Skip to main content

Advanced gene nanocarriers/scaffolds in nonviral-mediated delivery system for tissue regeneration and repair

Abstract

Tissue regeneration technology has been rapidly developed and widely applied in tissue engineering and repair. Compared with traditional approaches like surgical treatment, the rising gene therapy is able to have a durable effect on tissue regeneration, such as impaired bone regeneration, articular cartilage repair and cancer-resected tissue repair. Gene therapy can also facilitate the production of in situ therapeutic factors, thus minimizing the diffusion or loss of gene complexes and enabling spatiotemporally controlled release of gene products for tissue regeneration. Among different gene delivery vectors and supportive gene-activated matrices, advanced gene/drug nanocarriers attract exceptional attraction due to their tunable physiochemical properties, as well as excellent adaptive performance in gene therapy for tissue regeneration, such as bone, cartilage, blood vessel, nerve and cancer-resected tissue repair. This paper reviews the recent advances on nonviral-mediated gene delivery systems with an emphasis on the important role of advanced nanocarriers in gene therapy and tissue regeneration.

Introduction

Tissue defects including wound, infection and tumor removal have become a great challenge in clinic around the world [1,2,3]. Due to poor intrinsic repair capacity in somatic cells and differentiated cells, a myriad of tissues and organs are urgently desired for in vivo clinical transplantation to repair damaged tissues, organs and parts in recent years [4]. Therefore, tissue regeneration technology has been rapidly developed and widely applied in tissue engineering since it is firstly proposed by American National Science Foundation (NSF) in 1988 [5]. Tissue engineering is also defined to apply the basic criterion and manner of engineering science and life science to understand the structure-and-function relationship between physiological or pathological tissues and organs, and further develop artificial substitutes with biological activity to restore, maintain and improve the functions of tissues and organs [6]. Current strategies that are curative and work well consisting of autologous transplantation, allograft transplantation and artificial substitutes for tissue regeneration [7]. However, these methods suffer from some inevitable drawbacks, such as donor shortage and immune rejection response and so on [6, 8]. To circumambulate these problems, tremendous efforts have been contributed to explore novel tissue regeneration techniques with efficient repair ability and low immunogenicity [9, 10].

Gene therapy has provided a potential tactic to aid in the tissue regeneration and repair for destructive tissues [11, 12]. In this process, exogenous genes are delivered into targeting damage sites by gene carriers to achieve effective gene therapy of tissues and organs [13]. Generally, gene delivery carriers can be divided into viral and nonviral-mediated vectors [14]. Viral vectors can incorporate exogenous genes into DNA sequence of host cells to form stable transfection, consisting of genetic materials (RNA or DNA), protein-coated capsid and envelope of lipids surrounding the capsid [15]. Viral vectors with the advantages of stability, high transfection efficiency and extended expression window of related genes have been used in many clinical trials for cancer therapy, cardiovascular disease treatment and tissue regeneration [16,17,18,19,20]. Currently, the most commonly used viral vectors include retrovirus (RV), lentivirus (LV), adeno-associated virus (AAV), and adenovirus (AV) for gene delivery and therapy [21, 22]. Although viral-based gene delivery vector has gained a huge success in transfection efficiency and stability, the off-target effect, immune related adverse risk and restricted viral tropism are still inevitable events, which limit the further clinical application and development especially for AAV vectors [23,24,25]. These challenges motivate researchers to explore the alternative carriers to realize precise gene delivery. Advanced functional biomaterial-based gene delivery systems open a new insight for gene therapy by the means of encapsulation and immobilization of nonviral carriers [15]. Biomaterials-based gene delivery has attracted more and more attention in tissue regeneration and repair, due to their economic cost, low immunogenicity and more accurate local delivery ability [26]. Advanced biomaterials, including inorganic materials, polymer materials, lipids, proteins and peptides, exosomes and DNA nanostructures for gene therapy mainly have the characteristics of good biocompatibility, targeting accuracy ability and excellent degradability [27, 28]. Further, advanced biomaterials-based gene carriers have also been investigated and applied in tissue regeneration, such as bone, cartilage, blood vessel, nerve and cancer therapy [29,30,31,32,33,34]. Different tissue engineering application aspects with their specific features are matched with various nonviral biomaterials [35]. The biological outcome of gene delivery and therapy depends heavily on the selection of gene vectors. Better biomaterial-vector choice and collocation will pave a new way for gene therapy in the potential of tissue engineering and organ regeneration.

In this review, we emphasize the nonviral-vector biomaterials developed for gene delivery and their application in tissue regeneration and repair (Fig. 1). Advanced nonviral biomaterials, including lipid/lipid-like carriers, polymeric biomaterials, functional dendrimers, inorganic particles and exosomes, carry the targeting genes, such as encoding DNA (double-stranded DNA (dsDNA), single-stranded DNA (ssDNA), plasmids DNA (pDNA)) and RNA (Messenger RNA (mRNA), MicroRNA (miRNA), small interfering RNA (siRNA)), into damaged tissues or sites for profitable gene therapy. On the other hand, we also highlight therapeutic applications of these advanced gene delivery materials in tissue regeneration, such as bone regeneration, cartilage regeneration, blood vessel regeneration, neuroregeneration and its broad potential in cancer-resected tissue repair and therapy. Finally, the strengths and limitations of different delivery vehicles are summarized in tissue regeneration and the review also shares our future perspectives in gene therapy and tissue regeneration. This will provide important information for advanced nonviral gene delivery materials in the potential of tissue engineering.

Fig. 1
figure 1

Nonviral-mediated advanced functional carriers for precise and targeting delivery of different nucleic acids, including dsDNA, ssDNA, siRNA, miRNA, mRNA and so on. Advanced biomaterials, such as lipid/lipid-like carriers, polymeric biomaterials, functional dendrimers, inorganic particles and exosomes, are intensively investigated for efficient gene delivery and therapy

Nonviral-mediated gene delivery systems

The development of gene delivery and therapy

Gene delivery and therapy are considered to introduce exogenous normal or targeting genes (DNA or RNA) into objective cells to recompense for diseases aroused by defects and abnormal genes, so as to reach gene-related therapeutic purpose [36]. The conception of gene therapy is nominated in the early 1970s, the initial conceptualization is to interpose an accustomed gene to supersede a freakish gene [37, 38]. In 1990, the first clinical evaluation of a gene therapy approach to combating disease is launched at National Insititutes of Health (NIH) in the United States [39]. Since 2015, the worldwide gene therapy trade has flourished precipitately and also assessed that in 2025, the domestic and worldwide market of gene therapy will reach nearly 31 billion US dollars [40,41,42]. Therefore, gene therapy has been considered the third industrial revolution in bioloical medicine field after small molecule drugs and antibody drugs. Despite the rapid establishment of gene therapy, there are still rate-limiting steps, for example, the lack of efficient delivery system, weak sustained gene expression, and host immune reactions [43]. The Achilles’ heel of gene therapy is that how the targeting genes are effectively and accurately delivered into damaged sites for gene-mediated treatment. At present, gene delivery systems used in clinical trials are the basis for RV, LV, poxviruses, AV, AAV and herpes simplex virus (HSV) [44]. Nevertheless, inefficient delivery and transient gene expression are common issues with these vectors [45, 46]. The delivery system of advanced biomaterials will bring us a new dawn for desired gene therapy. Therefore, the development of drug delivery systems based on advanced functional biomaterials may serve as more effective drug and gene delivery vehicles for the treatment of harmful injury and malignant diseases. Moreover, we have also summarized the nonviral vectors under clinical trials or approved for human use currently (Table 1). The nonviral vectors have broadened human applications in the clinic.

Table 1 Nonviral vectors under clinical trials or approved for human use

Advanced gene nanocarriers/scaffolds in nonviral-mediated delivery systems

Many kinds of advanced biomaterials including lipid/lipid-like carriers, biopolymeric materials, functional dendrimers, inorganic biomaterials, exosomes and DNA nanostructures are designed and prepared to achieve efficient gene delivery and therapy [52, 53]. To improve gene delivery ability, there are various strategies based on targeting genes and advanced functional biomaterials to transfer exogenous genes and particles into impaired cells for transfection [54,55,56]. Common targeting genes contain encoding pDNA, mRNA, miRNA, siRNA, shRNA and circular RNAs (circRNAs) [57, 58]. Correspondingly, lipid carriers, cationic polymers, peptides, dendritic/branched materials and inorganic nanoparticles as gene delivery vectors are also performed well to enhance gene delivery and transfection. Moreover, extracellular vesicles (EVs) with the advantages of non-immunogenic, biodegradable, and biocompatible ability are considered as a promising delivery vehicle for various genetic therapeutics [59]. Lipid, lipid-like compounds and polymer compounds are one of the most intensively studied vectors for therapeutic gene delivery [60]. Herein, we will introduce these nonviral-mediated advanced biomaterials used for gene delivery and therapy in details.

Exosomes/DNA system

Extracellular vesicles (EVs) are regarded as small membrane vesicles to release into the extracellular matrix (ECM) and play an important role in regulating gene delivery and therapy [61, 62]. Exosomes with lipid-bilayer structure and the diameter of 40–160 nm are derived from different cells types, including stem cells, mast cells, epithelial cells, dendritic cells, B cells, T cells and cancer cells, to circulate in the extracellular environment [63, 64]. Exosomes from different cells have various functions and are applied in disease therapy [65, 66]. There are generally two forms of exosomes: One is biomimetic vesicles and the other is hybrid vescicles (Fig. 2A, B). For gene therapy, exosomes usually act as the efficient gene delivery system with versatile advantages, such as crossing biological barriers and genetically engineerable potential [67]. Yang and collaborators present an improved strategy to realize more efficient encapsulation and manageable release in recipient cells by encapsulating low-density lipoprotein receptor (Ldlr) mRNA into EVs [68]. There is an improved loading efficiency of Ldlr mRNA in EVs via MS2-MCP interaction [68]. Wood and collaborators also demonstrated that exosome-endogenous nanovesicles enable to transport short interfering (si)RNA to the brain in mice. GAPDH siRNA is specifically delivered by intravenously injected RVG-targeted exosomes to oligodendrocytes, microglia, neurons, causing a particular gene knockdown [69]. The integration of cell-derived EVs and synthesized polymers is further applied to complement targeting gene delivery capability. For instance, Gupta and collaborators successfully report a hybrid system of exosomes and PEI matrix (EPM) to provide high nucleic acid entrapment and protection from enzymatic degradation. The nanostructures can also enhance the specificity of delivery by combining a targeting ligand with no adverse response (Fig. 2C-E) [70]. In addition of the cell-derived EVs as gene delivery carriers, Li and collaborators develop a nanocarrier of CD44-specific ligand hyaluronic acid (HA)-modified milk-derived exosomes (mExo) encapsulated with folic acid (HA-mExo-FA). HA-mExo-FA could promote the apoptosis of activated LX2 cells and improve liver morphology and function alterations in zebrafish larvae (Fig. 2F, G) [71].

Exosomes are genetically engineered by the modification of polypeptide and nucleic acid to enhance the targeting gene therapy [63, 72]. For example, Liang and collaborators have fabricated an exosome-based chondrocyte-targeted miRNA (microRNA-140) delivery system for osteoarthritis (OA) therapy [73]. The chondrocyte-affinity peptide (CAP)-exosomes could carry miR-140 into profound cartilage sites through the dense mesochondrium to prevent cartilage-degrading proteases and slow down the OA progression. Moreover, exosomes can also transfer the CRISPR-Cas9 plasmids into targeting sites for gene therapy [74]. Kim and collaborators have reported that cancer-derived exosomes loaded CRISPR/Cas9 could restrain the expression of poly(ADP-ribose) polymerase-1 (PARP-1) to enhance the chemosensitivity of cisplatin and achieve satisfactory therapeutic effects in ovarian cancer tumors [75].

Fig. 2
figure 2

Exosomes as gene delivery vectors and their biomedical applications in gene therapy. (A) The production of biomimetic vesicles derived from cells through extrusion/filtration method. (B) The biomedical application of exosomes. (C) Representative illustration of FA covalently attached to the exosome-polyethyleneimine matrix (EPM) as a transfer carrier for nucleic acid delivery. (D) The entrapment ability of EPM with siKRAS and siRNA. (E) Tissue distribution of bovine colostrum exosomes and EPM, with and without FA-functionalized exosomes. (C-E) Reproduced with permission [70]. Copyright 2021, Elsevier. (F) Schematic graph of HA and FA loading into mExo. (G) Scanning electron microscope (SEM) photos of mExo and HA-mExo-FA. (F, G) Reproduced with permission [71]. Copyright 2023, Wiley

Based on structural and chemical traits of EVs and DNA, DNA nanostructure is constructed by the complementary nature of the four nucleotide bases as well [76]. By precisely controlling the DNA hybridization, stable branched DNA self-assembly, DNA programmability and desired DNA-sequence synthesis, DNA nanoparticles are fabricated by decorating DNA structures and sequences to deliver the designed genes for gene therapy [77, 78]. Moreover, EVs can also be modified by DNA aptamers or other molecules to realize the target activity or other functions for broad potentials in clinic [79, 80].

Lipids

Lipid-induced gene delivery platform works for gene therapy due to the effect of cationic lipid molecules [81, 82]. In principle, the positive charges on head group could interact with the negative charges of nucleic acids to establish targeting lipid/gene complexes [83, 84]. Then, the lipid/gene complexes could fuse with cell membrane and eventually enter into the targeting cell for gene silencing or integration in gene-related disease treatment [85]. The commercial Lipofectamine 2000 and Lipofectamine 3000 are the most commonly transfection reagent with their characteristics of simple operation, easy acquisition and high transfection efficiency [86, 87]. For instance, when these two reagents are applied to deliver mRNAs into dorsal root ganglion (DRG) neurons, the transfection efficiency can reach up to 25% under the assistance of Lipofectamine 2000 transfection reagent [88].

In addition, lipid and lipid-like carriers are also used for delivery of both small nucleic acids (siRNA, miRNA) and large ones (pDNA, mRNA) [89]. Kranz and collaborators have reported that RNA-lipoplexes (RNA-LPX) with adjusted net-like structures could enter into dendritic cell (DC) populations and macrophages to realize stable gene expression of the encoded antigen for cancer immunotherapy [90]. Rajala and collaborators have also designed and synthesized an artificial virus by using the functional nanoparticles of liposome-protamine-DNA (LPD) complexes and modifying with cell permeable peptide or nuclear localization signaling (NLS) peptide, to deliver the targeting engineered genes for eye disease treatment [91]. The results revealed that LPD could encourage effective gene delivery in a cell specific manner and achieve the long-term expression of Rpe65 gene to Rpe65-lacking mice, thus inducing blindness correction in vivo [91]. Due to the specific advantages of lipids in easier design for ideal properties by altering lipid components, particle size and surface charges, lipids have more broad application potentials in gene delivery and therapy [92].

Cationic polymers

Similar to the mechanism of liposomes in gene delivery, cationic polymers also perform the delivery functions by the interaction with the negatively charged nucleic acids [93]. Cationic polymers with ionizable head groups can bind and condense DNA into small molecular structures to carry targeting genes for disease treatment [94]. Efficient transfer and cell internalization of DNAs are regulated through electrostatic interactions between cationic lipids and negatively charged nucleic acids or plasma membrane components [95]. For example, the amine-based cationic materials, including polylysine, polyamidoamine (PAMAM), PBAEs, poly(ethyleneimine) (PEI), cationic dendrimers, and chitosan, have also been explored as DNA vectors for gene delivery and therapy [96,97,98]. Among these cationic polymers, biofunctional PEI (50 kDa) supramolecular structures have demonstrated excellent transfection efficiency in vivo, compared to other amine-based polymers [99]. PEI serves as a classic and effective transfection agent for gene delivery. PEI possesses two structure forms, one is linear chain and another is branched chain, which can be synthesized at different condition (Fig. 3A) [100]. About 20% or 16.7% of the amino nitrogen is protonated at physiological pH [99, 101,102,103]. Besides, PEI has a high buffering capacity in a very wide pH range to escape from lysosomes and enhance the delivery productivity [104, 105]. Rohiwal and collaborators have constructed the targeting delivery strategy of CRISPR/Cas9 complexed PEI magnetic nanoparticles (MNPs) for gene therapy [106]. The 20-nm-in-diameter PEI-MNPs with sufficient colloidal stability show higher gene-editing efficiency, even compared with standard Lipofectamine transfection. To further improve the transfection efficiency, Lemaire and collaborators design and synthesize a depolymerized chitosan-polyethyleneimine DNA complex (dCS-PEI/DNA) for effective gene delivery (Fig. 3B). The results reveal that dCS-PEI/DNA complexes present the improved nucleic acid encapsulation ability, cellular uptake efficiency and transfer potency in human hepatoma HuH-7cells and murine primary hepatocytes (Fig. 3C) [107].

In addition of PEI-based gene carriers, PAMAM amino biomaterials are also applied as gene delivery vectors for efficient gene therapy. Volonterio and collaborators report the functionalization of the outer primary amines of PAMAM G2 and PAMAM G4 by building blocks bearing fluorinated moieties with a guanidino functional group. The obtained conjugates have the prominent advantages of improved transfection efficiency and negligible cytotoxicity (Fig. 3D) [108]. Nevertheless, poor biocompatibility, low biodegradability and high manufacturing costs greatly hinder their clinical applications for gene delivery and gene therapy.

Fig. 3
figure 3

Cationic polymers as gene delivery vectors and their biomedical potentials. (A) PEI-based nanoparticle formation is designed by the admixing of PEI polymer derivatives (linear or branched) with nucleic acid cargos to encode for any genes of interest. (B) Schematic overview of ternary complex formation by sequential addition of core and shell polymer to the nanovector DNA. (C) In vitro evaluation of toxicity and transfection efficiency in HuH-7 cells. (B, C) Reproduced with permission [107]. Copyright 2023, Wiley. (D) Structural illustration of selectively fluorinated PAMAM–Arg conjugates and transfected cell micrographs. Reproduced with permission [108]. Copyright 2023, American Chemical Society

Inorganic materials

Inorganic materials are also implemented by biofunctional modification as gene delivery carriers for purposeful gene therapy. The commonly used inorganic nanoparticles (NPs) consist of carbon-based NPs (quantum dots, graphene, carbon nanotubes and nanofibers), metal-based NPs (gold, copper, silver), semiconductor-based NPs (germanium, iron oxide) and silica NPs (SiO2) [109]. These inorganic NPs can be functionalized by biophysical or biochemical modification to adjust specific size of inorganic NPs and endow unique surface properties for precisely targeted gene delivery and therapy [110].

To enhance gene delivery ability of inorganic NPs, gold NPs are modified with nucleic acids via covalent or non-covalent conjugation through thiol moieties [111,112,113,114]. Dai and collaborators have fabricated the Janus gold/chitosan (J-Au-CS) NPs to realize fluorescence imaging-guided photothermal therapy (PTT) and PTT-enhanced gene therapy for cancer therapy [115]. Mesoporous silica nanoparticles (MSNs) with the diameter of 100–250 nm have also been used in gene delivery due to its good biocompatibility, porous size controllability and stable aqueous dispersion [116,117,118]. These MSNs are usually loaded with targeting genes by the means of weak non-covalent interactions [119]. Rasool and collaborators have also developed a thiolated, bioactive mesoporous silica NPs to improve the cell adhesion property in bone tissue regeneration [120]. The results reveal that MSN-SH groups can increase calcium deposition and induce osteogenesis by altering the bone-related gene expression [120]. Although inorganic materials play a vital role in gene delivery and therapy, especially in the cargo-loading ability and gene release rate, the toxicity and degradability are still unavoidable problems for gene delivery in vivo [121]. Some studies have demonstrated that gold and iron oxide nanoparticles are regarded as ideal non-toxic carriers for tissue regeneration and disease therapy [116, 122]. Moreover, the metabolism of these inorganic materials remains still a matter of concern in gene delivery.

Proteins and peptides

Protein and peptide-based formulations are also important delivery carriers, which usually consist of 5–30 amino acids with mostly hydrophobic and/or positively charged side chains [123]. Cell-penetrating peptides (CPPs) or homing peptides are typically representatives for gene delivery [124]. Positive charge in proteins and peptides can interact with negative charge by the means of covalent or non-covalent binding to enter into the cells [125]. Wang and collaborators have designed and prepared a library of glutathione (GSH)-responsive silica nanocapsules (SNCs), which could systemically deliver SNCs conjugated with glucose [126]. Under glycemic control, rabies virus glycoprotein peptide enables to bypass the intact blood-brain barrier (BBB) to achieve brain-wide transfer of various biologics and CRISPR genome editors can target various genes in both wild-type mice and Ai14 reporter mice [126]. Nevertheless, endowing CPPs with targeting ability and decreased toxicity remain great advantages for gene delivery of proteins and peptides in clinical potential.

In summary, DNA condensation, protection from nuclease, cell targeting, endosomal disruption, and nuclear translocation are the key factors for a successful gene delivery vector. For mechanism of biomaterials-based vectors entry into cells, most polymer/DNA complexes are uptaken into membrane-bound compartments by pinocytosis, adsorptive endocytosis, receptor-mediated endocytosis, or phagocytosis. After entering into the cells, they will escape from endosomal degradation using endosomolytic peptides or osmosis, and possibly obtain nuclear entry via an active nuclear skeleton mechanism. Therefore, cell-specific targeting ligands have been designed to link polymeric vectors to enhance the cell targeting efficiency by improving receptor-mediated endocytosis. For example, transferrin, monoclonal antibody, mannose, galactose, lactose, folic acid, low-density lipoproteins, and RGD peptides are the cell-specific targeting ligands to improve cellular internalization and gene transfection. In addition, biomaterial-mediated nonviral gene vectors also show respective advantages and disadvantages in gene delivery, including exosomes, lipids, cationic polymers, inorganic materials, proteins and peptides (Table 2).

Table 2 Advantages and disadvantages of biomaterial-mediated efficient nonviral gene delivery

Gene therapy for tissue regeneration

Gene therapy serves as a crucial role in modulating tissue regeneration and repair [127]. Regeneration is the process of self-renewal and repair by which a partial organism replaces lost or damaged tissue. Regenerative capacity widely varies across species, tissues, and life stages, and the process is also genetically programmed to accomplish injured tissue repair [128]. In the evolution of tissue regeneration, gene therapy can achieve point-to-point precise and targeted therapeutic effect of tissue repair through the transfer of genetic materials [129, 130]. The activity of DNA regulatory elements, transcription factors and chromatin regulators co-determine regeneration programming for tissue engineering [128]. For tissue regeneration, the targeting precision is critical to obtain the high efficiency and safety of gene therapy for disease diagnosis indications, such as tissue damage [130]. Yan and collaborators have found that tissue regeneration enhancer elements (TREEs) isolated from zebrafish can direct targeted, injury-associated gene expression from viral DNA vectors delivered systemically in small and large adult mammalian species [131]. These findings indicate that targeting gene delivery and therapy present huge potentials in tissue regeneration and repair.

Tissue regeneration and repair

Tissue Engineering is a new discipline emerging in recent years, which belongs to the category of biological high technology [6, 132]. The term of tissue engineering is firstly proposed at the Bioengineering Group Meeting held by the American Science Foundation in Washington in 1987. It is officially defined in 1988 as: Combing the basic principles and new technologies of life science and engineering to correctly understand the normal and pathological conditions of mammals. On the basis of the relationship between tissue structure and function in the two states, it is an emerging discipline to study and develop biological substitutes for repairing, maintaining, and promoting the functions and forms of various tissues or organs of the human body after damage [6, 8]. Specially, tissue regeneration has ushered in vigorous process with the development of nanomedicine engineering and gene therapy in recent years.

Tissue engineering is divided into four aspects: Cell seeding and culture, advanced functional biomaterials for mimicking cell microniches, techniques for creating tissues and organs, and clinical utilization of tissue engineering [133, 134]. Currently, three approaches are constructed for tissue repair in clinical practice: (i) autologous tissue transplantation, (ii) allogeneic tissue transplantation, and (iii) the application of artificial substitutes [130, 135]. The three methods are respectively deficient and confronted with some mutual problems, such as immune rejection and donor shortage. The development of tissue engineering will fundamentally solve the problem of dysfunction or loss of treatment caused by tissue and organ defects [135]. The decisive essence of tissue engineering is to establish a three-dimensional (3D) space complex composed of cells and bioactive materials, which is radically different from traditional two-dimensional (2D) structures, such as cell culture patterns, nutritional supplement modes and 3D microenvironment niches [136, 137]. Therefore, the basic principles of tissue regeneration are: (1) simple reconstruction of tissue shape, structure and function to achieve permanent replacement, (2) abundant usage of least tissue cells to repair large tissue defects after in vitro culture and expansion, and (3) arbitrary geometry switching according to the tissue and organ defects to achieve a perfect shape repair [138].

Advanced gene nanocarriers/scaffolds for gene therapy in tissue engineering

Advanced functional biomaterials can act as gene carriers and regulators to deliver targeting genes or growth factors into corresponding damaged tissue sites to achieve tissue regeneration [139]. The tissue regeneration related cells, including progenitor cells, induced pluripotent stem cells (iPSCs), embryonic stem cells (ESCs) and mesenchymal stem cells (MSCs), can interact with the advanced biomaterials and promote the expression of some functional proteins related to cell anchoring junction (Fig. 4) [140,141,142,143,144]. The expression of collagen, fibronectin and laminin demonstrates the formation of adhesion anchorage between cells and extracellular matrix (ECM) [145, 146]. The transmembrane adhesion protein, integrin primarily binds collagen and fibronectin to produce adhesion sites for cell attachment, which is beneficial for migration, division and differentiation of stem cells in the process of tissue regeneration [147,148,149,150]. It can also connect a series of intracellular anchor proteins, such as talin, vinculin and β-catenin for cell spreading and adhesion [151,152,153,154]. Eventually, talin, vinculin, and β-catenin interacted with other molecules to form physical bridges, can alter the cytoskeleton and nuclear transcription to promote functional cell fate [155]. At the same time, gene editing (i.e. knock in and knock out of targeting genes) is also taken up in this process to enhance gene therapy in tissue regeneration [156, 157]. According to the discussion of gene therapy in tissue regeneration, gene therapy based on advanced functional gene delivery materials can play a vital role in tissue engineering and repair, such as bone regeneration, cartilage regeneration, blood vessel regeneration, nerve regeneration and repair (Table 3).

Fig. 4
figure 4

ECM remodeling and skeleton-mediated cellular nanomechanics enhance gene delivery through activation of force-sensing-related signaling pathways. Interaction of tissue regeneration related cells, including progenitor cells, iPSCs, ESCs and MSCs with advanced biomaterials to promote the expression of some functional proteins related to cell anchoring junction, such as collagen, fibronectin and laminin. These proteins can connect a series of intracellular anchor proteins, such as talin, vinculin and β-catenin for cell spreading and adhesion to alter the cytoskeleton and nuclear transcription to promote cell migration, division and differentiation and functional cell fate. Gene editing (knock in and knock out of targeting genes) also participate in this process to enhance gene therapy ability in tissue regeneration

Table 3 Advanced gene/drug nanocarriers for gene therapy in tissue engineering

Advanced gene nanocarriers/scaffolds for gene therapy in bone regeneration

Long-distance bone defects, caused by trauma, tumor resection, and infection remain a worldwide concern due to the autologous regeneration insufficiency and alloimmune reaction. To date, the “gold standard” for bone defect repair is autologous bone grafting in clinic [170, 171]. Nonetheless, the extended treatment time at the graft donor site, significant pain, and increased treatment costs are the major limitations for bone regeneration [172]. On the other hand, bone allografts have the inevitable risk of immune rejection and bone size mismatch [173]. Moreover, metal bone implants have the drawbacks of poor osseointegration [174]. Therefore, the development of tissue-engineered bones based on the combination of efficient gene therapy has broad prospects in recent years.

Gene therapy is considered as an ideal way in many ways to enhance bone regeneration ability due to precisely targeting bone repair. Gene therapy is interpreted to transfer specific gene products to precise anatomical positions [175, 176]. In addition, transgene expression level and expression duration can be modulated with existing techniques [177]. For bone regeneration, genes of interest are transferred to the fracture locations, and expressed at suitable levels, and then inactivated after fracture healing [178]. The transmission of biological factors, primarily bone morphogenetic proteins (BMPs), has produced encouraging results in both animal and clinical studies [179]. A great deal of work has also been done on exploring previously defined growth factors and inventing new growth factors [158]. To enhance gene therapy capacity, gene delivery systems are designed from viral and nonviral vectors to tissue-engineered scaffolds to make significant progress [159]. Despite some public hesitation about gene therapy, this technology has huge potential applications to broaden our horizons to cure a wide range of human skeletal or musculoskeletal disorders.

Nonviral-mediated gene vectors contain chemical delivery and physical delivery methods [180]. Chemical delivery is mainly to complex cationic biomacromolecules or lipids with DNA/RNA to induce the electrochemical repulsion with cell membranes [160]. Physical methods of electroporation and sonoporation take advantage of electrical and mechanical energy respectively to build transient pores in cell membranes for efficient internalization of exogenous DNAs [56, 181]. Curtin and collaborators have reported a nonviral dual delivery system of vascular endothelial growth factors (VEGF) and bone morphogenetic protein 2 (BMP2) in a collagen-nanohydroxyapatite (nHA) scaffold, which reveals a marked 36-fold increase in bone formation in nHA scaffold versus the empty defect as early as 4 weeks post-implantation underlining the immense ability in bone regeneration [182]. In this scaffold, nHA particles and PEI are used to serve as nonviral vectors to achieve pDNA delivery for efficient gene therapy [182].

Fig. 5
figure 5

Biomaterials-based gene delivery vectors for bone regeneration. (A) Non-coding-RNA-activated core/chitosan shell nanounits coated with polyetheretherketone with the ability to promote bone regeneration. (B) Micro-CT images demonstrating bone regeneration surrounding implants, as well as reconstructed 3D images of samples and new bone. The color yellow denotes samples, whereas the color red represents fresh bone regeneration. Quantitative bone analysis 8 weeks after implantation. (A-B) Reproduced with permission [183]. Copyright 2023, American Chemical Society. (C) CRISPR activation of DANCR promotes calvarial bone formation. (D) µCT analysis of in vivo calvarial bone repair. 3D projection of osteomized sites of mock and Bac-Sa-VPR groups (left) and frontal views of osteomized sites (right). (C-D) Reproduced with permission [184]. Copyright 2021, Elsevier. (E) Schematic diagram of conjugating Asp8 to exosomal amine groups by click chemistry reaction. (F) H&E staining and Masson’s trichrome staining analysis of defect sites at 2 weeks, scale bar = 100 μm. (E-F) Reproduced with permission [185]. Copyright 2023, Elsevier

Advanced functional biomaterials can serve as gene delivery carriers for gene therapy in bone regeneration. Li and collaborators develop non-coding-RNA-induced core/chitosan shell nanounits to improve bone incorporation by immunoregulation (Fig. 5A). miMT-PEEK contributed macrophage M2 polarization via the NF-κB pathway to enhance BMSCs osteogenic differentiation. In vivo, miMT-PEEK increases osteogenesis and osseointegration, verified by Micro-CT images and histological images (Fig. 5B) [183]. Moreover, Hu and collaborators firstly verify that DANCR expression promotes rASC chondrogenesis. Then, the CRISPR activation (CRISPRa) technology is conducted to upregulate endogenous DANCR, promote rASC chondrogenesis and increase calvarial bone healing in rats (Fig. 5C). The µCT analysis in vivo also verifies the effect of DANCR NPs on calvarial bone repair (Fig. 5D) [184]. Further, Zhao and collaborators have designed Asp8-EM/miR-26a to enhance the synergetic effects for bone regeneration with increased targeting ability by Asp8 modification in bone defect and osteoporosis models (Fig. 5E-F) [185]. Therefore, the nonviral-mediated gene delivery system can regulate gene therapy ability in bone regeneration and make great achievement for clinical potential.

Advanced gene nanocarriers/scaffolds for gene therapy in cartilage regeneration

Advanced gene delivery biomaterials are applied to control targeting gene delivery efficiency into desired damaged sites for gene therapy and cartilage regeneration [186]. For instance, untreated damage to articular cartilage often results in osteoarthritis to induce inevitable disability [187]. Bone-marrow stimulation procedures, autologous or allograft, and periosteum transplantation are the main treatment methods in clinic for cartilage repair, for example, CuS/TGF-β1@PC NPs are prepared to deliver genes for cartilage regeneration [161, 188]. Nevertheless, these processes are usually accompanied by surgical trauma and poor long-term efficacy. Besides, the limited availability of donor sites and immune rejection may also restrict the broad application of gene therapy in cartilage regeneration [189]. Therefore, new comprehensive cartilage repair strategies are urgently needed based on efficient gene delivery and therapy. Advanced biomaterials-guided gene delivery provides the novel solution for precise delivery to direct targeted gene therapy [190]. Up to now, inorganic NPs, liposomes and cationic polymers are the commonly used biomaterials for gene delivery in cartilage regeneration [190]. These NPs can spatiotemporally enhance the process of cartilage repair by improving gene delivery ability [191]. Based on the superiority, the designed strategies of different biomaterials are also introduced and their important role is highlighted in gene therapy for cartilage repair.

Fig. 6
figure 6

Biomaterials-based gene delivery vectors for cartilage regeneration. (A) Schematic illustration of the preparation of the Cu-based NPs. CuS NPs are reacted with ATPES to obtain CuS-NH2 NPs, and then electrostatically bonded with negatively charged TGF-β1 pDNA. Then PC is coated on the NPs for biomimetic modification. (B) Representative images of Safranin-O/Fast green staining and Toluidine blue staining of the knee joints after different treatments. (A-B) Reproduced with permission [161]. Copyright 2021, Elsevier. (C) The silencing of miR-221 in combination with TGF-β3 improved sGAG distribution. Representative histological images of cell-seeded scaffolds stained with (A) hematoxylin and eosin and (B) thionine after 28 days in culture +/− TGF-β3 supplementation. Scale bars represent 100 μm length. (D) Representation of GET peptide: a glycosaminoglycan (GAG)-binding peptide sequence (P21), fused to an amphiphilic region with an octa-arginine (8R) and a cell-penetrating peptide (CPP). (E) The incorporation of miR-221 inhibitor to CI/II-HyA scaffolds reduced the expression of miR-221 by hMSCs. (C-E) Reproduced with permission [163]. Copyright 2021, Wiley

Inorganic NPs occupy an important position in gene delivery and therapy for cartilage regeneration [192]. Cai and collaborators have designed and synthesized the biomimetic cupper sulfide@phosphatidylcholine (CuS@PC) NPs loaded with pDNA encoding transforming growth factor-beta 1 (TGF-β1) to operate MSCs for increased osteoarthritis treatment via cartilage regeneration (Fig. 6A, B) [162]. CuS/TGF-β1@PC NPs are more efficient than commercial transfection agent for gene delivery and present increased cell migration, chondrogenesis and inhibition of ECM degradation [162]. However, the low transfection efficiency and potential cytotoxicity restrict its further development and application in clinic. To detour these disturbances, liposomes have been widely recognized as a commercial transfection reagent for efficient gene delivery due to good biocompatibility [193]. At the same time, the influence of polymer-based carriers on gene delivery and therapy for cartilage repair are also studied and applied in gene delivery process [190]. For instance, hydrogels become research hot spot for gene delivery and therapy [194]. Dong and collaborators have developed an injectable chitosan/silk fibroin hydrogel for cartilage repair, which could release stromal cell-derived factor-1 (SDF-1) and kartogenin (KGN), in turn to enhance the recruitment and chondrogenic differentiation of MSCs spatiotemporally [195]. For autologous treatment, the poor regenerative ability of adult cells and the inflammatory state of the injured joint are the main obstacles. Bonato and collaborators have also encapsulated the TAK1-KO (TGF-β-activated kinase 1 knockout) chondrocytes into a hyaluronan hydrogel to deposit copious cartilage extracellular matrix proteins and facilitate integration onto native cartilage, even under proinflammatory conditions (Fig. 6C-D) [163]. In another study, Chen and collaborators have used polyamidoamine dendrimer with amino acids as efficient vectors to deliver microRNA-224-5p (miR-224-5p), and the vector could condense miR-224-5p into transfected nanoparticles, which can show higher cell internalization and transfer ability compared to Lipofectamine 3000, and also protect miR-224-5p from RNase degradation to enhance miR-224-5p delivery ability [196].

Organic/inorganic nanohybrids also serve as multifunctional gene delivery systems in cartilage regeneration [197]. In particular, nanohybrids comprising of cationic polymers and inorganic NPs are considered promising candidates as advanced functional gene delivery system [198]. To promote gene delivery and therapy efficiency, glycosaminoglycan is designed as advanced gene delivery carriers for cartilage regeneration [199]. Claudio and collaborators have fabricated the microRNA (miR)-activated scaffolds, in which glycosaminoglycan-binding enhanced transduction (GET) NPs system is used to encapsulate the miR-221 inhibitor [200]. The miR-activated scaffolds can successfully transfect human MSCs with the miR-221 cargos in a sustained and controlled manner up to 28 days, which is promising to enhance cartilage defect repair. Nucleic kinase substrate short peptide (pNNS)-conjugated chitosan (pNNS-CS) is also applied to improve gene delivery and therapy ability for cartilage repair [201]. Zhao and collaborators have reported the enhanced transfection efficiency of a pDNA-chitosan (pDNA-CS) complex using a phosphorylatable nuclear localization signal-linked nucleic kinase substrate short peptide (pNNS) conjugated to chitosan (pNNS-CS) [202]. On the other hand, they also investigate the effects of pNNS-CS-mediated miR-140 and IGF-1 in both rabbit chondrocytes and cartilage defects model. The results indicate that pNNS-CS serves as an excellent gene delivery vector in gene therapy for cartilage defects and that miR-140 combination with IGF-1 transfection has better biological effect on cartilage defects.

Exosomes are one of the most popular gene delivery vehicles, especially for cartilage regeneration and repair [203]. Mao and collaborators have suggested that exosomal microRNA-92a-3p (miR-92a-3p) can regulate cartilage development and homeostasis by directly targeting WNT5A [204]. Exosomal miR-92a-3p may act as a Wnt inhibitor and exhibit the potential as a disease-modifying osteoarthritis drug. In addition, Xu and collaborators have also reported that BMSC-Exos can deliver miR-326 to chondrocytes and cartilage to improve OA by targeting HDAC3 and STAT1/NF-kappa B p65 to inhibit pyroptosis of chondrocytes and cartilage [205].

Although biomaterial-based gene therapy has achieved great success in cartilage regeneration and repair, there are little nonviral vectors to achieve extremely high transfection efficiency [190, 206]. It is still an important task to improve transfection efficiency as much as possible on the premise of ensuring safety [190]. In addition, it is necessary to enhance the matching ability of gene release with the various stages of the chondrogenesis process [207]. Therefore, once these shortcomings are solved, it still can’t stop biomaterial-based gene therapy from representing the future direction for cartilage regeneration.

Advanced gene nanocarriers/scaffolds for gene therapy in blood vessel regeneration

Angiogenesis, the growth of new vessels, is a vital process for tissue regeneration of circulation research [208]. During the process of blood vessel regeneration, advanced biomaterial-based nonviral vectors play a vital role in regulating gene delivery and therapy. Among them, inorganic NPs, liposomes and cationic polymers are common gene carriers for blood vessel regeneration [52, 208].

Mesoporous silica NPs are considered as gene therapy vectors and used in blood vessel regeneration [164]. To enhance delivery ability of mesoporous silica NPs, Wang and collaborators have investigated that mesoporous silica NPs-encapsulated miR-124 inhibitor increases the expression of SCF/c-kit protein by targeting P2Y(12) to improve the regeneration of cerebral blood vessels in lacunar cerebral infarction [165]. Although inorganic materials present the significant power in controlling gene delivery and therapy, the degradation and metabolic toxicity are still not negligible [209]. Alternatively, polymer-based gene therapy is showing huge advantages at this time. For instance, hydrogel scaffolds have the important influence on gene delivery [210]. Hu and collaborators have constructed exosome-embedded GelMA (GelMA-Exo) hydrogels with the ability of promoting diabetic wound healing, which could also deliver circ-Snhg11 and thereby increase the survival of endothelial cells to maintain endothelial cell (EC) function by activation of miR-144-3p/NFE2L2/HIF1α signaling pathway (Fig. 7A) [166]. Further, Zheng and collaborators have also found that the miR-22 loading laponite hydrogels inhibit the neointimal formation to reduce the infiltration of the macrophages and reverse the adverse vascular ECM remodeling by the upregulation of miR-22 and downregulation of its target genes methyl-CpG binding protein 2 (MECP2), which may offer a novel strategy to treat cardiovascular diseases (Fig. 7B, C) [211]. Polymer-based nanofibers also serve as a common delivery vector for gene delivery [212, 213]. Zhou and collaborators have fabricated a bilayer vascular scaffold via emulsion electrospinning of poly(ethylene glycol)-b-poly(l-lactide-co-ε-caprolactone) (PELCL), dual-power electrospinning of poly(ε-caprolactone) (PCL) and gelatin (Fig. 7D) [214]. The results demonstrate that the inner layer of PELCL loading with miR-126 could improve endothelialization ability in vivo. Local delivery of miRNA-based nanofibers may be an effective method to facilitate blood vessel regeneration [213, 215]. Maryam and collaborators have found that incubation of human endothelial cells (HUVECs) with polyhedral oligomeric silsesquioxane NPs (POSS NPs) could contribute to the modulation of angiogenesis and exosome biogenesis, which benefits to angiogenesis [216]. Some different types of polymer materials, for example, poly(L-lactic acid) (PLLA), poly(L-lactic acid)-co-poly(-caprolactone) [P(LLA-CL)] and eN-(2-hydroxypropyl)methacrylamide (HPMA) copolymer, also play a crucial role in regulating gene delivery and therapy for blood vessel regeneration [217, 218].

Fig. 7
figure 7

Biomaterial-based gene delivery vectors for blood vessel regeneration. (A) Hypoxia-pretreated ADSC-derived exosome-embedded hydrogels promote angiogenesis and accelerate diabetic wound healing. Reproduced with permission [166]. Copyright 2023, Elsevier. (B) The gelation of the laponite hydrogels in vitro. (C) Co-staining of the injured rat common carotid arteries with eNOS and αSMA 14 days after the treatment with the miR-22 loading laponite hydrogels. (B, C) Reproduced with permission [211]. Copyright 2022, Elsevier. (D) Morphology of the prepared vascular scaffold. SEM micrograph of the cross section and larger magnification. Reproduced with permission [214]. Copyright 2016, Elsevier

Polymer-based gene delivery materials can not only take part in gene delivery and therapy, but also control many other processes of endothelial cells, for example, nanocarriers with good biological activity due to surface and volume effects are more conducive to gene loading, cell adhesion and proliferation [219]. For hydrogel scaffolds, they can offer the possibility of sustained blood vessel regeneration [220]. Further, the binding and migration ability of transfected cells is limited in hydrogels for blood vessel regeneration [221]. Therefore, polymer-based materials allow the combination of multiple components and forms, and the integration of multiple materials has become the next development direction of biomaterial-mediated gene therapy.

Advanced gene nanocarriers/scaffolds for gene therapy in neuroregeneration

Peripheral nerve injuries (PNIs) are frequently encountered in clinical practice [222]. Short distant defected peripheral nerve can regenerate by repairing and suturing, but the repair of long distant nerve defect is still a global problem [223]. Currently, the autologous nerve graft is the gold standard treatment for long-distance nerve defect [223]. However, some challenges remain pendent for PNIs repair, such as sensory loss, scarring, and neuroma formation at the donor site [135, 167]. Moreover, size and fascicle mismatch, tubercle, and fibrosis also limit their applications in neuroregeneration [167]. To solve these problems, gene therapy has been applied in neuroregeneration and reported for nerve regeneration [224]. It is imminent to develop efficient biomaterial-based gene delivery vectors for high transfection efficiency of neuronal cells [225]. An overview of studies about neuronal cell transfection approaches, nonviral-vectors and their transfection efficiencies are summarized in Table 4.

Table 4 Nonviral-mediated delivery of nucleic acids for nerve system repair

Liposomes as a commercial reagent work through the effects of cationic lipid molecules [236]. The molecules containing positive-charge head groups can interact with negative-charge nucleic acids to form the complexes [237, 238]. Then, the complexes fuse with the cell membrane and efficiently deliver the nucleic acids into cells for gene therapy [85]. Liposomes have low toxicity and stable transfection ability for gene delivery, and may be further applied for neuron regeneration and repair in vivo [89, 239].

Cationic polymers have caused intense attention in nerve regeneration [240]. Shi and collaborators have studied the effect of intrathecal administration of pDNA/PEI complex into the lumbar subarachnoid space and the result is satisfactory for high efficient gene delivery [226]. Some researchers have proposed the modification of PEI by polyethylene glycol (i.e. PEGylation) could improve biocompatibility [241]. Song and collaborators have reported a biodegradable and biocompatible cationic polymer (PDA/PEI) by cross-linking low molecular weight PEI (1.8 kDa) with 2,6-pyridinedicarboxaldehyde [242]. During the transfection process, PDA/PEI shows a lower cytotoxicity and higher transfection efficiency than PEI (25 kDa) in transfecting miR-221/222 into rat Schwann cells (SCs) [227]. Due to the toxicity of PEI and its modified polymers, natural polysaccharides (i.e. chitosan) are chosen as gene vectors for gene delivery and therapy [243]. Chitosan with biocompatibility, biodegradability and low toxicity have attended more and more attraction in gene delivery for nerve regeneration [244, 245]. Mittnacht and collaborators have fabricated the polymer filaments loaded with chitosan/siRNA nanoparticles to promote nerve regeneration. There is a reduction of 65–75% of targeted mRNA and neurite outgrowth is improved [228]. Scaffold-mediated nonviral gene delivery for nerve regeneration also takes an important position in regulating gene therapy [243]. Li and collaborators have incorporated the microRNA-222 loading chitosan NPs (miR-222/CS NPs) with silk fibroin (SF) nanofibrous scaffolds to enhance neuronal differentiation of neural stem cells [229]. Sustained release and delivery of miR-222 in SF-based scaffold can improve the chance of successful gene transfection.

Gene therapy for cancer-resected tissue repair

Cancer gene therapy

Cancer is an unsolved worldwide problem due to high morbidity and mortality [246]. Cancer gene therapy refers to the delivery of therapeutic genes to target cells through appropriate carriers to precisely control expression up-regulation/down-regulation of specific tumor-related genes, so as to achieve efficient cancer treatment [247, 248]. The regulation process of gene and protein expression using RNA/nucleic acid in cancer therapy is essential for cancer-resected tissue repair (Fig. 8). Once genes are delivered into the targeting cells, RNA macromolecules can utilize diverse intracellular molecules to adjust gene and protein expression [249]. Then, the tumor suppressor gene is highly expressed and acts on cells to induce tumor cell apoptosis for cancer gene therapy. After the treatment of cancers, a tissue defect is exposed and it is difficult to heal damaged tissue [250, 251]. Therefore, gene therapy for cancer-resected tissue repair is demanded to coordinate tumor treatment and tissue regeneration [252, 253].

To enhance the cancer therapy efficiency and tissue repair ability, many researchers pay more attention on the inorganic NPs as gene delivery vectors for gene therapy [254]. Cristofolini and collaborators have fabricated a multifunctional hybrid NPs as magnetic delivery system for siRNA targeting the HER2 gene in breast cancer cells for efficient gene therapy and wound healing [255]. Kara and collaborators have also developed a novel poly-L-lysine (PLL)-modified sericin-coated superparamagnetic iron oxide NPs (PLL/Ser-SPIONs) as siRNA carriers [256]. PLL/Ser-SPIONs can act as a safe and promising carrier candidate to deliver siRNA-based genes into cells to treat cancer cells and induce tissue development for further applications in gene therapy and tissue engineering [256]. However, the major disadvantages of inorganic NPs are non-degradable nature, and entrapment by the lungs, liver, and kidneys, which is difficult for in vivo gene therapy in clinic [257].

In addition of inorganic NPs, the commonly used positive-charge polymers including PLL, PEI, polyamide (PAMAM), and poly(β-amino ester) (PBAE), which can interact with the negative-charge phosphate groups on the main chain of RNA and form complexes into cells and realize escape of RNA-bound endosomes [258,259,260,261]. Among them, PEI serves as a classic and effective transfection agent for gene therapy in cancer and tissue regeneration [99, 101,102,103]. In addition, PEI has a high buffering capacity in a very wide pH range. Chen and collaborators have reported the cationic perfluorocarbon nanoemulsions with positive surface charge provided by a fluorinated PEI (F-PEI) [262]. The fluorinated emulsion (F-PEI@PFD) with reduced cytotoxicity can enhance cellular uptake and improve endosomal escape of the siRNA for gene delivery and therapy in cancer-resected tissue engineering.

Fig. 8
figure 8

Regulation of gene and protein expression using RNA/Nucleic acid in cancer therapy for tissue regeneration. Once delivered into the cells, RNA macromolecules can utilize diverse intracellular mechanism to control gene and protein expression in cancer cells for gene therapy. (I) Hybridization of antisense oligonucleotides (ASOs) to a targeting mRNA can result in specific inhibition of gene expression by the induction of RNase H endonuclease activity, which cleaves the mRNA-ASO heteroduplex. (II) Short interfering RNA (siRNA) is recognized by the RNA-induced silencing complex (RISC), guided by an antisense strand of the siRNA, specifically to bind and cleave targeting mRNA. (III) In vitro transcribed mRNA utilizes the protein synthesis machinery of host cells to translate the encoded genetic information into a protein. Ribosome subunits are recruited to mRNA with a cap and poly(A)-binding proteins, forming a translation initiation complex. (IV) In the clustered regularly interspaced short palindromic repeats (CRISPR-Cas9) system, co-delivery of a single guide RNA (sgRNA) together with the mRNA encoding the Cas9 DNA endonuclease allows site-specific cleavage of dsDNA, leading to the knockout of a target gene and its product. Then, this will realize cancer therapy by RNA/Nucleic acid delivery

EVs are small membrane vesicles released into the extracellular matrix and they also play an important role in gene delivery and therapy in cancer [59]. Mutifunctional exosomes are derived from different cells types, including cancer cells, dendritic cells, B cells, T cells, mast cells and epithelial cells to deliver targeting genes for cancer therapy and tissue repair [63,64,65,66]. For example, HEK293 cells with advantages of high transfection and easy to gene operation are allowed for gene manipulation on exosomes. The biggest advantage of cancer cell-derived exosomes involved in gene delivery is targeting and low immunogenicity [263]. Zhou and collaborators have reported a novel mechanism that targeting tumor exosomal circular RNA cSERPINE2 suppresses breast cancer progression by mediating MALT1-NF-𝜅B-IL-6 axis of tumor-associated macrophages [264]. Ellipilli and collaborators have also reported GalNAc decorated exosomes as cargo for targeted delivery of Paclitaxel (PTX) and miR122 to liver tumors as an effective means to inhibit the hepatocellular carcinoma [265]. Because of the excellent performance, it has also been applied in cancer clinical trials. Researchers can consult related data on the database www.ClinicalTrials.gov (accessed in April, 2022) [266, 267]. However, some limitations are still remained for cancer-induced tissue repair, such as pharmacokinetic characteristics and potential safety issues [268,269,270].

Protein or peptide with unique sequences can demonstrate a variety of properties, including siRNA binding, membrane penetration, endosomal destruction, targeting, etc [271]. Cell-penetrating peptides (CPPS) are amphiphilic oligomers with positive-charge residues (arginine or lysine) and it is usually applied in gene delivery [272]. Deng and collaborators have reported a transdermal peptide, #PKU12. The antitumor effect of PKU12-based siRNA against HPV is perfect for human papillomavirus in vivo [273]. Protein or polypeptide delivery RNA strategies have high selectivity and activity in gene therapy and tissue repair.

Cancer-resected gene therapy for tissue repair

Fig. 9
figure 9

Regulation of gene and protein expression using RNA/Nucleic acid in cancer therapy for tissue regeneration. (A) A schematic diagram of gene therapy in tissue repair after removing tumors. Reproduced with permission [274]. Copyright 2015, Elsevier. (B) A schematic diagram to depict the role of lnc030 in maintenance of BCSC stemness through regulating SQLE and cholesterol accumulation to stimulate PI3K/Akt signaling. (C) Intracellular cholesterol is measured in lnc030-knockdown (left) or SQLE-knockdown (right) Hs578T, BT549, MCF-7 derived spheres and their control spheres. (D) Tumor growth curve and tumor weight of each mouse group are measured. (B-D) Reproduced with permission [275]. Copyright 2020, Wiley. (E) Schematic illustration of therapeutic process of ADSC-loaded GO-GA-polymer scaffold with pH-triggered dual release of GO and GA. (F) The tumor growth curves after implantation of different formulations with and without NIR irradiation. (G) Histological analysis and histomorphometric measurements of adipose tissue regeneration in vivo. (E-G) Reproduced with permission [276]. Copyright 2019, Wiley. (H) A schematic diagram of nanocomposite multifunctional hydrogel for suppressing osteosarcoma recurrence and enhancing bone regeneration. Reproduced with permission [274]. Copyright 2022, Elsevier

As mentioned before, tissue defects after tumor surgery are difficult to repair [277]. For breast cancer patients, surgical removal of tumor tissue usually results in permanent breast defects, which brings great psychological pain to the patients [278]. Therefore, gene therapy plays an important role to induce post-cancer repair and tissue regeneration (Fig. 9A) [279]. Liu and collaborators report that a long non-coding RNA lnc030, cooperated with poly(rC) binding protein 2(PCBP2) to stabilize squalene epoxidase (SQLE) mRNA, resulting in an increase of cholesterol synthesis (Fig. 9B) [275]. The increased cholesterol also activates PI3K/Akt signaling and governs breast cancer stem cells stemness. Therefore, the lnc030-SQLE-cholesterol synthesis pathway is a therapeutic target for breast cancer treatment and adipose tissue regeneration (Fig. 9C, D) [275].

In addition, advanced functional scaffolds are also applied for cancer-resected tissue regeneration. Chen and collaborators fabricate a stimuli-responsive scaffold for breast cancer treatment and adipose tissue regeneration (Fig. 9E). This scaffold is composed of polyacrylic acid-g-polylactic acid (PAA-g-PLLA) modified graphene oxide (GO) with a cleavable bond in between (GO-PAA-g-PLLA), gambogic acid (GA), and polycaprolactone (PCL). These complex scaffolds can not only selectively induce the apoptosis of tumor cells, but also improve the capability to stimulate the differentiation of ADSCs into adipocytes (Fig. 9F, G) [276]. Chen and collaborators have also fabricated a composite scaffold of gelatin and Fe3O4 NPs for magnetic hyperthermia-based breast cancer treatment and adipose tissue regeneration. It can ablate breast cancer cell and facilitate the growth and adipogenesis of mesenchymal stem cells [280]. Currently, photothermal therapy is the commonly used method for most scaffolds for breast tumor treatment and adipose tissue regeneration [281]. If gene therapy is integrated into these scaffolds, better results will be achieved for adipose regeneration. The scaffolds with RNA lnc030 integration can realize breast cancer treatment and further contribute to the adipose tissue regeneration.

Similarly, bone tumor is clinically a malignant tumor and surgical resection is the common clinical treatment [282]. Although native bone tissue has the self-regeneration ability, inevitable damage to the bone is still unsettled [283, 284]. Therefore, gene therapy would be a preferable strategy to introduce targeting genes into cells for bone tumor treatment and bone tissue regeneration. Zhang and collaborators have fabricated a functional nanovector to deliver astrocyte elevated gene-1 (AEG-1) small interfering RNA (siRNA; siAEG-1) into osteosarcoma cells to silence the targeted gene both in vitro and in vivo. The results show that the Cs-g-PLLD-FA/siAEG-1 nanocomplexes could knockdown AEG-1 genes to inhibit the tumor cell proliferation, invasion and lung metastasis in tumor-bearing mice with low cytotoxicity and high efficacy [285]. Nasr-Esfahani and collaborators also prepare mesoporous bioactive glass (MBG) sub-micro particles to present the potential in treatment of osteosarcoma by releasing bioactive ions of Ca, Si and P, and contribute to bone regeneration by depositing hydroxylapatite (HA) [286]. Lai and collaborators demonstrate that multifunctional nanocomposite hydrogel has an excellent effect for suppressing osteosarcoma recurrence and enhancing bone regeneration (Fig. 9H) [274].

At present, many researchers have proposed the gene therapy method to treat bone tumor by preparing high-efficacy and low-toxicity vectors to discover novel long non-coding RNA, circular RNA or miRNA [287, 288]. Hu and collaborators have even mapped the single-cell RNA landscape of intratumoral heterogeneity and immunosuppressive microenvironment in advanced osteosarcoma, which provides potential therapeutic targets for osteosarcoma [288]. Moreover, there are also many novel methods to treat bone tumor and contribute to bone regeneration by fabricating multifunctional scaffolds or vectors [274, 289, 290]. Combining gene therapy vectors and bioactive scaffolds, advanced biomaterials will provide a promising insight for synergetic tumor treatment and tissue regeneration (Table 5).

Table 5 Cancer-resected gene therapy for tissue repair

Conclusions and future perspectives

Tissue repair and regeneration are long-term concerns for postoperative patients, especially after orthopedics, neurology, and cancer operations. In order to alleviate the burden on patients and expedite recovery, regional gene therapy has been developed and is being considered as a potential approach for tissue regeneration and repair. Over the past two decades, regional gene therapy has been rapidly developed for tissue regeneration, and encompassing advances in vector types, delivery methods, as well as the design and fabrication of novel materials are also prosperous for efficient gene delivery and therapy in bone regeneration, cartilage regeneration, blood vessel regeneration, nerve regeneration and even cancer-resected tissue repair. Although gene therapy has shown the explosive development for tissue regeneration and repair, current mainstream programmes for regional gene therapy rely on viral-based vectors. Despite their inherent advantages such as extended duration and high payload capacity, there are still concerns about their safety, particularly in terms of immunogenicity. To completely overcome potential risks from viral and further improve gene delivery ability, various gene delivery biomaterials have been exploited for tissue regeneration and repair, including lipids, cationic polymers, peptides or proteins, dendritic/branched biomaterials, inorganic NPs and their composites. These novel gene carriers can not only enhance targeting accuracy and delivery ability to replace defective genes in damaged sites, but also emphasize the creation of new drugs and treatment procedures that are effective and less painful to eradicate all kinds of genetic tissue diseases. To date, gene therapy is still considered as the most potential solution to tissue regeneration and repair.

Improving the transfection efficiency, enhancing the transfection stability and reducing the toxicity are the common development direction of these biomaterials-based nonviral vectors for gene therapy and tissue regeneration. It is essential to further enhance the comprehension of the molecular mechanisms underlying tissue reconstruction and the effects of polymeric biomaterials on cell fate. Maximizing the transfection capacity of polymeric biomaterials in gene therapy remains a challenge, as currently, nonviral vectors are unable to achieve the same high transfection efficiency as viral vectors. In addition, although polymeric biomaterials offer desired release kinetics and spatially defined architectures, the optimal release time and the dose of therapeutic genes required for effective cartilage repair are still unknown. With the development of advanced biofabrication techniques, the combination of various biomaterials to construct composites with tailorable properties and spatially controlled biological function will be realized, and more personalized polymeric biomaterial-mediated gene delivery systems should be explored to promote clinical treatment effect. The precise regulation of each stage matching to the process of tissue formation is still hardly achieved at present. Therefore, advanced biomaterial-mediated gene delivery with efficient, well-defined and targeting transduction system will represent the next frontier to enhance tissue regeneration and repair in vivo.

Data availability

The data that support the findings of this study are available from the corresponding authors upon reasonable request.

References

  1. Zhang Z, et al. Native tissue-based strategies for meniscus repair and regeneration. Cell Tissue Res. 2018;373:337–50.

    Article  PubMed  Google Scholar 

  2. Salhotra A, et al. Mechanisms of bone development and repair. Nat Rev Mol Cell Biol. 2020;21:696–711.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Tzahor E, Poss KD. Cardiac regeneration strategies: staying young at heart. Science. 2017;356:1035–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Rana D, et al. Development of decellularized scaffolds for stem cell-driven tissue engineering. J Tissue Eng Regen Med. 2017;11:942–65.

    Article  CAS  PubMed  Google Scholar 

  5. Nerem RM, et al. Tissue engineering in the USA. Med Biol Eng Comput. 1992;30:CE8–12.

    Article  CAS  PubMed  Google Scholar 

  6. Khademhosseini A, et al. A decade of progress in tissue engineering. Nat Protoc. 2016;11:1775–81.

    Article  CAS  PubMed  Google Scholar 

  7. Berthiaume F, et al. Tissue engineering and regenerative medicine: history, progress, and challenges. Annu Rev Chem Biomol Eng. 2011;2:403–30.

    Article  PubMed  Google Scholar 

  8. Griffith LG, et al. Tissue engineering–current challenges and expanding opportunities. Science. 2002;295:1009–14.

    Article  CAS  PubMed  Google Scholar 

  9. Hassanzadeh P, et al. Tissue engineering: still facing a long way ahead. J Control Release. 2018;279:181–97.

    Article  CAS  PubMed  Google Scholar 

  10. Xia H, et al. Tissue repair and regeneration with endogenous stem cells. Nat Rev Mater. 2018;3:174–93.

    Article  CAS  Google Scholar 

  11. Ferrari S, et al. Genetic engineering meets hematopoietic stem cell biology for next-generation gene therapy. Cell Stem Cell. 2023;30:549–70.

    Article  CAS  PubMed  Google Scholar 

  12. Fox T, et al. Access to gene therapy for rare diseases when commercialization is not fit for purpose. Nat Med. 2023;29:518–9.

    Article  CAS  PubMed  Google Scholar 

  13. Birch SM, et al. Assessment of systemic AAV-microdystrophin gene therapy in the GRMD model of Duchenne muscular dystrophy. Sci Transl Med. 2023;15:eabo1815.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Kaczmarek JC, et al. Advances in the delivery of RNA therapeutics: from concept to clinical reality. Genome Med. 2017;9:60.

    Article  PubMed  PubMed Central  Google Scholar 

  15. Wang Y, et al. Is viral vector gene delivery more effective using biomaterials? Adv Healthc Mater. 2021;10:2001238.

    Article  CAS  Google Scholar 

  16. Yin H, et al. Non-viral vectors for gene-based therapy. Nat Rev Genet. 2014;15:541–55.

    Article  CAS  PubMed  Google Scholar 

  17. Gao K, et al. Nonviral methods for siRNA delivery. Mol Pharm. 2009;6:651–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Finer M, et al. A brief account of viral vectors and their promise for gene therapy. Gene Ther. 2017;24:1–2.

    Article  CAS  PubMed  Google Scholar 

  19. Chien YH, Lee NC, Tseng SH, Tai CH, Muramatsu SI, Byrne BJ, et al. Efficacy and safety of AAV2 gene therapy in children with aromatic L-amino acid decarboxylase deficiency: an open-label, phase 1/2 trial. Lancet Child Adolesc. 2017;1:265–73.

    Article  Google Scholar 

  20. Fischer MD, Ochakovski GA, Beier B, Seitz IP, Vaheb Y, Kortuem C, et al. Efficacy and safety of retinal gene therapy using adeno-associated virus vector for patients with choroideremia: a randomized clinical trial. JAMA Ophthalmol. 2019;137:1247–54.

    Article  PubMed  PubMed Central  Google Scholar 

  21. Chen S-H, et al. Recombinant viral vectors as neuroscience tools. Curr Protoc Neurosci. 2019;87:e67.

    Article  PubMed  PubMed Central  Google Scholar 

  22. Choi J-H, et al. Optimization of AAV expression cassettes to improve packaging capacity and transgene expression in neurons. Mol Brain. 2014;7:17.

    Article  PubMed  PubMed Central  Google Scholar 

  23. Davidson BL, et al. Viral vectors for gene delivery to the nervous system. Nat Rev Neurosci. 2003;4:353–64.

    Article  CAS  PubMed  Google Scholar 

  24. Flotte TR. Gene therapy progress and prospects: recombinant adeno-associated virus (rAAV) vectors. Gene Ther. 2004;11:805–10.

    Article  CAS  PubMed  Google Scholar 

  25. Wang Y, et al. Characterisation of systemic dissemination of nonreplicating adenoviral vectors from tumours in local gene delivery. Br J Cancer. 2005;92:1414–20.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Lagan A, Shasha D, Croce CM. Synthetic RNAs for gene regulation: design principles and computational tools. Front Bioeng Biotechnol. 2014;2:65.

  27. Buck J, et al. Lipid-based DNA therapeutics: hallmarks of non-viral gene delivery. ACS Nano. 2019;13:3754–82.

    Article  CAS  PubMed  Google Scholar 

  28. Anderson DG, et al. Semi-automated synthesis and screening of a large library of degradable cationic polymers for gene delivery. Angew Chem Int Ed. 2003;42:3153–8.

    Article  CAS  Google Scholar 

  29. Yan H. Nucleic acid nanotechnology. Science. 2004;306:2048–9.

    Article  CAS  PubMed  Google Scholar 

  30. Kinard LA, et al. Bone morphogenetic protein-2 release from composite hydrogels of oligo(poly(ethylene glycol) fumarate) and gelatin. Pharm Res. 2013;30:2332–43.

    Article  CAS  PubMed  Google Scholar 

  31. Elangovan S, et al. The enhancement of bone regeneration by gene activated matrix encoding for platelet derived growth factor. Biomaterials. 2014;35:737–47.

    Article  CAS  PubMed  Google Scholar 

  32. Odabas S, Feichtinger GA, Korkusuz P, Inci I, Bilgic E, Yar AS, et al. Auricular cartilage repair using cryogel scaffolds loaded with BMP-7-expressing primary chondrocytes. J Tissue Eng Regen Med. 2013;7(10):831–40.

  33. El-Aneed A. Current strategies in cancer gene therapy. Eur J Pharmacol. 2004;498:1–8.

    Article  CAS  PubMed  Google Scholar 

  34. Hanz S, et al. Integration of retrograde axonal and nuclear transport mechanisms in neurons: implications for therapeutics. Neuroscientist. 2004;10:404–8.

    Article  CAS  PubMed  Google Scholar 

  35. Zielińska A, et al. Scaffolds for drug delivery and tissue engineering: the role of genetics. JOCR. 2023;359:207–23.

    Google Scholar 

  36. Mulligan RC. The basic science of gene therapy. Science. 1993;260:926–32.

    Article  CAS  PubMed  Google Scholar 

  37. Petit C, Bonnet C, Safieddine S. Deafness: from genetic architecture to gene therapy. Nat Rev Genet. 2023;24:665–86.

  38. Dunbar CE, et al. Gene therapy comes of age. Science. 2018;359:eaan4672.

    Article  PubMed  Google Scholar 

  39. Juengst ET. The NIH points to consider and the limits of human gene therapy. Hum Gene Ther. 1990;1:425–33.

    Article  CAS  PubMed  Google Scholar 

  40. Naldini L. Gene therapy returns to centre stage. Nature. 2015;526:351–60.

    Article  CAS  PubMed  Google Scholar 

  41. van Overbeeke E, et al. Market access of gene therapies across Europe, USA, and Canada: challenges, trends, and solutions. Drug Discov Today. 2021;26:399–415.

    Article  PubMed  Google Scholar 

  42. Shukla V, et al. The landscape of cellular and gene therapy products: authorization, discontinuations, and cost. Hum Gene Ther Clin Dev. 2019;30:102–13.

    Article  CAS  PubMed  Google Scholar 

  43. Somia N, et al. Gene therapy: trials and tribulations. Nat Rev Genet. 2000;1:91–9.

    Article  CAS  PubMed  Google Scholar 

  44. Ginn SL, et al. Gene therapy clinical trials worldwide to 2017: an update. J Gene Med. 2018;20:e3015.

    Article  PubMed  Google Scholar 

  45. Wanisch K, et al. Integration-deficient lentiviral vectors: a slow coming of age. Mol Ther. 2009;17:1316–32.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Bleiziffer O, et al. Gene transfer strategies in tissue engineering. J Cell Mol Med. 2007;11:206–23.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Rizk M, et al. Patisiran for the treatment of patients with familial amyloid polyneuropathy. Drugs Today. 2019;55:315–27.

    Article  CAS  Google Scholar 

  48. Baden LR, et al. Efficacy and safety of the mRNA-1273 SARS-CoV-2 vaccine. N Engl J Med. 2021;384:403–16.

    Article  CAS  PubMed  Google Scholar 

  49. Polack FP, et al. Safety and efficacy of the BNT162b2 mRNA Covid-19 vaccine. N Engl J Med. 2020;383:2603–15.

    Article  CAS  PubMed  Google Scholar 

  50. Rojas LA, et al. Personalized RNA neoantigen vaccines stimulate T cells in pancreatic cancer. Nature. 2023;618:144–50.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Zuckerman JE, et al. Correlating animal and human phase Ia/Ib clinical data with CALAA-01, a targeted, polymer-based nanoparticle containing siRNA. PNAS. 2014;111:11449–54.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Han X, et al. Biomaterial-assisted biotherapy: a brief review of biomaterials used in drug delivery, vaccine development, gene therapy, and stem cell therapy. Bioact Mater. 2022;17:29–48.

    CAS  PubMed  PubMed Central  Google Scholar 

  53. Han S, Mahato RI, Sung YK, Kim SW. Development of biomaterials for gene therapy. Mol Ther. 2000;2:302–17.

    Article  CAS  PubMed  Google Scholar 

  54. Wong SY, et al. Polymer systems for gene delivery—past, present, and future. Prog Polym Sci. 2007;32:799–837.

    Article  CAS  Google Scholar 

  55. Dobson J. Gene therapy progress and prospects: magnetic nanoparticle-based gene delivery. Gene Ther. 2006;13:283–7.

    Article  CAS  PubMed  Google Scholar 

  56. Nayerossadat N, Maedeh T, Ali PA. Viral and nonviral delivery systems for gene delivery. Adv Biomed Res. 2012;1:27.

  57. Lostalé-Seijo I, et al. Synthetic materials at the forefront of gene delivery. Nat Rev Chem. 2018;2:258–77.

    Article  Google Scholar 

  58. Sano A, et al. Atelocollagen for protein and gene delivery. Adv Drug Deliv Rev. 2003;55:1651–77.

    Article  CAS  PubMed  Google Scholar 

  59. Cecchin R, Troyer Z, Witwer k, Morris KV. Extracellular vesicles: the next generation in gene therapy delivery. Mol Ther. 2023;31:1225–30.

  60. Angela Barba A, et al. Novel lipid and polymeric materials as delivery systems for nucleic acid based drugs. Curr Drug Metab. 2015;16:427–52.

    Article  Google Scholar 

  61. Vader P, et al. Extracellular vesicles for drug delivery. Adv Drug Deliv Rev. 2016;106:148–56.

    Article  CAS  PubMed  Google Scholar 

  62. Elsharkasy OM, et al. Extracellular vesicles as drug delivery systems: why and how? Adv Drug Deliv Rev. 2020;159:332–43.

    Article  CAS  PubMed  Google Scholar 

  63. Tan A, et al. Exosomes as nano-theranostic delivery platforms for gene therapy. Adv Drug Deliv Rev. 2013;65:357–67.

    Article  CAS  PubMed  Google Scholar 

  64. Azmi AS, et al. Exosomes in cancer development, metastasis, and drug resistance: a comprehensive review. Cancer Metastasis Rev. 2013;32:623–42.

    Article  CAS  PubMed  Google Scholar 

  65. Vlassov AV, et al. Exosomes: current knowledge of their composition, biological functions, and diagnostic and therapeutic potentials. Biochim Biophys Acta Gen Subj. 2012;1820:940–8.

    Article  CAS  Google Scholar 

  66. Mathivanan S, et al. Exosomes: extracellular organelles important in intercellular communication. J Proteom. 2010;73:1907–20.

    Article  CAS  Google Scholar 

  67. Hudry E, et al. Exosome-associated AAV vector as a robust and convenient neuroscience tool. Gene Ther. 2016;23:380–92.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Ahmadi SE, et al. Viral vectors and extracellular vesicles: innate delivery systems utilized in CRISPR/Cas-mediated cancer therapy. Cancer Gene Ther. 2023;30:936–54.

    Article  CAS  PubMed  Google Scholar 

  69. Alvarez-Erviti L, et al. Delivery of siRNA to the mouse brain by systemic injection of targeted exosomes. Nat Biotechnol. 2011;29:341–5.

    Article  CAS  PubMed  Google Scholar 

  70. Munagala R, et al. Exosome-mediated delivery of RNA and DNA for gene therapy. Cancer Lett. 2021;505:58–72.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  71. Gong L, et al. CD44-targeting drug delivery system of exosomes loading forsythiaside a combats liver fibrosis via regulating NLRP3‐mediated pyroptosis. Adv Healthc Mater. 2023;12:2202228.

    Article  CAS  Google Scholar 

  72. Duan L, et al. Exosome-mediated delivery of gene vectors for gene therapy. Nanoscale. 2021;13:1387–97.

    Article  CAS  PubMed  Google Scholar 

  73. Liang Y, Xu X, Li X, Xiong J, Li B, Duan L, et al. Chondrocyte-targeted MicroRNA delivery by engineered exosomes toward a cell-free osteoarthritis therapy. ACS Appl Mater Interfaces. 2020;12:36938–47.

    Article  CAS  PubMed  Google Scholar 

  74. Gori JL, Hsu PD, Maeder ML, Shen S, Welstead GG, Bumcrot D. Delivery and specificity of CRISPR/Cas9 genome editing technologies for human gene therapy. Hum Gene Ther. 2015;26:443–51.

    Article  CAS  PubMed  Google Scholar 

  75. Kim SM, et al. Cancer-derived exosomes as a delivery platform of CRISPR/Cas9 confer cancer cell tropism-dependent targeting. J Control Release. 2017;266:8–16.

    Article  CAS  PubMed  Google Scholar 

  76. Kallenbach NR, Rong-Ine Ma, Seeman NC. An immobile nucleic acid junction constructed from oligonucleotides. Nature. 1983;305:829–31.

    Article  CAS  Google Scholar 

  77. Rothemund PWK. Folding DNA to create nanoscale shapes and patterns. Nature. 2006;440:297–302.

    Article  CAS  PubMed  Google Scholar 

  78. Chen J, et al. Synthesis from DNA of a molecule with the connectivity of a cube. Nature. 1991;350:631–3.

    Article  CAS  PubMed  Google Scholar 

  79. Andersen ES, et al. Self-assembly of a nanoscale DNA box with a controllable lid. Nature. 2009;459:73–6.

    Article  CAS  PubMed  Google Scholar 

  80. Zadegan RM, et al. Construction of a 4 zeptoliters switchable 3D DNA box origami. ACS Nano. 2012;6:10050–3.

    Article  CAS  PubMed  Google Scholar 

  81. Li SD, Huang L. Nonviral gene therapy: promises and challenges. Gene Ther. 2000;7:31–4.

    Article  CAS  PubMed  Google Scholar 

  82. Albertsen CH, et al. The role of lipid components in lipid nanoparticles for vaccines and gene therapy. Adv Drug Deliv Rev. 2022;188:114416.

    Article  Google Scholar 

  83. Cullis PR, et al. Lipid nanoparticle systems for enabling gene therapies. Mol Ther. 2017;25:1467–75.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  84. del Pozo-Rodriguez A, et al. Applications of lipid nanoparticles in gene therapy. Eur J Pharm Biopharm. 2016;109:184–93.

    Article  PubMed  Google Scholar 

  85. Dalby B, et al. Advanced transfection with lipofectamine 2000 reagent: primary neurons, siRNA, and high-throughput applications. Methods. 2004;33:95–103.

    Article  CAS  PubMed  Google Scholar 

  86. Kulkarni JA, et al. Lipid nanoparticles enabling gene therapies: from concepts to clinical utility. Nucleic Acid Ther. 2018;28:146–57.

    Article  CAS  PubMed  Google Scholar 

  87. Li L, et al. Multifunctional nucleus-targeting nanoparticles with ultra-high gene transfection efficiency for in vivo gene therapy. Theranostics. 2017;7:1633.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  88. Williams DJ, Puhl HL, Ikeda SR. A simple, highly efficient method for heterologous expression in mammalian primary neurons using cationic lipid-mediated mRNA transfection. Front Neurosci. 2010;4:181.

  89. Tönges L, et al. Stearylated octaarginine and artificial virus-like particles for transfection of siRNA into primary rat neurons. RNA. 2006;12:1431–8.

    Article  PubMed  PubMed Central  Google Scholar 

  90. Kranz LM, et al. Systemic RNA delivery to dendritic cells exploits antiviral defence for cancer immunotherapy. Nature. 2016;534:396–401.

    Article  PubMed  Google Scholar 

  91. Rajala A, et al. Nanoparticle-assisted targeted delivery of eye-specific genes to eyes significantly improves the vision of blind mice in vivo. Nano Lett. 2014;14:5257–63.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  92. Li M, et al. Composition design and medical application of liposomes. Eur J Med Chem. 2019;164:640–53.

    Article  CAS  PubMed  Google Scholar 

  93. De Smedt SC, et al. Cationic polymer based gene delivery systems. Pharm Res. 2000;17:113–26.

    Article  PubMed  Google Scholar 

  94. Zhang S, et al. Cationic compounds used in lipoplexes and polyplexes for gene delivery. J Control Release. 2004;100:165–80.

    Article  CAS  PubMed  Google Scholar 

  95. Chen H, Guo L, Ding J, Zhou W, Qi Y. A general and efficient strategy for gene delivery based on tea polyphenols intercalation and self-polymerization. Adv Sci. 2023:2302620.

  96. Zhou Z, et al. Nonviral cancer gene therapy: delivery cascade and vector nanoproperty integration. Adv Drug Deliv Rev. 2017;115:115–54.

    Article  CAS  PubMed  Google Scholar 

  97. Zeng M, et al. Efficient and robust highly branched poly (β-amino ester)/minicircle COL7A1 polymeric nanoparticles for gene delivery to recessive dystrophic epidermolysis bullosa keratinocytes. ACS Appl Mater Interfaces. 2019;11:30661–72.

    Article  CAS  PubMed  Google Scholar 

  98. Borchard G. Chitosans for gene delivery. Adv Drug Deliv Rev. 2001;52:145–50.

    Article  CAS  PubMed  Google Scholar 

  99. Boussif O, et al. A versatile vector for gene and oligonucleotide transfer into cells in culture and in vivo: polyethylenimine. PNAS. 1995;92:7297–301.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  100. Jäger M, et al. Branched and linear poly (ethylene imine)-based conjugates: synthetic modification, characterization, and application. Chem Soc Rev. 2012;41:4755–67.

    Article  PubMed  Google Scholar 

  101. Qin L, et al. Gold nanorod-assembled ZnGa 2 O 4: cr nanofibers for LED-amplified gene silencing in cancer cells. Nanoscale. 2018;10:13432–42.

    Article  CAS  PubMed  Google Scholar 

  102. Laroui H, et al. Fab’-bearing siRNA TNFα-loaded nanoparticles targeted to colonic macrophages offer an effective therapy for experimental colitis. J Control Release. 2014;186:41–53.

    Article  CAS  PubMed  Google Scholar 

  103. Tian Y, et al. Polyvinylamine with moderate binding affinity as a highly effective vehicle for RNA delivery. J Control Release. 2022;345:20–37.

    Article  CAS  PubMed  Google Scholar 

  104. Godbey W, et al. Poly (ethylenimine) and its role in gene delivery. J Control Release. 1999;60:149–60.

    Article  CAS  PubMed  Google Scholar 

  105. Kievit FM, et al. PEI–PEG–chitosan-copolymer‐coated iron oxide nanoparticles for safe gene delivery: synthesis, complexation, and transfection. Adv Funct Mater. 2009;19:2244–51.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  106. Rohiwal SS, et al. Polyethylenimine based magnetic nanoparticles mediated non-viral CRISPR/Cas9 system for genome editing. Sci Rep. 2020;10:4619.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  107. Casper J, et al. Core–shell structured chitosan-polyethylenimine nanoparticles for gene delivery: improved stability, cellular uptake, and transfection efficiency. Macromol Biosci. 2023;23:2200314.

    Article  CAS  Google Scholar 

  108. Romani C, et al. Selectively fluorinated PAMAM–arginine conjugates as gene delivery vectors. Bioconjug Chem. 2023;34:1084–95.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  109. Huang HC, et al. Inorganic nanoparticles for cancer imaging and therapy. J Control Release. 2011;155:344–57.

    Article  CAS  PubMed  Google Scholar 

  110. Mitchell MJ, et al. Engineering precision nanoparticles for drug delivery. Nat Rev Drug Discov. 2021;20:101–24.

    Article  CAS  PubMed  Google Scholar 

  111. Cutler JI, et al. Spherical nucleic acids. J Am Chem Soc. 2012;134:1376–91.

    Article  CAS  PubMed  Google Scholar 

  112. Conde J, et al. Dual targeted immunotherapy via in vivo delivery of biohybrid RNAi-peptide nanoparticles to tumour-associated macrophages and cancer cells. Adv Funct Mater. 2015;25:4183–94.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  113. Conde J, et al. Implantable hydrogel embedded dark-gold nanoswitch as a theranostic probe to sense and overcome cancer multidrug resistance. PNAS. 2015;112:E1278–1287.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  114. Gilam A, et al. Local microRNA delivery targets Palladin and prevents metastatic breast cancer. Nat Commun. 2016;7:12868.

    Article  PubMed  PubMed Central  Google Scholar 

  115. Dai X, et al. Controlled synthesis and surface engineering of janus chitosan-gold nanoparticles for photoacoustic imaging-guided synergistic gene/photothermal therapy. Small. 2021;17:e2006004.

    Article  PubMed  Google Scholar 

  116. Luther DC, et al. Delivery of drugs, proteins, and nucleic acids using inorganic nanoparticles. Adv Drug Deliv Rev. 2020;156:188–213.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  117. Zhou Y, et al. Mesoporous silica nanoparticles for drug and gene delivery. Acta Pharm Sin B. 2018;8:165–77.

    Article  PubMed  PubMed Central  Google Scholar 

  118. Hussein HA,Nazir MS, Azra N, Qamar Z, Seeni A, Tengku Din TADA, Abdullah MA. Novel drug and gene delivery system and imaging agent based on marine diatom biosilica nanoparticles. Mar Drugs 2022;20(8):480.

  119. Huang XQ, et al. Hollow mesoporous zirconia delivery system for biomineralization precursors. Acta Biomater. 2018;67:366–77.

    Article  CAS  PubMed  Google Scholar 

  120. Rasool N, et al. Thiol-functionalized, antioxidant, and osteogenic mesoporous silica nanoparticles for osteoporosis. ACS Biomater Sci Eng. 2023;9:3535–45.

    Article  CAS  PubMed  Google Scholar 

  121. Mohammadpour R, et al. Subchronic and chronic toxicity evaluation of inorganic nanoparticles for delivery applications. Adv Drug Deliv Rev. 2019;144:112–32.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  122. Arami H, et al. In vivo delivery, pharmacokinetics, biodistribution and toxicity of iron oxide nanoparticles. Chem Soc Rev. 2015;44:8576–607.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  123. Nam SH, et al. Recent advances in selective and targeted drug/gene delivery systems using cell-penetrating peptides. Arch Pharm Res. 2023;46:18–34.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  124. Snyder EL, et al. Cell penetrating peptides in drug delivery. Pharm Res. 2004;21:389–93.

    Article  CAS  PubMed  Google Scholar 

  125. Wu H, et al. Cell-penetrating peptide: a powerful delivery tool for DNA-free crop genome editing. Plant Sci. 2022;324:111436.

    Article  CAS  PubMed  Google Scholar 

  126. Wang Y, et al. Overcoming the blood-brain barrier for gene therapy via systemic administration of GSH-responsive silica nanocapsules. Adv Mater. 2023;35:e2208018.

    Article  PubMed  Google Scholar 

  127. Cutroneo KR. Gene therapy for tissue regeneration. J Cell Biochem. 2003;88:418–25.

    Article  CAS  PubMed  Google Scholar 

  128. Goldman JA, et al. Gene regulatory programmes of tissue regeneration. Nat Rev Genet. 2020;21:511–25.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  129. Kusano KF, et al. Sonic hedgehog myocardial gene therapy: tissue repair through transient reconstitution of embryonic signaling. Nat Med. 2005;11:1197–204.

    Article  CAS  PubMed  Google Scholar 

  130. Bonadio J, et al. Gene therapy for tissue repair and regeneration. Adv Drug Deliv Rev. 1998;33:53–69.

    Article  CAS  PubMed  Google Scholar 

  131. Yan R, Cigliola V, Oonk KA, Petrover Z, DeLuca S, Wolfson DW, et al. An enhancer-based gene-therapy strategy for spatiotemporal control of cargoes during tissue repair. Cell Stem Cell. 2023;30:96–111.

    Article  CAS  PubMed  Google Scholar 

  132. Ikada Y. Challenges in tissue engineering. J R Soc Interface. 2006;3:589–601.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  133. Almeida GHD, et al. Uterine tissue engineering: where we stand and the challenges ahead. Tissue Eng Part B Rev. 2022;28:861–90.

    Article  PubMed  Google Scholar 

  134. Wang S, et al. Advances in translational 3D printing for cartilage, bone, and osteochondral tissue engineering. Small. 2022;18:e2201869.

    Article  PubMed  Google Scholar 

  135. Yu L, et al. Gradient degradable nerve guidance conduit with multilayer structure prepared by electrospinning. Mater Lett. 2020;276:128238.

    Article  CAS  Google Scholar 

  136. Cai J, Wang J, Sun C, Dai J, Zhang C. Biomaterials with stiffness gradient for interface tissue engineering. Biomed Mater 2022;17.

  137. Wang Y, et al. Collagen-based biomaterials for tissue engineering. ACS Biomater Sci Eng. 2023;9:1132–50.

    Article  CAS  PubMed  Google Scholar 

  138. Forbes SJ, et al. Preparing the ground for tissue regeneration: from mechanism to therapy. Nat Med. 2014;20:857–69.

    Article  CAS  PubMed  Google Scholar 

  139. Zhuang Y, et al. Biomaterial-based delivery of nucleic acids for tissue regeneration. Adv Drug Deliv Rev. 2021;176:113885.

    Article  CAS  PubMed  Google Scholar 

  140. Kwon SG, et al. Recent advances in stem cell therapeutics and tissue engineering strategies. Biomater Res. 2018;22:1–8.

    Article  Google Scholar 

  141. Wang Y, et al. Micropattern-controlled cell density and its effect on gene transfection of mesenchymal stem cells. Adv Mater Interfaces. 2022;9:2101978.

    Article  Google Scholar 

  142. Wang Y, et al. Regulation of gene transfection by cell size, shape and elongation on micropatterned surfaces. J Mater Chem B. 2021;9:4329–39.

    Article  CAS  PubMed  Google Scholar 

  143. Wang Y, et al. The varied influences of cell adhesion and spreading on gene transfection of mesenchymal stem cells on a micropatterned substrate. Acta Biomater. 2021;125:100–11.

    Article  CAS  PubMed  Google Scholar 

  144. Zhang W, et al. A PDA-functionalized 3D lung scaffold bioplatform to construct complicated breast tumor microenvironment for anticancer drug screening and immunotherapy. Adv Sci. 2023;10:e2302855.

    Article  Google Scholar 

  145. Dalton CJ, Lemmon CA. Fibronectin: molecular structure, fibrillar structure and mechanochemical signaling. Cells. 2021;10:2443.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  146. Singh P, et al. Assembly of fibronectin extracellular matrix. Annu Rev Cell Dev Bi. 2010;26:397–419.

    Article  CAS  Google Scholar 

  147. Campbell ID, Humphries MJ. Integrin structure, activation, and interactions. Cold Spring Harb Perspect Biol. 2011;3:a004994.

    Article  PubMed  PubMed Central  Google Scholar 

  148. Koivisto L, et al. Integrin αvβ6: structure, function and role in health and disease. Int J Biochem Cell Biol. 2018;99:186–96.

    Article  CAS  PubMed  Google Scholar 

  149. Wang Y, et al. Regulation of micropatterned curvature-dependent FA heterogeneity on cytoskeleton tension and nuclear DNA synthesis of malignant breast cancer cells. J Mater Chem B. 2023;11:99–108.

    Article  CAS  Google Scholar 

  150. Wang Y, et al. Cellular nanomechanics derived from pattern-dependent focal adhesion and cytoskeleton to balance gene transfection of malignant osteosarcoma. J Nanobiotechnol. 2022;20:499.

    Article  CAS  Google Scholar 

  151. Sun Z, et al. Integrin activation by talin, kindlin and mechanical forces. Nat Cell Biol. 2019;21:25–31.

    Article  CAS  PubMed  Google Scholar 

  152. Zhao Y, et al. Talin–1 interaction network in cellular mechanotransduction. Int J Mol Med. 2022;49:60.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  153. Bays JL, DeMali KA. Vinculin in cell-cell and cell-matrix adhesions. Cell Mol Life Sci. 2017;74:2999–3009.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  154. Liu J, et al. Wnt/β-catenin signalling: function, biological mechanisms, and therapeutic opportunities. Signal Transduct Target Ther. 2022;7:3.

    Article  PubMed  PubMed Central  Google Scholar 

  155. Kobielak A, et al. α-catenin: at the junction of intercellular adhesion and actin dynamics. Nat Rev Mol Cell Biol. 2004;5:614–25.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  156. Takata N, et al. Stem cells and genome editing: approaches to tissue regeneration and regenerative medicine. J Hum Genet. 2018;63:165–78.

    Article  CAS  PubMed  Google Scholar 

  157. Gaharwar AK, et al. Engineered biomaterials for in situ tissue regeneration. Nat Rev Mater. 2020;5:686–705.

    Article  CAS  Google Scholar 

  158. Vo TN, et al. Strategies for controlled delivery of growth factors and cells for bone regeneration. Adv Drug Deliv Rev. 2012;64:1292–309.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  159. Kofron MD, et al. Bone tissue engineering by gene delivery. Adv Drug Deliv Rev. 2006;58:555–76.

    Article  CAS  PubMed  Google Scholar 

  160. Jin L, et al. Current progress in gene delivery technology based on chemical methods and nano-carriers. Theranostics. 2014;4:240.

    Article  PubMed  PubMed Central  Google Scholar 

  161. Kwon H, et al. Surgical and tissue engineering strategies for articular cartilage and meniscus repair. Nat Rev Rheumatol. 2019;15:550–70.

    Article  PubMed  PubMed Central  Google Scholar 

  162. Cai Y, et al. Enhanced osteoarthritis therapy by nanoengineered mesenchymal stem cells using biomimetic CuS nanoparticles loaded with plasmid DNA encoding TGF-β1. Bioact Mater. 2023;19:444–57.

    CAS  PubMed  Google Scholar 

  163. Bonato A, et al. Engineering inflammation-resistant cartilage: bridging gene therapy and tissue engineering. Adv Healthc Mater. 2023;12:e2202271.

    Article  PubMed  Google Scholar 

  164. Rosenholm JM, et al. Mesoporous silica nanoparticles in tissue engineering–a perspective. Nanomedicine. 2016;11:391–402.

    Article  CAS  PubMed  Google Scholar 

  165. Wang J, et al. Mesoporous silica nanoparticles carrying microRNA-124 to target P2Y(12) facilitates cerebral angiogenesis in lacunar cerebral infarction through stem cell factor/c-kit signaling pathway. J Biomed Nanotechnol. 2022;18:218–24.

    Article  CAS  PubMed  Google Scholar 

  166. Hu N, et al. Hypoxia-pretreated ADSC-derived exosome-embedded hydrogels promote angiogenesis and accelerate diabetic wound healing. Acta Biomater. 2023;157:175–86.

    Article  CAS  PubMed  Google Scholar 

  167. Grinsell D, et al. Peripheral nerve reconstruction after injury: a review of clinical and experimental therapies. Biomed Res Int. 2014;2014:698256.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  168. Ma Y, et al. Exosomes released from neural progenitor cells and induced neural progenitor cells regulate neurogenesis through miR-21a. Cell Commun Signal. 2019;17:96.

    Article  PubMed  PubMed Central  Google Scholar 

  169. Cao L, et al. Neural progenitor cell-derived nanovesicles promote hair follicle growth via miR-100. J Nanobiotechnol. 2021;19:20.

    Article  CAS  Google Scholar 

  170. Majidinia M, et al. The roles of signaling pathways in bone repair and regeneration. J Cell Physiol. 2018;233:2937–48.

    Article  CAS  PubMed  Google Scholar 

  171. Yudoh K, Sugishita Y, Suzuki-Takahashi Y. Bone development and regeneration 2.0. Int J Mol Sci. 2023;24:8761.

  172. Lee J, et al. Current advances in immunomodulatory biomaterials for bone regeneration. Adv Healthc Mater. 2019;8:e1801106.

    Article  PubMed  Google Scholar 

  173. Elgali I, et al. Guided bone regeneration: materials and biological mechanisms revisited. Eur J Oral Sci. 2017;125:315–37.

    Article  PubMed  PubMed Central  Google Scholar 

  174. Este MD, Eglin D, Alini M, Kyllonen L. Bone regeneration with biomaterials and active molecules delivery. Curr Pharm Biotechnol. 2015;16:582–605.

  175. Moriarty N, et al. A combined cell and gene therapy approach for homotopic reconstruction of midbrain dopamine pathways using human pluripotent stem cells. Cell Stem Cell. 2022;29:434–48.

    Article  CAS  PubMed  Google Scholar 

  176. Winn SR, et al. Gene therapy approaches for modulating bone regeneration. Adv Drug Deliv Rev. 2000;42:121–38.

    Article  CAS  PubMed  Google Scholar 

  177. Deckers R, et al. Image-guided, noninvasive, spatiotemporal control of gene expression. PNAS. 2009;106:1175–80.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  178. Collon K, et al. Musculoskeletal tissue engineering: Regional gene therapy for bone repair. Biomaterials. 2021;275:120901.

    Article  CAS  PubMed  Google Scholar 

  179. Bessa PC, et al. Bone morphogenetic proteins in tissue engineering: the road from laboratory to clinic, part II (BMP delivery). J Tissue Eng Regen Med. 2008;2:81–96.

    Article  CAS  PubMed  Google Scholar 

  180. Santos J, et al. Non-viral gene delivery to mesenchymal stem cells: methods, strategies and application in bone tissue engineering and regeneration. Curr Gene Ther. 2011;11:46–57.

    Article  PubMed  Google Scholar 

  181. Alsaggar M, et al. Physical methods for gene transfer. Adv Genet. 2015;89:1–24.

    Article  CAS  PubMed  Google Scholar 

  182. Curtin CM, et al. Combinatorial gene therapy accelerates bone regeneration: non-viral dual delivery of VEGF and BMP2 in a collagen‐nanohydroxyapatite scaffold. Adv Healthc Mater. 2015;4:223–7.

    Article  CAS  PubMed  Google Scholar 

  183. Liu S, et al. Non-coding-RNA-activated core/chitosan shell nanounits coated with polyetheretherketone for promoting bone regeneration and osseointegration via osteoimmunology. ACS Appl Mater Interfaces. 2023;15:12653–68.

    Article  CAS  PubMed  Google Scholar 

  184. Nguyen NTK, Chang YH, Truong VA, Hsu MN, Pham NN, Chang CW, et al. CRISPR activation of long non-coding RNA DANCR promotes bone regeneration. Biomaterials. 2021;275:120965.

  185. Hao L, et al. Bone targeting miR-26a loaded exosome-mimetics for bone regeneration therapy by activating wnt signaling pathway. Chem Eng J. 2023;471:144594.

    Article  CAS  Google Scholar 

  186. Steinert AF, et al. Concepts in gene therapy for cartilage repair. Injury. 2008;39:97–113.

    Article  Google Scholar 

  187. Guo X, et al. Regeneration of articular cartilage defects: therapeutic strategies and perspectives. J Tissue Eng. 2023;14:20417314231164765.

    Article  PubMed  PubMed Central  Google Scholar 

  188. Lee HP, et al. Mechanical confinement regulates cartilage matrix formation by chondrocytes. Nat Mater. 2017;16:1243–51.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  189. Huo Y, et al. In vivo stable allogenic cartilage regeneration in a goat model based on immunoisolation strategy using electrospun semipermeable membranes. Adv Healthc Mater. 2023;12:2203084.

    Article  CAS  Google Scholar 

  190. Yang R, et al. Recent advances in polymeric biomaterials-based gene delivery for cartilage repair. Bioact Mater. 2020;5:990–1003.

    PubMed  PubMed Central  Google Scholar 

  191. Gonzalez-Fernandez T, et al. Pore-forming bioinks to enable spatio-temporally defined gene delivery in bioprinted tissues. J Control Release. 2019;301:13–27.

    Article  CAS  PubMed  Google Scholar 

  192. Liu Q, et al. Inorganic nanoparticles applied as functional therapeutics. Adv Funct Mater. 2021;31:2008171.

    Article  CAS  Google Scholar 

  193. Saraf A, et al. Gene delivery strategies for cartilage tissue engineering. Adv Drug Deliv Rev. 2006;58:592–603.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  194. Madry H, et al. Thermosensitive hydrogel based on PEO–PPO–PEO poloxamers for a controlled in situ release of recombinant adeno-associated viral vectors for effective gene therapy of cartilage defects. Adv Mater. 2020;32:1906508.

    Article  CAS  Google Scholar 

  195. Dong Y, et al. Spatiotemporal regulation of endogenous MSCs using a functional injectable hydrogel system for cartilage regeneration. NPG Asia Mater. 2021;13:71.

    Article  CAS  Google Scholar 

  196. Li M, et al. The immune microenvironment in cartilage injury and repair. Acta Biomater. 2022;140:23–42.

    Article  CAS  PubMed  Google Scholar 

  197. Liu X, et al. Injectable catalyst-free click organic-inorganic nanohybrid (click-ON) cement for minimally invasive in vivo bone repair. Biomaterials. 2021;276:121014.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  198. Kim HS, Lee JH, Mandakhbayar N, Jin GZ, Kim SJ, Yoon JY, et al. Therapeutic tissue regenerative nanohybrids self-assembled from bioactive inorganic core/chitosan shell nanounits. Biomaterials. 2021;274:120857.

    Article  CAS  PubMed  Google Scholar 

  199. Hayes AJ, et al. Glycosaminoglycan and proteoglycan biotherapeutics in articular cartilage protection and repair strategies: novel approaches to visco-supplementation in orthobiologics. Adv Ther. 2019;2:1900034.

    Article  Google Scholar 

  200. Intini C, et al. An innovative mir-activated scaffold for the delivery of a miR-221 inhibitor to enhance cartilage defect repair. Adv Ther. 2023;6:2200329.

    Article  CAS  Google Scholar 

  201. Zhao RL, Zhang XM, Jia LN, Song W, Sun YL, Meng XY, Peng XX. p NNS-conjugated chitosan mediated IGF-1 and miR-140 overexpression in articular chondrocytes improves cartilage repair. Biomed Res Int. 2019;2019:2761241.

  202. Zhao R, et al. Interleukin-1 receptor antagonist protein (IL-1Ra) and miR-140 overexpression via pNNS-conjugated chitosan-mediated gene transfer enhances the repair of full-thickness cartilage defects in a rabbit model. Bone Joint Res. 2019;8:165–78.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  203. Huang J, et al. Advances in the study of exosomes as drug delivery systems for bone-related diseases. Pharmaceutics. 2023;15:220.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  204. Mao GP, et al. Exosomes derived from miR-92a-3p-overexpressing human mesenchymal stem cells enhance chondrogenesis and suppress cartilage degradation via targeting WNT5A. Stem Cell Res Ther. 2018;9:247.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  205. Xu HG, Xu B. BMSC-derived exosomes ameliorate osteoarthritis by inhibiting pyroptosis of cartilage via delivering miR-326 targeting HDAC3 and STAT1//NF-kappa B p65 to chondrocytes. Mediators Inflamm. 2021;2021:9972805.

  206. Zhu W, et al. Advances in biomaterial-mediated gene therapy for articular cartilage repair. Bioengineering. 2022;9:502.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  207. Yang J, et al. Cell-laden hydrogels for osteochondral and cartilage tissue engineering. Acta Biomater. 2017;57:1–25.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  208. Deng C et al. Application of chitosan-based biomaterials for blood vessel regeneration. Macromol Symp. 2010, 297: 138–146.

  209. Chen Y, et al. In vivo bio-safety evaluations and diagnostic/therapeutic applications of chemically designed mesoporous silica nanoparticles. Adv Mater. 2013;25:3144–76.

    Article  CAS  PubMed  Google Scholar 

  210. Paul A, et al. Injectable graphene oxide/hydrogel-based angiogenic gene delivery system for vasculogenesis and cardiac repair. ACS Nano. 2014;8:8050–62.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  211. Zheng SS, et al. Inhibition of neointimal hyperplasia in balloon-induced vascular injuries in a rat model by miR-22 loading Laponite hydrogels. Biomater Adv. 2022;142:213140.

    Article  CAS  PubMed  Google Scholar 

  212. Theoharis S, et al. Gene delivery to vascular endothelium using chemical vectors: implications for cardiovascular gene therapy. Expert Opin Biol Ther. 2007;7:627–43.

    Article  CAS  PubMed  Google Scholar 

  213. Kagaya H, et al. Impact of polyplex micelles installed with cyclic RGD peptide as ligand on gene delivery to vascular lesions. Gene Ther. 2012;19:61–9.

    Article  CAS  PubMed  Google Scholar 

  214. Zhou F, et al. Nanofiber-mediated microRNA-126 delivery to vascular endothelial cells for blood vessel regeneration. Acta Biomater. 2016;43:303–13.

    Article  CAS  PubMed  Google Scholar 

  215. Peng B, et al. MicroRNA delivery for regenerative medicine. Adv Drug Deliv Rev. 2015;88:108–22.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  216. Feghhi M, et al. Effect of multi-functional polyhydroxylated polyhedral oligomeric silsesquioxane (POSS) nanoparticles on the angiogenesis and exosome biogenesis in human umbilical vein endothelial cells (HUVECs). Mater Des. 2021;197:109227.

    Article  CAS  Google Scholar 

  217. Dang JM, et al. Natural polymers for gene delivery and tissue engineering. Adv Drug Deliv Rev. 2006;58:487–99.

    Article  CAS  PubMed  Google Scholar 

  218. Cojocaru E, et al. Electrospun-fibrous-architecture-mediated non-viral gene therapy drug delivery in regenerative medicine. Polymers. 2022;14:2647.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  219. Al-Dosari MS, et al. Nonviral gene delivery: principle, limitations, and recent progress. AAPS J. 2009;11:671–81.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  220. Carballo-Pedrares N, et al. Hydrogel-based localized nonviral gene delivery in regenerative medicine approaches—An overview. Pharmaceutics. 2020;12:752.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  221. Hu H, et al. miR-23a-3p-abundant small extracellular vesicles released from Gelma/nanoclay hydrogel for cartilage regeneration. J Extracell Vesicles. 2020;9:1778883.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  222. Campbell WW. Evaluation and management of peripheral nerve injury. Neurophysiol Clin. 2008;119:1951–65.

    Article  Google Scholar 

  223. Scheib J, et al. Advances in peripheral nerve regeneration. Nat Rev Neurol. 2013;9:668–76.

    Article  CAS  PubMed  Google Scholar 

  224. Mason RJ. Gene therapy for the peripheral nervous system: a strategy to repair the injured nerve? Curr Gene Ther. 2011;11:75–89.

    Article  CAS  PubMed  Google Scholar 

  225. Zhang N, et al. Localised non-viral delivery of nucleic acids for nerve regeneration in injured nervous systems. Exp Neurol. 2019;319:112820.

    Article  CAS  PubMed  Google Scholar 

  226. Shi L, et al. Repeated intrathecal administration of plasmid DNA complexed with polyethylene glycol-grafted polyethylenimine led to prolonged transgene expression in the spinal cord. Gene Ther. 2003;10:1179–88.

    Article  CAS  PubMed  Google Scholar 

  227. Song J, Li X, Li Y, Che J, Li X, Zhao X, et al. Biodegradable and biocompatible cationic polymer delivering microRNA-221/222 promotes nerve regeneration after sciatic nerve crush. Int J Nanomed. 2017;12:4195–208.

    Article  Google Scholar 

  228. Mittnacht U, et al. Chitosan/siRNA nanoparticles biofunctionalize nerve implants and enable neurite outgrowth. Nano Lett. 2010;10:3933–9.

    Article  CAS  PubMed  Google Scholar 

  229. Li Z, et al. Silk fibroin nanofibrous scaffolds incorporated with microRNA-222 loaded chitosan nanoparticles for enhanced neuronal differentiation of neural stem cells. Carbohyd Polym. 2022;277:118791.

    Article  CAS  Google Scholar 

  230. Uchida S, et al. In vivo messenger RNA introduction into the central nervous system using polyplex nanomicelle. PLoS ONE. 2013;8:e56220.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  231. Rong Y, et al. Engineered extracellular vesicles for delivery of siRNA promoting targeted repair of traumatic spinal cord injury. Bioact Mater. 2023;23:328–42.

    CAS  PubMed  Google Scholar 

  232. Nguyen LH, et al. Three-dimensional aligned nanofibers-hydrogel scaffold for controlled non-viral drug/gene delivery to direct axon regeneration in spinal cord injury treatment. Sci Rep. 2017;7:42212.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  233. Uchida S, et al. Treatment of spinal cord injury by an advanced cell transplantation technology using brain-derived neurotrophic factor-transfected mesenchymal stem cell spheroids. Biomaterials. 2016;109:1–11.

    Article  CAS  PubMed  Google Scholar 

  234. Tawfik M, et al. Gene therapy with caspase-3 small interfering RNA-nanoparticles is neuroprotective after optic nerve damage. Neural Regen Res. 2021;16:2534–41.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  235. Chang MF, Hsieh JH, Chiang H, Kan HW, Huang CM, Chellis L, et al. Effective gene expression in the rat dorsal root ganglia with a non-viral vector delivered via spinal nerve injection. Sci Rep. 2016;6:35612.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  236. Zou L, et al. Liposome-mediated NGF gene transfection following neuronal injury: potential therapeutic applications. Gene Ther. 1999;6:994–1005.

    Article  CAS  PubMed  Google Scholar 

  237. Kim YH, Lee K, Li S. Nucleic acids based polyelectrolyte complexes: their complexation mechanism, morphology, and stability. Chem Mater. 2021;33:7923–43.

  238. Caplen NJ. Nucleic acid transfer using cationic lipids. Gene Target Protoc. 2000;133:1–19.

    CAS  Google Scholar 

  239. Han Z, et al. Establishment of lipofection protocol for efficient miR-21 transfection into cortical neurons in vitro. DNA Cell Biol. 2015;34:703–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  240. Jiang XC, et al. Neural stem cells transfected with reactive oxygen species–responsive polyplexes for effective treatment of ischemic stroke. Adv Mater. 2019;31:1807591.

    Article  Google Scholar 

  241. Ogris M, et al. Tumor-targeted gene therapy: strategies for the preparation of ligand–polyethylene glycol–polyethylenimine/DNA complexes. J Control Release. 2003;91:173–81.

    Article  CAS  PubMed  Google Scholar 

  242. Song J, et al. Biodegradable and biocompatible cationic polymer delivering microRNA-221/222 promotes nerve regeneration after sciatic nerve crush. Int J Nanomed. 2017;12:4195–208.

    Article  CAS  Google Scholar 

  243. Raviña M, et al. Hyaluronic acid/chitosan-g-poly (ethylene glycol) nanoparticles for gene therapy: an application for pDNA and siRNA delivery. Pharm Res. 2010;27:2544–55.

    Article  PubMed  Google Scholar 

  244. Lu H, et al. Porous chitosan scaffolds with embedded hyaluronic acid/chitosan/plasmid-DNA nanoparticles encoding TGF-β1 induce DNA controlled release, transfected chondrocytes, and promoted cell proliferation. PLoS ONE. 2013;8:e69950.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  245. Malmo J, et al. Nanoparticle mediated P-glycoprotein silencing for improved drug delivery across the blood-brain barrier: a siRNA-chitosan approach. PLoS ONE. 2013;8:e54182.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  246. Højris I, et al. Morbidity and mortality of ischaemic heart disease in high-risk breast-cancer patients after adjuvant postmastectomy systemic treatment with or without radiotherapy: analysis of DBCG 82b and 82c randomised trials. Lancet. 1999;354:1425–30.

    Article  PubMed  Google Scholar 

  247. Roma-Rodrigues C, et al. Gene therapy in cancer treatment: why go nano? Pharmaceutics. 2020;12:233.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  248. Witzigmann D, et al. Lipid nanoparticle technology for therapeutic gene regulation in the liver. Adv Drug Deliver Rev. 2020;159:344–63.

    Article  CAS  Google Scholar 

  249. Caplen N. Gene therapy progress and prospects. Downregulating gene expression: the impact of RNA interference. Gene Ther. 2004;11:1241–8.

    Article  CAS  PubMed  Google Scholar 

  250. Luo Y, et al. 3D printing of hydrogel scaffolds for future application in photothermal therapy of breast cancer and tissue repair. Acta Biomater. 2019;92:37–47.

    Article  CAS  PubMed  Google Scholar 

  251. Suárez-González D, et al. Controlled nucleation of hydroxyapatite on alginate scaffolds for stem cell-based bone tissue engineering. J Biomed Mater Res A. 2010;95A:222–34.

    Article  Google Scholar 

  252. Yang P, et al. Functionalized mesoporous silica materials for controlled drug delivery. Chem Soc Rev. 2012;41:3679–98.

    Article  CAS  PubMed  Google Scholar 

  253. Lin G, et al. Inorganic nanomaterial-mediated gene therapy in combination with other antitumor treatment modalities. Adv Funct Mater. 2021;31:2007096.

    Article  CAS  PubMed  Google Scholar 

  254. Loh XJ, et al. Utilising inorganic nanocarriers for gene delivery. Biomater Sci. 2016;4:70–86.

    Article  CAS  PubMed  Google Scholar 

  255. Cristofolini T, et al. Multifunctional hybrid nanoparticles as magnetic delivery systems for siRNA targeting the HER2 gene in breast cancer cells. Mat Sci Eng C. 2020;109:110555.

    Article  CAS  Google Scholar 

  256. Kara G, et al. Development of novel poly-l-lysine-modified sericin-coated superparamagnetic iron oxide nanoparticles as siRNA carrier. Colloid Surf A. 2021;630:127622.

    Article  CAS  Google Scholar 

  257. Sokolova V, et al. Inorganic nanoparticles as carriers of nucleic acids into cells. Angew Chem Int Ed Engl. 2008;47:1382–95.

    Article  CAS  PubMed  Google Scholar 

  258. Labala S, et al. Effective melanoma cancer suppression by iontophoretic co-delivery of STAT3 siRNA and imatinib using gold nanoparticles. Int J Pharmaceut. 2017;525:407–17.

    Article  CAS  Google Scholar 

  259. Ball RL, et al. Lipid nanoparticle formulations for enhanced co-delivery of siRNA and mRNA. Nano lett. 2018;18:3814–22.

    Article  CAS  PubMed  Google Scholar 

  260. Tseng YC, Mozumdar S, Huang L. Lipid-based systemic delivery of siRNA. Adv Drug Deliv Rev. 2009;61:721–31.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  261. Subhan MA, et al. Advances in siRNA delivery strategies for the treatment of MDR cancer. Life Sci. 2021;274:119337.

    Article  CAS  PubMed  Google Scholar 

  262. Chen G, et al. Development of fluorinated polyplex nanoemulsions for improved small interfering RNA delivery and cancer therapy. Nano Res. 2018;11:3746–61.

    Article  CAS  Google Scholar 

  263. Bellavia D, et al. Interleukin 3-receptor targeted exosomes inhibit in vitro and in vivo chronic myelogenous leukemia cell growth. Theranostics. 2017;7:1333.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  264. Zhou B, Mo Z, Lai G, Chen X, Li R, Wu R, et al. Targeting tumor exosomal circular RNA cSERPINE2 suppresses breast cancer progression by modulating MALT1-NF-kB-IL-6 axis of tumor-associated macrophages. J Exp Clin Cancer Res. 2023;42:48.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  265. Ellipilli S, et al. Ligand-displaying-exosomes using RNA nanotechnology for targeted delivery of multi-specific drugs for liver cancer regression. Nanomed-Nanotechnol. 2023;50:102667.

    Article  CAS  Google Scholar 

  266. Dai S, et al. Phase I clinical trial of autologous ascites-derived exosomes combined with GM-CSF for colorectal cancer. Mol Ther. 2008;16:782–90.

    Article  CAS  PubMed  Google Scholar 

  267. Lee BC, Kang I, Yu KR. Therapeutic features and updated clinical trials of mesenchymal stem cell (MSC)-derived exosomes. J Clin Med. 2021;10:711.

    Article  PubMed  PubMed Central  Google Scholar 

  268. Raguraman R, et al. Tumor-targeted exosomes for delivery of anticancer drugs. Cancer Lett. 2023;558:216093.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  269. Kar R, et al. Exosome-based smart drug delivery tool for cancer theranostics. Acs Biomater Sci Eng. 2023;9:577–94.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  270. Wang T, et al. Exosome-based drug delivery systems in cancer therapy. Chin Chem Lett. 2023;34:107508.

    Article  CAS  Google Scholar 

  271. Copolovici DM, et al. Cell-penetrating peptides: design, synthesis, and applications. ACS Nano. 2014;8:1972–94.

    Article  CAS  PubMed  Google Scholar 

  272. Torchilin VP. Cell penetrating peptide-modified pharmaceutical nanocarriers for intracellular drug and gene delivery. Pept Sci. 2008;90:604–10.

    Article  CAS  Google Scholar 

  273. Deng Y, Song Y, Du Q, Wang CC, Li H, Sui Y, et al. Anti-HPV16 oncoproteins siRNA therapy for cervical cancer using a novel transdermal peptide PKU12. Front Oncol. 2023;13:1175958.

  274. Li C, et al. Nanocomposite multifunctional hydrogel for suppressing osteosarcoma recurrence and enhancing bone regeneration. Chem Eng J. 2022;435:134896.

    Article  CAS  Google Scholar 

  275. Qin Y, et al. A novel long non-coding RNA lnc030 maintains breast cancer stem cell stemness by stabilizing SQLE mRNA and increasing cholesterol synthesis. Adv Sci. 2021;8:2002232.

    Article  CAS  Google Scholar 

  276. Bai G, et al. Stimuli-responsive scaffold for breast cancer treatment combining accurate photothermal therapy and adipose tissue regeneration. Adv Funct Mater. 2019;29:1904401.

    Article  Google Scholar 

  277. Renneker R, Cutler M. Psychological problems of adjustment to cancer of the breast. JAMA. 1952;148:833–8.

    Article  CAS  Google Scholar 

  278. Lovelace DL, et al. Long-term effects of breast cancer surgery, treatment, and survivor care. J Midwifery Womens Health. 2019;64:713–24.

    Article  PubMed  Google Scholar 

  279. Chhaya MP, et al. Sustained regeneration of high-volume adipose tissue for breast reconstruction using computer aided design and biomanufacturing. Biomaterials. 2015;52:551–60.

    Article  CAS  PubMed  Google Scholar 

  280. Sun R, et al. Composite scaffolds of gelatin and Fe3O4 nanoparticles for magnetic hyperthermia-based breast cancer treatment and adipose tissue regeneration. Adv Healthc Mater. 2023;12:2202604.

    Article  CAS  Google Scholar 

  281. Wang X, et al. Bifunctional scaffolds for the photothermal therapy of breast tumor cells and adipose tissue regeneration. J Mater Chem B. 2018;6:7728–36.

    Article  CAS  PubMed  Google Scholar 

  282. Liao J, et al. Review of a new bone tumor therapy strategy based on bifunctional biomaterials. Bone Res. 2021;9:18.

    Article  PubMed  PubMed Central  Google Scholar 

  283. Huang X, et al. Surface engineering of nanoparticles with ligands for targeted delivery to osteosarcoma. Colloid Surf B. 2020;190:110891.

    Article  CAS  Google Scholar 

  284. Wang H, et al. Integrative treatment of anti-tumor/bone repair by combination of MoS2 nanosheets with 3D printed bioactive borosilicate glass scaffolds. Chem Eng J. 2020;396:125081.

    Article  CAS  Google Scholar 

  285. Maleki Dana P, et al. Chitosan applications in studying and managing osteosarcoma. Int J Biol Macromol. 2021;169:321–9.

    Article  CAS  PubMed  Google Scholar 

  286. Ravanbakhsh M, et al. Mesoporous bioactive glasses for the combined application of osteosarcoma treatment and bone regeneration. Mater Sci Eng C. 2019;104:109994.

    Article  CAS  Google Scholar 

  287. Li Z, et al. An update on the roles of circular RNAs in osteosarcoma. Cell Prolif. 2021;54:e12936.

    Article  CAS  PubMed  Google Scholar 

  288. Zhou Y, et al. Single-cell RNA landscape of intratumoral heterogeneity and immunosuppressive microenvironment in advanced osteosarcoma. Nat Commun. 2020;11:6322.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  289. He C, et al. Photosynthetic oxygen-self-generated 3D-printing microbial scaffold enhances osteosarcoma elimination and prompts bone regeneration. Nano Today. 2021;41:101297.

    Article  CAS  Google Scholar 

  290. Geng B, et al. Carbon dot/WS2 heterojunctions for NIR-II enhanced photothermal therapy of osteosarcoma and bone regeneration. Chem Eng J. 2020;383:123102.

    Article  CAS  Google Scholar 

  291. Han J, et al. Controllable assembly/disassembly of polyphenol-DNA nanocomplex for cascade-responsive drug release in cancer cells. Biomaterials. 2021;273:120846.

    Article  CAS  PubMed  Google Scholar 

  292. Han J, et al. Responsive disassembly of nucleic acid nanocomplex in cells for precision medicine. Nano Today. 2021;39:101160.

    Article  CAS  Google Scholar 

  293. Ding J, et al. Stealth and fully-laden drug carriers: self-assembled nanogels encapsulated with epigallocatechin gallate and siRNA for drug-resistant breast cancer therapy. ACS Appl Mater Inter. 2018;10:9938–48.

    Article  CAS  Google Scholar 

  294. Xiong L, et al. Magnetic core–shell silica nanoparticles with large radial mesopores for siRNA delivery. Small. 2016;12:4735–42.

    Article  CAS  PubMed  Google Scholar 

  295. Chen Z, et al. Mg(II)-Catechin nanoparticles delivering siRNA targeting EIF5A2 inhibit bladder cancer cell growth in vitro and in vivo. Biomaterials. 2016;81:125–34.

    Article  CAS  PubMed  Google Scholar 

  296. Chen J, et al. Assembly of bioactive nanoparticles via metal–phenolic complexation. Adv Mater. 2022;34:2108624.

    Article  CAS  Google Scholar 

  297. Kim C-J, et al. Engineering programmable DNA particles and capsules using catechol-functionalized DNA block copolymers. Chem Mater. 2022;34:7468–80.

    Article  CAS  Google Scholar 

  298. Wang Y, et al. Sustained gene delivery from inflammation-responsive anti-inflammatory hydrogels promotes extracellular matrix metabolism balance in degenerative nucleus pulposus. Compos Part B-Eng. 2022;236:109806.

    Article  CAS  Google Scholar 

  299. Eelen G, et al. Basic and therapeutic aspects of angiogenesis updated. Circ Res. 2020;127:310–29.

    Article  CAS  PubMed  Google Scholar 

  300. Majidinia M, et al. The roles of signaling pathways in bone repair and regeneration. J Cell Physiol. 2018, 233:2937–2948.

  301. Pfister BJ, et al. Biomedical engineering strategies for peripheral nerve repair: surgical applications, state of the art, and future challenges. Crit Rev Biomed Eng. 2011;39:81–124.

    Article  PubMed  Google Scholar 

  302. Yudoh K, Sugishita Y, Suzuki-Takahashi Y. Bone development and regeneration 2.0. Int J Mol Sci. 2023;24:8761.

  303. Dhandapani RK, et al. Development of catechin, poly-l-lysine, and double-stranded RNA nanoparticles. ACS Appl Bio Mater. 2021;4:4310–8.

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

This work was supported by National Natural Science Foundation of China (No. 82302401), Shanghai Overseas High-Level Talent Project and High-End Foreign Experts Introduction Plan (No. G2023013020L).

Funding

Open access funding provided by Shanghai University.

Author information

Authors and Affiliations

Authors

Contributions

Conceptualization, W.Z. X.Z. and Y.W.; formal analysis, W.Z. and Y.W.; funding acquisition, Y.W.; methodology, all co-authors; project administration, Y.W.; resources, Y.W.; data curation, all co-authors; supervision, Y.W.; writing-original draft preparation, W.Z. and Y.W.; and writing-review and editing, all co-authors.

Corresponding authors

Correspondence to Xiaojian Zhou or Yongtao Wang.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Competing interests

The authors declare no competing interests.

Additional information

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Zhang, W., Hou, Y., Yin, S. et al. Advanced gene nanocarriers/scaffolds in nonviral-mediated delivery system for tissue regeneration and repair. J Nanobiotechnol 22, 376 (2024). https://doi.org/10.1186/s12951-024-02580-8

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s12951-024-02580-8

Keywords