activated macrophages contain abundant folate receptor-β, vasoactive intestinal peptide receptor, scavenger receptor class A, TLRs, transforming growth factor-β receptor, CD44, CD64 and other receptors ( Figure 4 ), whereas the surfaces of endothelial cells contain abundant αvβ3-integrin, E-selectin, vascular cell adhesion molecule-1 and intercellular adhesion molecule-1 [142, 143]. If drug-loaded nanoparticles can selectively bind these receptors, they can target their cargo to inflamed tissue, provided that the nanoparticles can evade clearance by the reticuloendothelial system and penetrate the inflamed vascular endothelium ( Figure 5 ) [55].

            Table 3 |

            Drug-loaded nanoparticles targeting receptors highly expressed in rheumatoid arthritis tissue.

            Targeted receptorDrugCarrierAnimal model
            Folate receptorMTXChitosan nanoparticlesCIA mice [130]
            Folate receptorMTXLiposomesCIA mice [131]
            Folate receptorMyeloid cell leukemia-1 siRNA/DexMicellesCIA mice [132]
            Scavenger receptorMTXDouble layered hydroxide nanocompositesAIA rats [133]
            Scavenger receptorCelastrolMicellesAIA rats [134]
            CD44DexPolymeric nanoparticlesAIA rats [135]
            CD44MTX/teriflunomideHydroxyapatite nanoparticlesCIA mice [136]
            CD44TripterineBilosomesCAIA rats [137]
            Mannose receptor p-Coumaric acidLiposomesAIA rats [138]
            Mannose receptorMorinLiposomesAIA rats [139]

            AIA, adjuvant-induced arthritis; CAIA, collagen antibody-induced arthritis; CIA, collagen-induced arthritis; MTX, methotrexate; Dex, dexamethasone.

            Figure 4 |

            Active targeting of cells in inflamed joints through specific binding between ligand-modified nanoparticles and targeted receptors.

            Figure 5 |

            Active targeting of inflamed tissues by (1) drug-loaded nanoparticles modified with specific ligands that can evade (2) reticuloendothelial clearance and penetrate (3) inflamed vascular endothelium.

            5.2.2.1 Folate receptors

            Folate receptors are glycopolypeptides with high affinity for folic acid that have four isoforms (FR-α, FR-β, FR-γ and FR-δ) that are differentially expressed among tissues [144]. FR-β is widely expressed on the surfaces of activated synovial macrophages in patients with RA and therefore may be a potential target for disease treatment [145]. MTX encapsulated in hydrophobically modified ethylene glycol chitosan nanoparticles conjugated to folic acid, compared with nanoparticles without folic acid conjugation, has been found to result in significantly lower pro-inflammatory cytokine expression, paw thickness and arthritic scores in CIA mice [130]. In another study, folic acid has been coupled with PEG100 monostearate to prepare liposomes co-loaded with MTX and catalase. The obtained liposomes show enhanced cellular uptake through folate-mediated endocytosis and strong toxicity against activated RAW264.7 cells, as well as enhanced accumulation in inflamed joints and therapeutic efficacy in CIA mice, while causing minimal toxicity to major organs [131]. Myeloid cell leukemia-1 (Mcl-1) is overexpressed in the macrophages of arthritic joints. Inhibiting MCL-1 induces apoptosis of macrophages and thus alleviates inflammation. Li et al. have developed folate-modified polymeric micelles co-loaded with MCL-1 siRNA and DEX for RA therapy [132]. In the CIA mouse model, these DEX/siRNA co-loaded polymeric micelles have been found to significantly decrease MCL-1 mRNA levels in macrophages, as well as levels of inflammatory factors such as TNF-α and IL-1β.

            5.2.2.2 Scavenger receptors

            Scavenger receptors belong to a superfamily of structurally heterogeneous proteins, including transmembrane proteins and soluble secretory extracellular domains, and effectively regulate the uptake of oxidized and acetylated low-density lipoprotein [146]. Scavenger receptor class A (SR-A) is expressed mainly on the surfaces of mature macrophages and has been associated with atherogenesis [147]. This receptor has been targeted for RA treatment: MTX-loaded layered double hydroxide nanocomposites have been modified with dextran sulfate (LDH-MTX-DS), a hydrophilic block that specifically binds SR-A [148]. LDH-MTX-DS releases MTX more rapidly at pH 5.5 than at pH 7.4, and actively targets scavenger receptors on the surfaces of activated RAW 264.7 cells, thus leading to significantly stronger therapeutic effects than those of free MTX in AIA rats [133]. In another study, micelles modified with dextran sulfate have also been prepared for celastrol delivery. These micelles have been found to effectively accumulate and release celastrol in RAW264.7 cells by binding SR-A, and to significantly improve the therapeutic effect of celastrol in vivo without causing clear systemic toxicity [134].

            5.2.2.3 CD44

            CD44 is a non-kinase transmembrane glycoprotein that is overexpressed in inflammatory synovial macrophages and fibroblasts [149]. It promotes pathological angiogenesis by regulating the proliferation, migration and adhesion of endothelial cells. HA is a natural polysaccharide in the extracellular matrix that specifically binds the CD44 receptor [54, 150]. Therefore, HA-coated acid-sensitive polymeric nanoparticles composed of egg phosphatidylcholine, polyethyleneimine and poly(cyclohexane-1,4-diyl acetone dimethylene ketal) have been loaded with dexamethasone to target activated macrophages overexpressing CD44 [135]. The obtained nanoparticles have been found to significantly decrease inflammatory cell infiltration as well as damage to bone and cartilage in the ankles of AIA rats, showing stronger therapeutic efficacy than the free drug [135]. In another study, HA-functionalized hydroxyapatite nanoparticles loaded with MTX and teriflunomide have shown high cellular uptake and cytotoxicity in vitro; prevented the progression of arthritis and promoted joint regeneration in CIA rats; and led to less hepatotoxicity than commercially available formulations [136]. Bilosomes loaded with tripterine, a natural compound with strong antioxidant, antiangiogenic and antirheumatic properties, have been prepared with cationic lipids through the thin film hydration method, then coated with HA to form HA@Tri-BLs [137]. The resulting nanosystem has been found to improve the circulation time of tripterine in vivo, and increase its systemic and intra-arthritic bioavailability, in a collagen antibody-induced arthritis model. The in vivo anti-arthritic efficacy of HA@Tri-BLs has also been found to be significantly higher than that of uncoated bilosomes, owing to high drug accumulation in the articular cavity.

            5.2.2.4 Mannose receptors

            Mannose receptors belong to the C-type lectin family and are overexpressed on the surfaces of macrophages during inflammation; they mediate endocytosis of glycoproteins [151]. Mannose-conjugated liposomes loaded with p-coumaric acid, a compound with anti-inflammatory and osteoclastic effects, prolongs the drug’s residence time in joints and downregulates pro-inflammatory TNF-α, IL-1β, IL-6 and IL-23, as well as the transcription factor NFATc1, which triggers osteoclast differentiation [138]. In another study, mannose-modified liposomes have been used to encapsulate morin, a bioflavonoid with anti-inflammatory, antitumor and anti-oxidant activity. These mannose-modified liposomes are preferentially internalized by macrophages from arthritic rats, and inhibit the inflammatory immune response and osteoclastogenesis better than the reference drug dexamethasone palmitate encapsulated in mannosylated liposomes, according to clinical and histological analysis [139].

            5.3. Targeting based on environmental stimuli

            The microenvironment of inflamed tissues differs from that of healthy tissue in pH, redox conditions, and the levels of oxygen, reactive oxygen species (ROS), glucose, enzymes and ATP. These differences can be exploited to trigger drug release from responsive delivery systems, thereby avoiding premature release into the circulation or healthy tissues, and decreasing the risk of adverse or other off-target effects [152154].

            5.3.1 pH-responsive drug delivery

            Under normal conditions, the pH of the extracellular medium and blood is usually ∼7.4, but the pH falls to 6.0 in the synovial microenvironment [155]. This pH difference between the inflammatory microenvironment and normal tissues has been exploited to design pH-responsive targeted drug delivery systems based either on ionizable polymers (polyacids or polybases), whose conformation or solubility changes with pH, or on polymer systems with acid-sensitive bonds [155, 156]. For example, pH-sensitive polymeric micelles prepared through self-assembly of PEG-based derivatives and the hydrophobic drug prednisolone through acid-labile hydrazone bonds have been found to promote the accumulation of prednisolone in inflamed joints and to show stronger anti-inflammatory effects in vivo than the free drug [157]. Another study has reported the preparation of pH-sensitive polymer nanoparticles that deliver siRNA when the surrounding pH is 5.0, thus allowing selective drug delivery to sites of inflammation, and leading to therapeutic efficacy in AIA rats [52]. A small library of biocompatible amphiphilic polymers based on methoxy poly(ethylene glycol)-poly(cyclohexane-1,4-diyl acetone dimethyleneketal) and methoxy poly(ethylene glycol)-poly((cyclohexane86.7%, 1,5-pentane-diol13.3%)-1,4-diyl acetone dimethylene ketal) has been synthesized for the targeted delivery of superoxide dismutase. The novel polymers release the enzyme in a pH-dependent manner, thus allowing the enzyme, which is normally rapidly cleared or degraded in vivo, to show good anti-oxidant and anti-inflammatory activities in AIA rats [158].

            5.3.2 Redox-responsive drug delivery

            Activated T cells produce 10–100 times more ROS at sites of inflammation than in normal tissues. These ROS trigger an anti-oxidant glutathione (GSH) response, which prevents further increases in ROS and resultant cellular damage [159]. The increased GSH concentrations in inflammatory microenvironments result in cleavage of disulfide bonds [160, 161]. Reduction-responsive polyprodrug amphiphiles have been prepared through “reversible addition fragmentation chain transfer” polymerization of indomethacin-based redox-responsive prodrug monomers bearing disulfide bonds that are GSH responsive and phenylboronic acid ester bonds that are H2O2 responsive [162]. The resulting polymers have been found to efficiently antagonize the effects of lipopolysaccharide on RAW264.7 macrophages.

            5.3.3 ROS-responsive drug delivery

            ROS, such as H2O2, superoxide (O2 • −), hydroxyl radical (OH), peroxynitrite (ONOO) and hypochlorite (OCl), play important roles in normal cellular signaling pathways and oxidative metabolism. However, their excessive production in cells or tissues can cause oxidative stress, thus leading to various diseases, including inflammation, cancer and atherosclerosis [154, 163, 164]. To target the high ROS levels in RA tissues, 4-phenylboronic acid pinacol ester-conjugated cyclodextrin biomaterials have been used to prepare ROS-responsive dexamethasone-loaded nanoparticles [165]. The particles have been found to be efficiently internalized by activated macrophages, to accumulate in inflamed joints, and to decrease joint swelling and cartilage destruction in CIA mice.

            5.3.4 Enzyme-responsive drug delivery

            Enzymes such as proteases, glycosidases, metalloproteases, lipases and phospholipases are biocatalysts that play key roles in countless normal processes [166], but their expression may be altered in disease [167]. Such changes can be exploited as triggers for environmentally responsive drug delivery [168]. For example, phospholipase A2 is overexpressed under inflammatory conditions, and it specifically hydrolyzes sn-2 ester bonds in phospholipids [169]. Thus, this enzyme attacks liposomes of the appropriate composition and consequently enables controlled release of drug cargo [170], Similarly, high phospholipase A2 levels can release colchicine from phosphatidylcholinase-responsive liposomes that are otherwise stable in the blood circulation [171].

            5.4 Local injection strategies

            Local intra-articular injection is also an important therapeutic strategy for RA. Intra-articular injection has several advantages in RA therapy, including good bioavailability, diminished systemic exposure and adverse events, and lower costs. The intra-articular injection of corticosteroids is the first-line treatment for RA. Researchers have found that intra-articular injection of etanercept is a safer and more promising treatment than corticosteroid treatment [172]. Drugs currently used for intra-articular injection are dissolved in solution. Consequently, the drug rapidly diffuses into the systemic circulation after intra-articular injection, thereby resulting in rapid removal from the joint cavity and short retention time. Frequent intra-articular injection is required for good therapeutic effects against RA, thus increasing the risk of local pain, joint swelling and infection [173]. Therefore, several nanoplatforms have been developed to improve drug accumulation at inflamed sites. For example, injection of tetramethylpyrazine as a nanosuspension with hydrophobic ions prolongs its retention in the articular cavity in rats and leads to greater anti-arthritic efficacy than that of the free drug [174]. Moreover siRNAs targeting TNF loaded into lipid-polymer hybrid nanoparticles of lipidoid and PLGA have been found to inhibit inflammation in murine experimental arthritis models, even at the low siRNA dose of 1 μg [175]. Intra-articular injection of HA-modified liposomes loaded with diclofenac and dexamethasone has been observed to effectively decrease inflammation and paw swelling in mice for 4 weeks [176].

            6. SUMMARY AND PERSPECTIVES

            Although great progress has been made in treating RA, the clinical application of traditional therapies is limited by high costs and requirements for frequent, long-term dosing. Nanocarriers can actively or passively deliver drugs to inflamed sites, thereby prolonging drug half-life, improving drug accumulation in target tissues and decreasing drug systemic toxicity. Although large numbers of nanocarriers have been developed for RA therapy, few have entered clinical trials. These nanocarriers face several challenges. (1) Most nanocarriers are unstable in the circulation and usually leak the drug before reaching inflamed sites. (2) Nanocarriers are easily captured by the reticuloendothelial system after intravenous injection. (3) The targeting activity of nanocarriers for inflamed regions is generally too low. Consequently, developing safer, more efficient nanocarriers remains the most important research frontier at present.

            In addition, although nanocarriers can greatly improve therapeutic effects in RA, several shortcomings remain. For example, the modification of ligands or targeting molecules is complex and costly. How to simplify modification procedures and decrease costs will be key for clinical transformation. PEG modification is usually used to prolong the blood circulation time in RA targeting by nanocarriers. However, this modification can also hinder interactions between nanoparticles and cells, and inhibit cellular uptake. In addition, after multiple intravenous injections, the PEG-modified nanoparticles are rapidly removed from the blood circulation [177]. Therefore, developing new simply and easily prepared biological materials that can avoid phagocytosis by the reticuloendothelial system, prolong the circulation time and target arthritis sites will be critical for RA therapy.

            Recent research has focused on the construction of targeting nanocarriers modified with ligands that bind specific receptors on the surfaces of cells involved in RA. Most of these nanocarriers target RA by targeting the receptors on vascular endothelial cells, fibroblast-like synoviocytes and macrophages. Beyond these cells, autoreactive T cells and B cells also substantially contribute to the inflammatory process. Activated T cells activate monocytes, macrophages and synovial fibroblasts. Developing nanocarriers targeting these cells will also be important.

            A recent approach for targeted treatment of RA involves delivery of RA-associated antibodies and antigens to auto-reactive lymphocytes or dendritic cells, with the aim of inducing antigen-specific immune tolerance [178] while maintaining protective immunity [179, 180]. To this end, future work should explore as many RA-associated antibodies and antigens as possible. For example, RA-associated cell injury and death result in the release of cell-free DNA (cfDNA) into the peripheral blood and synovial fluid [181, 182]. Auto-antibodies bind the cfDNA, and the resulting complex activates TLRs, thereby leading to the secretion of inflammatory cytokines [183]. Thus, using cationic polymers to eliminate cfDNA may be a promising treatment against RA. In a step in this direction, one study has shown that cationic dimethylamino-group-modified polydopamine nanoparticles strongly bind cfDNA and efficiently inhibit cfDNA-induced inflammation in an animal model [184]. Decreasing the systemic toxicity of these cationic polymeric nanoparticles is an important priority for future work.

            RA-associated disability and mortality currently affect large numbers of people. Despite decades of exploration of nanocarriers for RA treatment, many areas have yet to be investigated. Developing superior nanocarriers for RA therapy is important and urgently needed.

            ACKNOWLEDGMENTS

            This work was financially supported by the National Natural Science Foundation of China (82073803 and 82274372) and the Applied Basic Research Programs of Sichuan Province Science and Technology Department (2019YJ0663).

            CONFLICTS OF INTEREST

            The authors declare that they have no conflicts of interest.

            REFERENCES

            1. Scott DL, Wolfe F, Huizinga TW. Rheumatoid Arthritis. Lancet. 2010. Vol. 376:1094–1108

            2. Cush J. Rheumatoid Arthritis: Early Diagnosis and Treatment. The Medical Clinics of North America. 2021. Vol. 105:355–365

            3. Deane KD, Demoruelle MK, Kelmenson LB, Kuhn KA, Norris JM, Holers VM. Genetic and Environmental Risk Factors for Rheumatoid Arthritis. Best Practice & Research. Clinical Rheumatology. 2017. Vol. 31:3–18

            4. Firestein GS. Evolving Concepts of Rheumatoid Arthritis. Nature. 2003. Vol. 423:356–361

            5. Ahlmén M, Svensson B, Albertsson K, Forslind K, Hafström I. BARFOT Study Group: Influence of Gender on Assessments of Disease Activity and Function in Early Rheumatoid Arthritis in Relation to Radiographic Joint Damage. Annals of the Rheumatic Disease. 2010. Vol. 69:230–233

            6. Zhang J, Jiang L, Sun L, Wang P, Sun S, Xu M, et al.. Targeted Drug Delivery Strategies for the Treatment of Rheumatoid Arthritis. Science China. Life Sciences. 2021. Vol. 64:1187–1189

            7. Qamar N, Arif A, Bhatti A, John P. Nanomedicine: An Emerging Era of Theranostics and Therapeutics for Rheumatoid Arthritis. Rheumatology (Oxford, England). 2019. Vol. 58:1715–1721

            8. Chen M, Daddy JCKA, Xiao Y, Ping Q, Zong L. Advanced Nanomedicine for Rheumatoid Arthritis Treatment: Focus on Active Targeting. Expert Opinion on Drug Delivery. 2017. Vol. 14:1141–1144

            9. Chemin K, Gerstner C, Malmström V. Effector Functions of CD4+ T Cells at the Site of Local Autoimmune Inflammation-Lessons From Rheumatoid Arthritis. Frontiers in Immunology. 2019. Vol. 10:353

            10. Wang Q, Sun X. Recent Advances in Nanomedicines for the Treatment of Rheumatoid Arthritis. Biomaterials Science. 2017. Vol. 5:1407–1420

            11. Jang S, Kwon EJ, Lee JJ. Rheumatoid Arthritis: Pathogenic Roles of Diverse Immune Cells. International Journal of Molecular Sciences. 2022. Vol. 23:905

            12. Alivernini S, MacDonald L, Elmesmari A, Finlay S, Tolusso B, Gigante MR, et al.. Distinct Synovial Tissue Macrophage Subsets Regulate Inflammation and Remission in Rheumatoid Arthritis. Nature Medicine. 2020. Vol. 26:1295–1306

            13. Ishiguro N, Moriyama M, Furusho K, Furukawa S, Shibata T, Murakami Y, et al.. Activated M2 Macrophages Contribute to the Pathogenesis of IgG4-Related Disease via Toll-Like Receptor 7/Interleukin-33 Signaling. Arthritis & Rheumatology (Hoboken, NJ). 2020. Vol. 72:166–178

            14. Jiang C, Zhu W, Xu J, Wang B, Hou W, Zhang R, et al.. MicroRNA-26a Negatively Regulates Toll-Like Receptor 3 Expression of Rat Macrophages and Ameliorates Pristane Induced Arthritis in Rats. Arthritis Research & Therapy. 2014. Vol. 16:R9

            15. Wehr P, Purvis H, Law SC, Thomas R. Dendritic Cells, T Cells and their Interaction in Rheumatoid Arthritis. Clinical and Experimental Immunology. 2019. Vol. 196:12–27

            16. O’Neil LJ, Kaplan MJ. Neutrophils in Rheumatoid Arthritis: Breaking Immune Tolerance and Fueling Disease. Trends in Molecular Medicine. 2019. Vol. 25:215–227

            17. Xu H, Chen F, Liu T, Xu J, Li J, Jiang L, et al.. Ellagic Acid Blocks RANKL-RANK Interaction and Suppresses RANKL-Induced Osteoclastogenesis by Inhibiting RANK Signaling Pathways. Chemico-Biological Interactions. 2020. Vol. 331:109235

            18. Fang Q, Zhou C, Nandakumar KS. Molecular and Cellular Pathways Contributing to Joint Damage in Rheumatoid Arthritis. Mediators of Inflammation. 2020. Vol. 2020:3830212

            19. So T. The immunological Significance of Tumor Necrosis Factor Receptor-Associated Factors (TRAFs). International Immunology. 2022. Vol. 34:7–20

            20. Hassine HB, Zemni R, Nacef IB, Boumiza A, Slama F, Baccouche K, et al.. A TRAF6 Genetic Variant is Associated with Low Bone Mineral Density in Rheumatoid Arthritis. Clinical Rheumatology. 2019. Vol. 38:1067–1074

            21. Koziolová E, Venclíková K, Etrych T. Polymer-Drug Conjugates in Inflammation Treatment. Physiological Research. 2018. Vol. 67 Suppl 2:S281–S292

            22. Yu S, Lu Y, Zong M, Tan Q, Fan L. Hypoxia-Induced miR-191-C/EBPβ Signaling Regulates Cell Proliferation and Apoptosis of Fibroblast-Like Synoviocytes from Patients with Rheumatoid Arthritis. Arthritis Research & Therapy. 2019. Vol. 21:78

            23. Wang Y, Wu H, Deng R. Angiogenesis as a Potential Treatment Strategy for Rheumatoid Arthritis. European Journal of Pharmacology. 2021. Vol. 910:174500

            24. O’Brien MJ, Shu Q, Stinson WA, Tsou PS, Ruth JH, Isozaki T, et al.. A Unique Role for Galectin-9 in Angiogenesis and Inflammatory Arthritis. Arthritis Research & Therapy. 2018. Vol. 20:31

            25. Jaiswal PK, Goel A, Mittal RD. Survivin: A Molecular Biomarker in Cancer. Indian Journal of Medical Research. 2015. Vol. 141:389–397

            26. Ma S, Wang J, Lin J, Jin S, He F, Mei J, et al.. Survivin Promotes Rheumatoid Arthritis Fibroblast-Like Synoviocyte Cell Proliferation, and the Expression of Angiogenesis-Related Proteins by Activating the NOTCH Pathway. International Journal of Rheumatic Diseases. 2021. Vol. 24:922–929

            27. Avouac J, Pezet S, Vandebeuque E, Orvain C, Gonzalez V, Marin G, et al.. Semaphorins: From Angiogenesis to Inflammation in Rheumatoid Arthritis. Arthritis & Rheumatology (Hoboken, NJ). 2021. Vol. 73:1579–1588

            28. Riemann A, Wußling H, Loppnow H, Fu H, Reime S, Thews O. Acidosis Differently Modulates the Inflammatory Program in Monocytes and Macrophages. Biochimica et Biophysica Acta. 2016. Vol. 1862:72–81

            29. Carstensen SMD, Terslev L, Jensen MP, Østergaard M. Future use of Musculoskeletal Ultrasonography and Magnetic Resonance Imaging in Rheumatoid Arthritis. Current Opinion in Rheumatology. 2020. Vol. 32:264–272

            30. van Delft MAM, Huizinga TWJ. An Overview of Autoantibodies in Rheumatoid Arthritis. Journal of Autoimmunity. 2020. Vol. 110:102392

            31. Sieghart D, Platzer A, Studenic P, Alasti F, Grundhuber M, Swiniarski S, et al.. Determination of Autoantibody Isotypes Increases the Sensitivity of Serodiagnostics in Rheumatoid Arthritis. Frontiers in Immunology. 2018. Vol. 9:876

            32. Mun S, Lee J, Park M, Shin J, Lim MK, Kang HG. Serum Biomarker Panel for the Diagnosis of Rheumatoid Arthritis. Arthritis Research & Therapy. 2021. Vol. 23:31

            33. Radu AF, Bungau SG. Management of Rheumatoid Arthritis: An Overview. Cells. 2021. Vol. 10:2857

            34. Lin YJ, Anzaghe M, Schülke S. Update on the Pathomechanism, Diagnosis, and Treatment Options for Rheumatoid Arthritis. Cells. 2020. Vol. 9:880

            35. Littlejohn EA, Monrad SU. Early Diagnosis and Treatment of Rheumatoid Arthritis. Primary Care. 2018. Vol. 45:237–255

            36. Bindu S, Mazumder S, Bandyopadhyay U. Non-Steroidal Anti-Inflammatory Drugs (NSAIDs) and Organ Damage: A Current Perspective. Biochemical Pharmacology. 2020. Vol. 180:114147

            37. Johnson AG, Quinn DI, Day RO. Non-Steroidal Anti-Inflammatory Drugs. Medical Journal of Australia. 1995. Vol. 163:155–158

            38. Schjerning AM, McGettigan P, Gislason G. Cardiovascular Effects and Safety of (Non-Aspirin) NSAIDs. Nature Reviews. Cardiology. 2020. Vol. 17:574–584

            39. Ling S, Bluett J, Barton A. Prediction of Response to Methotrexate in Rheumatoid Arthritis. Expert Review of Clinical Immunology. 2018. Vol. 14:419–429

            40. Wang W, Zhou H, Liu L. Side Effects of Methotrexate Therapy for Rheumatoid Arthritis: A Systematic Review. European Journal of Medicinal Chemistry. 2018. Vol. 158:502–516

            41. Salliot C, van der Heijde D. Long-Term Safety of Methotrexate Monotherapy in Patients with Rheumatoid Arthritis: A Systematic Literature Research. Annals of the Rheumatic Disease. 2009. Vol. 68:1100–1104

            42. Sinniah A, Yazid S, Flower RJ. From NSAIDs to Glucocorticoids and Beyond. Cells. 2021. Vol. 10:3524

            43. Sparks JA. Rheumatoid Arthritis. Annals of Internal Medicine. 2019. Vol. 170:ITC1–ITC16

            44. Hua C, Buttgereit F, Combe B. Glucocorticoids in Rheumatoid Arthritis: Current Status and Future Studies. RMD Open. 2020. Vol. 6:e000536

            45. Strehl C, van der Goes MC, Bijlsma JW, Jacobs JW, Buttgereit F. Glucocorticoid-Targeted Therapies for the Treatment of Rheumatoid Arthritis. Expert Opinion on Investigational Drugs. 2017. Vol. 26:187–195

            46. Law ST, Taylor PC. Role of Biological Agents in Treatment of Rheumatoid Arthritis. Pharmacological Research. 2019. Vol. 150:104497

            47. Cañete JD, Hernández MV, Sanmartí R. Safety Profile of Biological Therapies for Treating Rheumatoid Arthritis. Expert Opinion on Biological Therapy. 2017. Vol. 17:1089–1103

            48. Zavvar M, Assadiasl S, Soleimanifar N, Pakdel FD, Abdolmohammadi K, Fatahi Y, et al.. Gene Therapy in Rheumatoid Arthritis: Strategies to Select Therapeutic Genes. Journal of Cellular Physiology. 2019. Vol. 234:16913–16924

            49. Feng N, Guo F. Nanoparticle-siRNA: A Potential Strategy for Rheumatoid Arthritis Therapy? Journal of Controlled Release. 2020. Vol. 325:380–393

            50. Chang C, Xu L, Zhang R, Jin Y, Jiang P, Wei K, et al.. MicroRNA-Mediated Epigenetic Regulation of Rheumatoid Arthritis Susceptibility and Pathogenesis. Frontiers in Immunology. 2022. Vol. 13:838884

            51. Song P, Yang C, Thomsen JS, Dagnæs-Hansen F, Jakobsen M, Brüel A, et al.. Lipidoid-siRNA Nanoparticle-Mediated IL-1β Gene Silencing for Systemic Arthritis Therapy in a Mouse Model. Molecular Therapy. 2019. Vol. 27:1424–1435

            52. Sun X, Dong S, Li X, Yu K, Sun F, Lee RJ, et al.. Delivery of siRNA using Folate Receptor-Targeted pH-Sensitive Polymeric Nanoparticles for Rheumatoid Arthritis Therapy. Nanomedicine. 2019. Vol. 20:102017

            53. Shrestha S, Zhao J, Yang C, Zhang J. Iguratimod Combination Therapy Compared with Methotrexate Monotherapy for the Treatment of Rheumatoid Arthritis: A Systematic Review and Meta-Analysis. Clinical Rheumatology. 2021. Vol. 40:4007–4017

            54. Tang Q, Yin D, Wang Y, Du W, Qin Y, Ding A, et al.. Cancer Stem Cells and Combination Therapies to Eradicate Them. Current Pharmaceutical Design. 2020. Vol. 26:1994–2008

            55. Xiao S, Tang Y, Lv Z, Lin Y, Chen L. Nanomedicine - Advantages for their use in Rheumatoid Arthritis Theranostics. Journal of Controlled Release. 2019. Vol. 316:302–316

            56. Su T, Feng X, Yang J, Xu W, Liu T, Zhang M, et al.. Polymer Nanotherapeutics to Correct Autoimmunity. Journal of Controlled Release. 2022. Vol. 343:152–174

            57. Feng X, Xu W, Li Z, Song W, Ding J, Chen X. Immunomodulatory Nanosystems. Advanced Science (Weinheim, Baden-wurttemberg, Germany). 2019. Vol. 6:1900101

            58. Li H, Xu Y, Tong Y, Dan Y, Zhou T, He J, et al.. Sucrose Acetate Isobutyrate as an In Situ Forming Implant for Sustained Release of Local Anesthetics. Current Drug Delivery. 2019. Vol. 16:331–340

            59. Zheng C, Li M, Ding J. Challenges and Opportunities of Nanomedicines in Clinical Translation. BIO Integration. 2021. Vol. 2:57–60

            60. Wu D, Tang L, Zeng Z, Zhang J, Hu X, Pan Q, et al.. Delivery of Hyperoside by using a Soybean Protein Isolated-Soy Soluble Polysaccharide Nanocomplex: Fabrication, Characterization, and In Vitro Release Properties. Food Chemistry. 2022. Vol. 386:132837

            61. Wang Q, He L, Fan D, Liang W, Fang J. Improving the Anti-Inflammatory Efficacy of Dexamethasone in the Treatment of Rheumatoid Arthritis with Polymerized Stealth Liposomes as a Delivery Vehicle. Journal of Materials Chemistry. B. 2020. Vol. 8:1841–1851

            62. Meka RR, Venkatesha SH, Acharya B, Moudgil KD. Peptide-Targeted Liposomal Delivery of Dexamethasone for Arthritis Therapy. Nanomedicine (London, England). 2019. Vol. 14:1455–1469

            63. Shen Q, Shu H, Xu X, Shu G, Du Y, Ying X. Tofacitinib Citrate-Based Liposomes for Effective Treatment of Rheumatoid Arthritis. Die Pharmazie. 2020. Vol. 75:131–135

            64. Sujitha S, Dinesh P, Rasool M. Berberine Encapsulated PEG-Coated Liposomes Attenuate Wnt1/β-catenin Signaling in Rheumatoid Arthritis via miR-23a Activation. European Journal of Pharmaceutics and Biopharmaceutics. 2020. Vol. 149:170–191

            65. Li X, Hou Y, Meng X, Li G, Xu F, Teng L, et al.. Folate Receptor-Targeting Mesoporous Silica-Coated Gold Nanorod Nanoparticles for the Synergistic Photothermal Therapy and Chemotherapy of Rheumatoid Arthritis. RSC Advances. 2021. Vol. 11:3567–3574

            66. Abdollahi AR, Firouzian F, Haddadi R, Nourian A. Indomethacin Loaded Dextran Stearate Polymeric Micelles Improve Adjuvant-Induced Arthritis in Rats: Design and In Vivo Evaluation. Inflammopharmacology. 2021. Vol. 29:107–121

            67. Wang X, Feng Y, Fu J, Wu C, He B, Zhang H, et al.. A Lipid Micellar System Loaded with Dexamethasone Palmitate Alleviates Rheumatoid Arthritis. AAPS PharmSciTech. 2019. Vol. 20:316

            68. Helal HM, Samy WM, Kamoun EA, El-Fakharany EM, Abdelmonsif DA, Aly RG, et al.. Potential Privilege of Maltodextrin-α-Tocopherol Nano-Micelles in Seizing Tacrolimus Renal Toxicity, Managing Rheumatoid Arthritis and Accelerating Bone Regeneration. International Journal of Nanomedicine. 2021. Vol. 16:4781–4803

            69. Hao F, Lee RJ, Zhong L, Dong S, Yang C, Teng L, et al.. Hybrid Micelles Containing Methotrexate-Conjugated Polymer and Co-Loaded with microRNA-124 for Rheumatoid Arthritis Therapy. Theranostics. 2019. Vol. 9:5282–5297

            70. Wang Q, Jiang H, Li Y, Chen W, Li H, Peng K, et al.. Targeting NF-kB Signaling with Polymeric Hybrid Micelles that Co-Deliver siRNA and Dexamethasone for Arthritis Therapy. Biomaterials. 2017. Vol. 122:10–22

            71. Simón-Vázquez R, Tsapis N, Lorscheider M, Rodríguez A, Calleja P, Mousnier L, et al.. Improving Dexamethasone Drug Loading and Efficacy in Treating Arthritis through a Lipophilic Prodrug Entrapped into PLGA-PEG Nanoparticles. Drug Delivery and Translational Research. 2022. Vol. 12:1270–1284

            72. Zhao G, Liu A, Zhang Y, Zuo ZQ, Cao ZT, Zhang HB, et al.. Nanoparticle-Delivered siRNA Targeting Bruton’s Tyrosine Kinase for Rheumatoid Arthritis Therapy. Biomaterials Science. 2019. Vol. 7:4698–4707

            73. Zhang F, Liu Z, He X, Li Z, Shi B, Cai F. β-Sitosterol-Loaded Solid Lipid Nanoparticles Ameliorate Complete Freund’s Adjuvant-Induced Arthritis in Rats: Involvement of NF-κB and HO-1/Nrf-2 Pathway. Drug Delivery. 2020. Vol. 27:1329–1341

            74. Zhou M, Hou J, Zhong Z, Hao N, Lin Y, Li C. Targeted Delivery of Hyaluronic Acid-Coated Solid Lipid Nanoparticles for Rheumatoid Arthritis Therapy. Drug Delivery. 2018. Vol. 25:716–722

            75. Cui P, Qu F, Sreeharsha N, Sharma S, Mishra A, Gubbiyappa SK. Antiarthritic Effect of Chitosan Nanoparticle Loaded with Embelin Against Adjuvant-Induced Arthritis in Wistar Rats. IUBMB Life. 2020. Vol. 72:1054–1064

            76. Jabbari N, Eftekhari Z, Roodbari NH, Parivar K. Evaluation of Encapsulated Eugenol by Chitosan Nanoparticles on the Aggressive Model of Rheumatoid Arthritis. International Immunopharmacology. 2020. Vol. 85:106554

            77. Ansari MM, Ahmad A, Mishra RK, Raza SS, Khan R. Zinc Gluconate-Loaded Chitosan Nanoparticles Reduce Severity of Collagen-Induced Arthritis in Wistar Rats. ACS Biomaterials Science & Engineering. 2019. Vol. 5:3380–3397

            78. Gholijani N, Abolmaali SS, Kalantar K, Ravanrooy MH. Therapeutic Effect of Carvacrol-loaded Albumin Nanoparticles on Arthritic Rats. Iranian Journal of Pharmaceutical Research. 2020. Vol. 19:312–320

            79. Gong T, Tan T, Zhang P, Li H, Deng C, Huang Y, et al.. Palmitic Acid-Modified Bovine Serum Albumin Nanoparticles Target Scavenger Receptor-A on Activated Macrophages to Treat Rheumatoid Arthritis. Biomaterials. 2020. Vol. 258:120296

            80. Liu L, Hu F, Wang H, Wu X, Eltahan AS, Stanford S, et al.. Secreted Protein Acidic and Rich in Cysteine Mediated Biomimetic Delivery of Methotrexate by Albumin-Based Nanomedicines for Rheumatoid Arthritis Therapy. ACS Nano. 2019. Vol. 13:5036–5048

            81. Yan F, Li H, Zhong Z, Zhou M, Lin Y, Tang C, et al.. Co-Delivery of Prednisolone and Curcumin in Human Serum Albumin Nanoparticles for Effective Treatment of Rheumatoid Arthritis. International Journal of Nanomedicine. 2019. Vol. 14:9113–9125

            82. Shi Y, Xie F, Rao P, Qian H, Chen R, Chen H, et al.. TRAIL-Expressing Cell Membrane Nanovesicles as an Anti-Inflammatory Platform for Rheumatoid Arthritis Therapy. Journal of Controlled Release. 2020. Vol. 320:304–313

            83. Wang QS, Xu BX, Fan KJ, Li YW, Wu J, Wang TY. Dexamethasone-Loaded Thermosensitive Hydrogel Suppresses Inflammation and Pain in Collagen-Induced Arthritis Rats. Drug Design, Development and Therapy. 2020. Vol. 14:4101–4113

            84. Yin N, Tan X, Liu H, He F, Ding N, Gou J, et al.. A Novel Indomethacin/Methotrexate/MMP-9 siRNA In Situ Hydrogel with Dual Effects of Anti-Inflammatory Activity and Reversal of Cartilage Disruption for the Synergistic Treatment of Rheumatoid Arthritis. Nanoscale. 2020. Vol. 12:8546–8562

            85. Li H, Li Y, Wang X, Hou Y, Hong X, Gong T, et al.. Rational Design of Polymeric Hybrid Micelles to Overcome Lymphatic and Intracellular Delivery Barriers in Cancer Immunotherapy. Theranostics. 2017. Vol. 7:4383–4398

            86. Ahmed KS, Hussein SA, Ali AH, Korma SA, Lipeng Q, Jinghua C. Liposome: Composition, Characterisation, Preparation, and Recent Innovation in Clinical Applications. Journal of Drug Targeting. 2019. Vol. 27:742–761

            87. Li H, Tang Q, Wang Y, Li M, Wang Y, Zhu H, et al.. Injectable Thermosensitive Lipo-Hydrogels Loaded with Ropivacaine for Prolonging Local Anesthesia. International Journal of Pharmaceutics. 2022. Vol. 611:121291

            88. Guimarães D, Cavaco-Paulo A, Nogueira E. Design of Liposomes as Drug Delivery System for Therapeutic Applications. International Journal of Pharmaceutics. 2021. Vol. 601:120571

            89. Cheung BM, Ng PY, Liu Y, Zhou M, Yu V, Yang J, et al.. Pharmacokinetics and Safety of Liposomal Bupivacaine after Local Infiltration in Healthy Chinese Adults: A Phase 1 Study. BMC Anesthesiology. 2021. Vol. 21:197

            90. Pirmardvand Chegini S, Varshosaz J, Taymouri S. Recent Approaches for Targeted Drug Delivery in Rheumatoid Arthritis Diagnosis and Treatment. Artificial Cells, Nanomedicine, and Biotechnology. 2018. Vol. 46:502–514

            91. Boisselier E, Astruc D. Gold Nanoparticles in Nanomedicine: Preparations, Imaging, Diagnostics, Therapies and Toxicity. Chemical Society Reviews. 2009. Vol. 38:1759–1782

            92. Kirdaite G, Leonaviciene L, Bradunaite R, Vasiliauskas A, Rudys R, Ramanaviciene A, et al.. Antioxidant Effects of Gold Nanoparticles on Early Stage of Collagen-Induced Arthritis in Rats. Research in Veterinary Science. 2019. Vol. 124:32–37

            93. Tatovic D, McAteer MA, Barry J, Barrientos A, Rodríguez Terradillos K, Perera I, et al.. Safety of the use of Gold Nanoparticles Conjugated with Proinsulin Peptide and Administered by Hollow Microneedles as an Immunotherapy in Type 1 Diabetes. Immunotherapy Advances. 2022. Vol. 2:ltac002

            94. Xiang J, Shen Y, Zhang Y, Liu X, Zhou Q, Zhou Z, et al.. Multipotent Poly(Tertiary Amine-Oxide) Micelles for Efficient Cancer Drug Delivery. Advanced Science (Weinheim, Baden-wurttemberg, Germany). 2022. Vol. 9:e2200173

            95. Ghezzi M, Pescina S, Padula C, Santi P, Del Favero E, Cantù L, et al.. Polymeric Micelles in Drug Delivery: An Insight of the Techniques for their Characterization and Assessment in Biorelevant Conditions. Journal of Controlled Release. 2021. Vol. 332:312–336

            96. Ghosh B, Biswas S. Polymeric Micelles in Cancer Therapy: State of the Art. Journal of Controlled Release. 2021. Vol. 332:127–147

            97. Mohammadi-Samani S, Taghipour B. PLGA Micro and Nanoparticles in Delivery of Peptides and Proteins; Problems and Approaches. Pharmaceutical Development and Technology. 2015. Vol. 20:385–393

            98. Ding D, Zhu Q. Recent Advances of PLGA Micro/Nanoparticles for the Delivery of Biomacromolecular Therapeutics. Materials Science & Engineering. C, Materials for Biological Applications. 2018. Vol. 92:1041–1060

            99. Anwer MK, Mohammad M, Ezzeldin E, Fatima F, Alalaiwe A, Iqbal M. Preparation of Sustained Release Apremilast-Loaded PLGA Nanoparticles: In Vitro Characterization and In Vivo Pharmacokinetic Study in Rats. International Journal of Nanomedicine. 2019. Vol. 14:1587–1595

            100. Trujillo-Nolasco RM, Morales-Avila E, Ocampo-García BE, Ferro-Flores G, Gibbens-Bandala BV, Escudero-Castellanos A, et al.. Preparation and In Vitro Evaluation of Radiolabeled HA-PLGA Nanoparticles as Novel MTX Delivery System for Local Treatment of Rheumatoid Arthritis. Materials Science & Engineering. C, Materials for Biological Applications. 2019. Vol. 103:109766

            101. Duan Y, Dhar A, Patel C, Khimani M, Neogi S, Sharma P, et al.. A Brief Review on Solid Lipid Nanoparticles: Part and Parcel of Contemporary Drug Delivery Systems. RSC Advances. 2020. Vol. 10:26777–26791

            102. Pandey S, Shaikh F, Gupta A, Tripathi P, Yadav JS. A Recent Update: Solid Lipid Nanoparticles for Effective Drug Delivery. Advanced Pharmaceutical Bulletin. 2022. Vol. 12:17–33

            103. Ahsan SM, Thomas M, Reddy KK, Sooraparaju SG, Asthana A, Bhatnagar I. Chitosan as Biomaterial in Drug Delivery and Tissue Engineering. International Journal of Biological Macromolecules. 2018. Vol. 110:97–109

            104. Rizeq BR, Younes NN, Rasool K, Nasrallah GK. Synthesis, Bioapplications, and Toxicity Evaluation of Chitosan-Based Nanoparticles. International Journal of Molecular Sciences. 2019. Vol. 20:5776

            105. Tan YL, Ho HK. Navigating Albumin-based Nanoparticles through Various Drug Delivery Routes. Drug Discovery Today. 2018. Vol. 23:1108–1114

            106. Kianfar E. Protein Nanoparticles in Drug Delivery: Animal Protein, Plant Proteins and Protein Cages, Albumin Nanoparticles. Journal of Nanobiotechnology. 2021. Vol. 19:159

            107. Copolovici DM, Langel K, Eriste E, Langel Ü. Cell-Penetrating Peptides: Design, Synthesis, and Applications. ACS Nano. 2014. Vol. 8:1972–1994

            108. Li H, Wang Y, Tang Q, Yin D, Tang C, He E, et al.. The Protein Corona and its Effects on Nanoparticle-based Drug Delivery Systems. Acta Biomaterialia. 2021. Vol. 129:57–72

            109. Wang J, Ding Y, Zhou W. Albumin Self-Modified Liposomes for Hepatic Fibrosis Therapy via SPARC-Dependent Pathways. International Journal of Pharmaceutics. 2020. Vol. 574:118940

            110. Zhang Y, Guo Z, Cao Z, Zhou W, Zhang Y, Chen Q, et al.. Endogenous Albumin-Mediated Delivery of Redox-Responsive Paclitaxel-Loaded Micelles for Targeted Cancer Therapy. Biomaterials. 2018. Vol. 183:243–257

            111. Wei Q, Su Y, Xin H, Zhang L, Ding J, Chen X. Immunologically Effective Biomaterials. ACS Applied Materials & Interfaces. 2021. Vol. 13:56719–56724

            112. Chen HY, Deng J, Wang Y, Wu CQ, Li X, Dai HW. Hybrid Cell Membrane-Coated Nanoparticles: A Multifunctional Biomimetic Platform for Cancer Diagnosis and Therapy. Acta Biomaterialia. 2020. Vol. 112:1–13

            113. Kunder SS, Wairkar S. Platelet Membrane Camouflaged Nanoparticles: Biomimetic Architecture for Targeted Therapy. International Journal of Pharmaceutics. 2021. Vol. 598:120395

            114. Zhang R, Wu S, Ding Q, Fan Q, Dai Y, Guo S, et al.. Recent Advances in Cell Membrane-Camouflaged Nanoparticles for Inflammation Therapy. Drug Delivery. 2021. Vol. 28:1109–1119

            115. Oroojalian F, Beygi M, Baradaran B, Mokhtarzadeh A, Shahbazi MA. Immune Cell Membrane-Coated Biomimetic Nanoparticles for Targeted Cancer Therapy. Small (Weinheim an der Bergstrasse, Germany). 2021. Vol. 17:e2006484

            116. Chen L, Hong W, Ren W, Xu T, Qian Z, He Z. Recent Progress in Targeted Delivery Vectors based on Biomimetic Nanoparticles. Signal Transduction and Targeted Therapy. 2021. Vol. 6:225

            117. Beh CY, Prajnamitra RP, Chen LL, Hsieh PC. Advances in Biomimetic Nanoparticles for Targeted Cancer Therapy and Diagnosis. Molecules. 2021. Vol. 26:5052

            118. Ben-Akiva E, Meyer RA, Yu H, Smith JT, Pardoll DM, Green JJ. Biomimetic Anisotropic Polymeric Nanoparticles Coated with Red Blood Cell Membranes for Enhanced Circulation and Toxin Removal. Science Advances. 2020. Vol. 6:eaay9035

            119. Zhang Q, Dehaini D, Zhang Y, Zhou J, Chen X, Zhang L, et al.. Neutrophil Membrane-Coated Nanoparticles Inhibit Synovial Inflammation and Alleviate Joint Damage in Inflammatory Arthritis. Nature Nanotechnology. 2018. Vol. 13:1182–1190

            120. Bisht R, Nirmal S, Agrawal R, Jain GK, Nirmal J. Injectable In-Situ Gel Depot System for Targeted Delivery of Biologics to the Retina. Journal of Drug Targeting. 2021. Vol. 29:46–59

            121. Durymanov M, Kamaletdinova T, Lehmann SE, Reineke J. Exploiting Passive Nanomedicine Accumulation at Sites of Enhanced Vascular Permeability for Non-Cancerous Applications. Journal of Controlled Release. 2017. Vol. 261:10–22

            122. Mitragotri S, Yoo JW. Designing Micro- and Nano-Particles for Treating Rheumatoid Arthritis. Archives of Pharmacal Research. 2011. Vol. 34:1887–1897

            123. Katsuki S, Matoba T, Koga JI, Nakano K, Egashira K. Anti-Inflammatory Nanomedicine for Cardiovascular Disease. Frontiers in Cardiovascular Medicine. 2017. Vol. 4:87

            124. Mozar FS, Chowdhury EH. Impact of PEGylated Nanoparticles on Tumor Targeted Drug Delivery. Current Pharmaceutical Design. 2018. Vol. 24:3283–3296

            125. Yadav D, Dewangan HK. PEGYLATION: An Important Approach for Novel Drug Delivery System. Journal of Biomaterials Science. Polymer Edition. 2021. Vol. 32:266–280

            126. Nguyen A, Ando H, Böttger R, DurgaRao Viswanadham KK, Rouhollahi E, Ishida T, et al.. Utilization of Click Chemistry to Study the Effect of Poly(ethylene)glycol Molecular Weight on the Self-Assembly of PEGylated Gambogic Acid Nanoparticles for the Treatment of Rheumatoid Arthritis. Biomaterials Science. 2020. Vol. 8:4626–4637

            127. Ren H, He Y, Liang J, Cheng Z, Zhang M, Zhu Y, et al.. Role of Liposome Size, Surface Charge, and PEGylation on Rheumatoid Arthritis Targeting Therapy. ACS Applied Materials & Interfaces. 2019. Vol. 11:20304–20315

            128. Gabizon A, Shiota R, Papahadjopoulos D. Pharmacokinetics and Tissue Distribution of Doxorubicin Encapsulated in Stable Liposomes with Long Circulation Times. Journal of the National Cancer Institute. 1989. Vol. 81:1484–1488

            129. Chen H, Wang L, Yeh J, Wu X, Cao Z, Wang YA, et al.. Reducing Non-Specific Binding and Uptake of Nanoparticles and Improving Cell Targeting with an Antifouling PEO-b-PgammaMPS Copolymer Coating. Biomaterials. 2010. Vol. 31:5397–5407

            130. Wu Z, Xu K, Min J, Chen M, Shen L, Xu J, et al.. Folate-Conjugated Hydrophobicity Modified Glycol Chitosan Nanoparticles for Targeted Delivery of Methotrexate in Rheumatoid Arthritis. Journal of Applied Biomaterials & Functional Materials. 2020. Vol. 18:2280800020962629

            131. Chen M, Amerigos JCKD, Su Z, Guissi NEI, Xiao Y, Zong L, et al.. Folate Receptor-Targeting and Reactive Oxygen Species-Responsive Liposomal Formulation of Methotrexate for Treatment of Rheumatoid Arthritis. Pharmaceutics. 2019. Vol. 11:582

            132. Li Y, Wei S, Sun Y, Zong S, Sui Y. Nanomedicine-based Combination of Dexamethasone Palmitate and MCL-1 siRNA for Synergistic Therapeutic Efficacy against Rheumatoid Arthritis. Drug Delivery and Translational Research. 2021. Vol. 11:2520–2529

            133. Wang X, Yang B, Xu X, Su M, Xi M, Yin Z. Dextran Sulfate-Modified pH-Sensitive Layered Double Hydroxide Nanocomposites for Treatment of Rheumatoid Arthritis. Drug Delivery and Translational Research. 2021. Vol. 11:1096–1106

            134. Yu C, Liu H, Guo C, Chen Q, Su Y, Guo H, et al.. Dextran Sulfate-based MMP-2 Enzyme-Sensitive SR-A Receptor Targeting Nanomicelles for the Treatment of Rheumatoid Arthritis. Drug Delivery. 2022. Vol. 29:454–465

            135. Yu C, Li X, Hou Y, Meng X, Wang D, Liu J, et al.. Hyaluronic Acid Coated Acid-Sensitive Nanoparticles for Targeted Therapy of Adjuvant-Induced Arthritis in Rats. Molecules. 2019. Vol. 24:146

            136. Pandey S, Rai N, Mahtab A, Mittal D, Ahmad FJ, Sandal N, et al.. Hyaluronate-Functionalized Hydroxyapatite Nanoparticles Laden with Methotrexate and Teriflunomide for the Treatment of Rheumatoid Arthritis. International Journal of Biological Macromolecules. 2021. Vol. 171:502–513

            137. Yang H, Liu Z, Song Y, Hu C. Hyaluronic Acid-Functionalized Bilosomes for Targeted Delivery of Tripterine to Inflamed Area with Enhancive Therapy on Arthritis. Drug Delivery. 2019. Vol. 26:820–830

            138. Neog MK, Rasool M. Targeted Delivery of p-coumaric Acid Encapsulated Mannosylated Liposomes to the Synovial Macrophages Inhibits Osteoclast Formation and Bone Resorption in the Rheumatoid Arthritis Animal Model. European Journal of Pharmaceutics and Biopharmaceutics. 2018. Vol. 133:162–175

            139. Sultana F, Neog MK, Rasool M. Targeted Delivery of Morin, a Dietary Bioflavanol Encapsulated Mannosylated Liposomes to the Macrophages of Adjuvant-Induced Arthritis Rats Inhibits Inflammatory Immune Response and Osteoclastogenesis. European Journal of Pharmaceutics and Biopharmaceutics. 2017. Vol. 115:229–242

            140. Wang Q, Qin X, Fang J, Sun X. Nanomedicines for the Treatment of Rheumatoid Arthritis: State of art and Potential Therapeutic Strategies. Acta Pharmaceutica Sinica. B. 2021. Vol. 11:1158–1174

            141. Kondo N, Kuroda T, Kobayashi D. Cytokine Networks in the Pathogenesis of Rheumatoid Arthritis. International Journal of Molecular Sciences. 2021. Vol. 22:10922

            142. Kim HJ, Lee SM, Park KH, Mun CH, Park YB, Yoo KH. Drug-Loaded Gold/Iron/Gold Plasmonic Nanoparticles for Magnetic Targeted Chemo-Photothermal Treatment of Rheumatoid Arthritis. Biomaterials. 2015. Vol. 61:95–102

            143. Xiao Q, Li X, Li Y, Wu Z, Xu C, Chen Z, et al.. Biological Drug and Drug Delivery-Mediated Immunotherapy. Acta Pharmaceutica Sinica. B. 2021. Vol. 11:941–960

            144. Kumar P, Huo P, Liu B. Formulation Strategies for Folate-Targeted Liposomes and Their Biomedical Applications. Pharmaceutics. 2019. Vol. 11:381

            145. Steinz MM, Ezdoglian A, Khodadust F, Molthoff CFM, Srinivasarao M, Low PS, et al.. Folate Receptor Beta for Macrophage Imaging in Rheumatoid Arthritis. Frontiers in Immunology. 2022. Vol. 13:819163

            146. Alquraini A, El Khoury J. Scavenger Receptors. Current Biology. 2020. Vol. 30:R790–R795

            147. Ahmed M, Baumgartner R, Aldi S, Dusart P, Hedin U, Gustafsson B, et al.. Human Serum Albumin-Based Probes for Molecular Targeting of Macrophage Scavenger Receptors. International Journal of Nanomedicine. 2019. Vol. 14:3723–3741

            148. Nishinaka T, Mori S, Yamazaki Y, Niwa A, Wake H, Yoshino T, et al.. A Comparative Study of Sulphated Polysaccharide Effects on Advanced Glycation End-Product Uptake and Scavenger Receptor Class A Level in Macrophages. Diabetes & Vascular Disease Research. 2020. Vol. 17:1479164119896975

            149. Gorantla S, Gorantla G, Saha RN, Singhvi G. CD44 Receptor-Targeted Novel Drug Delivery Strategies for Rheumatoid Arthritis Therapy. Expert Opinion on Drug Delivery. 2021. Vol. 18:1553–1557

            150. Choi KY, Han HS, Lee ES, Shin JM, Almquist BD, Lee DS, et al.. Hyaluronic Acid-Based Activatable Nanomaterials for Stimuli-Responsive Imaging and Therapeutics: Beyond CD44-Mediated Drug Delivery. Advanced Materials (Deerfield Beach, Fla). 2019. Vol. 31:e1803549

            151. van der Zande HJP, Nitsche D, Schlautmann L, Guigas B, Burgdorf S. The Mannose Receptor: From Endocytic Receptor and Biomarker to Regulator of (Meta)Inflammation. Frontiers in Immunology. 2021. Vol. 12:765034

            152. Qiao Y, Wan J, Zhou L, Ma W, Yang Y, Luo W, et al.. Stimuli-Responsive Nanotherapeutics for Precision Drug Delivery and Cancer Therapy. Wiley Interdisciplinary Reviews. Nanomedicine and Nanobiotechnology. 2019. Vol. 11:e1527

            153. Wells CM, Harris M, Choi L, Murali VP, Guerra FD, Jennings JA. Stimuli-Responsive Drug Release from Smart Polymers. Journal of Functional Biomaterials. 2019. Vol. 10:34

            154. Gulfam M, Sahle FF, Lowe TL. Design Strategies for Chemical-Stimuli-Responsive Programmable Nanotherapeutics. Drug Discovery Today. 2019. Vol. 24:129–147

            155. Webb BA, Chimenti M, Jacobson MP, Barber DL. Dysregulated pH: A Perfect Storm for Cancer Progression. Nature Reviews, Cancer. 2011. Vol. 11:671–677

            156. El-Sawy HS, Al-Abd AM, Ahmed TA, El-Say KM, Torchilin VP. Stimuli-Responsive Nano-Architecture Drug-Delivery Systems to Solid Tumor Micromilieu: Past, Present, and Future Perspectives. ACS Nano. 2018. Vol. 12:10636–10664

            157. Li C, Li H, Wang Q, Zhou M, Li M, Gong T, et al.. pH-Sensitive Polymeric Micelles For Targeted Delivery To Inflamed Joints. Journal of Controlled Release. 2017. Vol. 246:133–141

            158. Sun X, Yu K, Zhou Y, Dong S, Hu W, Sun Y, et al.. Self-Assembled pH-Sensitive Polymeric Nanoparticles for the Inflammation-Targeted Delivery of Cu/Zn-Superoxide Dismutase. ACS Applied Materials & Interfaces. 2021. Vol. 13:18152–18164

            159. Das SS, Bharadwaj P, Bilal M, Barani M, Rahdar A, Taboada P, et al.. Stimuli-Responsive Polymeric Nanocarriers for Drug Delivery, Imaging, and Theragnosis. Polymers (Basel). 2020. Vol. 12:1397

            160. Hsu PH, Almutairi A. Recent Progress of Redox-Responsive Polymeric Nanomaterials for Controlled Release. Journal of Materials Chemistry. B. 2021. Vol. 9:2179–2188

            161. Chen J, Yang J, Ding H. Rational Construction of Polycystine-based Nanoparticles for Biomedical Applications. Journal of Materials Chemistry. B. 2022. Vol. 10:7173–7182

            162. Tan J, Deng Z, Liu G, Hu J, Liu S. Anti-Inflammatory Polymersomes of Redox-Responsive Polyprodrug Amphiphiles with Inflammation-Triggered Indomethacin Release Characteristics. Biomaterials. 2018. Vol. 178:608–619

            163. Tao W, He Z. ROS-Responsive Drug Delivery Systems for Biomedical Applications. Asian Journal of Pharmaceutical Sciences. 2018. Vol. 13:101–112

            164. Liu J, Li Y, Chen S, Lin Y, Lai H, Chen B, et al.. Biomedical Application of Reactive Oxygen Species-Responsive Nanocarriers in Cancer, Inflammation, and Neurodegenerative Diseases. Frontiers in Chemistry. 2020. Vol. 8:838

            165. Ni R, Song G, Fu X, Song R, Li L, Pu W, et al.. Reactive Oxygen Species-Responsive Dexamethasone-Loaded Nanoparticles for Targeted Treatment of Rheumatoid Arthritis via Suppressing the iRhom2/TNF-α/BAFF Signaling Pathway. Biomaterials. 2020. Vol. 232:119730

            166. Paruchuri BC, Gopal V, Sarupria S, Larsen J. Toward Enzyme-Responsive Polymersome Drug Delivery. Nanomedicine (London, England). 2021. Vol. 16:2679–2693

            167. Liu M, Du H, Zhang W, Zhai G. Internal Stimuli-Responsive Nanocarriers for Drug Delivery: Design Strategies and Applications. Materials Science & Engineering. C, Materials for Biological Applications. 2017. Vol. 71:1267–1280

            168. Li M, Zhao G, Su WK, Shuai Q. Enzyme-Responsive Nanoparticles for Anti-tumor Drug Delivery. Frontiers in Chemistry. 2020. Vol. 8:647

            169. Shayman JA, Tesmer JJG. Lysosomal Phospholipase A2. Biochimica et Biophysica Acta. Molecular and Cell Biology of Lipids. 2019. Vol. 1864:932–940

            170. Filkin SY, Lipkin AV, Fedorov AN. Phospholipase Superfamily: Structure, Functions, and Biotechnological Applications. Biochemistry (Moscow). 2020. Vol. 85:S177–S195

            171. Shchegravina ES, Tretiakova DS, Alekseeva AS, Galimzyanov TR, Utkin YN, Ermakov YA, et al.. Phospholipidic Colchicinoids as Promising Prodrugs Incorporated into Enzyme-Responsive Liposomes: Chemical, Biophysical, and Enzymological Aspects. Bioconjugate Chemistry. 2019. Vol. 30:1098–1113

            172. Salem RM, El-Deeb AE, Elsergany M, Elsaadany H, El-Khouly R. Intra-Articular Injection of Etanercept Versus Glucocorticoids in Rheumatoid Arthritis Patients. Clinical Rheumatology. 2021. Vol. 40:557–564

            173. Liang Y, Xu X, Xu L, Prasadam I, Duan L, Xiao Y, et al.. Non-Surgical Osteoarthritis Therapy, Intra-Articular Drug Delivery towards Clinical Applications. Journal of Drug Targeting. 2021. Vol. 29:609–616

            174. Li H, Zhuo H, Yin D, Li W, Zhang Y, Li P, et al.. Intra-Articular Injection of a Nanosuspension of Tetramethylpyrazine Dihydroxynaphthalenate for Stronger and Longer-Lasting Effects Against Osteoarthritis. Journal of Biomedical Nanotechnology. 2021. Vol. 17:1199–1207

            175. Jansen MAA, Klausen LH, Thanki K, Lyngsø J, Skov Pedersen J, Franzyk H, et al.. Lipidoid-Polymer Hybrid Nanoparticles Loaded with TNF siRNA Suppress Inflammation after Intra-Articular Administration in a Murine Experimental Arthritis Model. European Journal of Pharmaceutics and Biopharmaceutics. 2019. Vol. 142:38–48

            176. Chang MC, Chiang PF, Kuo YJ, Peng CL, Chen KY, Chiang YC. Hyaluronan-Loaded Liposomal Dexamethasone-Diclofenac Nanoparticles for Local Osteoarthritis Treatment. International Journal of Molecular Sciences. 2021. Vol. 22:665

            177. Ishida T, Ichihara M, Wang X, Kiwada H. Spleen Plays an Important Role in the Induction of Accelerated Blood Clearance of PEGylated Liposomes. Journal of Controlled Release. 2006. Vol. 115:243–250

            178. Feng X, Liu J, Xu W, Li G, Ding J. Tackling Autoimmunity with Nanomedicines. Nanomedicine (London, England). 2020. Vol. 15:1585–1597

            179. Li H, Yang YG, Sun T. Nanoparticle-Based Drug Delivery Systems for Induction of Tolerance and Treatment of Autoimmune Diseases. Frontiers in Bioengineering and Biotechnology. 2022. Vol. 10:889291

            180. Jung SM, Kim WU. Targeted Immunotherapy for Autoimmune Disease. Immune Network. 2022. Vol. 22:e9

            181. Duvvuri B, Lood C. Cell-Free DNA as a Biomarker in Autoimmune Rheumatic Diseases. Frontiers in Immunology. 2019. Vol. 10:502

            182. Hashimoto T, Yoshida K, Hashiramoto A, Matsui K. Cell-Free DNA in Rheumatoid Arthritis. International Journal of Molecular Sciences. 2021. Vol. 22:8941

            183. Zhu Y, Zhao T, Liu M, Wang S, Liu S, Yang Y, et al.. Rheumatoid Arthritis Microenvironment Insights into Treatment Effect of Nanomaterials. Nano Today. 2022. Vol. 42:101358

            184. Chen Y, Wang Y, Jiang X, Cai J, Chen Y, Huang H, et al.. Dimethylamino Group Modified Polydopamine Nanoparticles with Positive Charges to Scavenge Cell-Free DNA for Rheumatoid Arthritis Therapy. Bioactive Materials. 2022. Vol. 18:409–420

            Author and article information

            Journal
            amm
            Acta Materia Medica
            Compuscript (Ireland )
            2737-7946
            18 January 2023
            : 2
            : 1
            : 23-41
            Affiliations
            [a ]Key Laboratory of Coarse Cereal Processing (Ministry of Agriculture and Rural Affairs), School of Food and Biological Engineering, Chengdu University, Chengdu 610106, China
            [b ]School of Mechanical Engineering, Chengdu University, Chengdu 610106, China
            [c ]Sichuan Industrial Institute of Antibiotics, School of Pharmacy, Chengdu University, Chengdu 610106, China
            [d ]State Key Laboratory of Southwestern Chinese Medicine Resources, School of Pharmacy, Chengdu University of Traditional Chinese Medicine, Chengdu 611137, China
            [e ]Department of Pharmaceutical Sciences and the Biointerfaces Institute, University of Michigan, Ann Arbor, MI 48109, USA
            Author notes

            1These authors contributed equally.

            Article
            10.15212/AMM-2022-0039
            314fb04b-ecc7-4fa9-b0d8-7bb90bd56cbc
            Copyright © 2023 The Authors.

            Creative Commons Attribution 4.0 International License

            History
            : 18 October 2022
            : 20 December 2022
            : 22 December 2022
            Page count
            Figures: 5, Tables: 3, References: 184, Pages: 19
            Categories
            Review Article

            Toxicology,Pathology,Biochemistry,Clinical chemistry,Pharmaceutical chemistry,Pharmacology & Pharmaceutical medicine
            Rheumatoid arthritis,targeting drug delivery,inflammation,liposomes,nanoparticles

            Comments

            Comment on this article