Abstract
Intelligent nanomedicine is currently one of the most active frontiers in cancer therapy development. Empowered by the recent progresses of nanobiotechnology, a new generation of multifunctional nanotherapeutics and imaging platforms has remarkably improved our capability to cope with the highly heterogeneous and complicated nature of cancer. With rationally designed multifunctionality and programmable assembly of functional subunits, the in vivo behaviors of intelligent nanosystems have become increasingly tunable, making them more efficient in performing sophisticated actions in physiological and pathological microenvironments. In recent years, intelligent nanomaterial-based theranostic platforms have showed great potential in tumor-targeted delivery, biological barrier circumvention, multi-responsive tumor sensing and drug release, as well as convergence with precise medication approaches such as personalized tumor vaccines. On the other hand, the increasing system complexity of anti-cancer nanomedicines also pose significant challenges in characterization, monitoring and clinical use, requesting a more comprehensive and dynamic understanding of nano-bio interactions. This review aims to briefly summarize the recent progresses achieved by intelligent nanomaterials in tumor-targeted drug delivery, tumor immunotherapy and temporospatially specific tumor imaging, as well as important advances of our knowledge on their interaction with biological systems. In the perspective of clinical translation, we have further discussed the major possibilities provided by disease-oriented development of anti-cancer nanomaterials, highlighting the critical importance clinically-oriented system design.
Introduction
The rapid advances of nanobiotechnology have been introducing revolutionary changes in cancer medicine. Compared with traditional formulations, nanomaterial-based therapeutics provide a much wider range of biological functionalities as well as their combinations, allowing them to cope with the complexity of living systems more efficiently [1], [2], [3], [4]. In anti-cancer drug development, nanocarriers have long been used to attain tumor-targeted delivery, controlled drug release, optimized pharmacokinetics, or to deliver multiple drugs at the same time [1], [2], [3, 5], [6], [7]. However, the highly heterogeneous and complicated nature of tumor microenvironment (TME), and the presence of diverse biological barriers against drug delivery, are still major challenges for the development of nanomedicine [2, 8], [9], [10], [11]. Intelligent nanomaterials, namely nanostructures that can be precisely controlled to perform a complex series of actions (e.g., guided navigation, assembly or disassembly, topological change, surface charge reversal) in vivo, either in response to specific biological conditions (e.g., pH, oxidative stress, inflammation, cell phenotype abnormalities), or modulated by external input signals (e.g., radiation, ultrasound, magnetic field), are therefore attracting much research interest and opening new possibilities for nanotechnology-enabled cancer treatment [1, 3, 12], [13], [14], [15].
Early study on anticancer nanomaterials mainly focused on their application as drug carrier to enhance the efficacy of traditional therapeutics. In the recent years, the immense multifunctional capacity of new-generation intelligent nanomaterials has allowed them to be exploited in an increasingly diversified manner. Intelligent nanosystems have demonstrated great potential in emerging fields such as immunotherapy [4, 16], [17], [18] and real-time imaging [3, 19, 20], and provided excellent platforms for personalized medication [21], [22], [23], [24], [25], [26], [27]. Significant effort has also been made to promote our understanding of the in vivo behavior of complex nanosystems, which helped to pave the way for the rational design of intelligent nanotherapeutic as well as their effective clinical translation [8, 28], [29], [30], [31]. This review would like to provide a brief summary on the recent advances in using intelligent nanomaterials for cancer imaging and treatment, and on our knowledge regarding their interaction with the living system (Figure 1).

Schematic overview of the development of intelligent nanomaterials for anti-cancer applications. Anti-cancer nanomaterials should be rationally designed according to the specific needs of different diseases and patients, while their multifunctionality and biomedical properties should be precisely manipulated during construction. To address the current challenges in clinical translation, deep understanding of their interactions with biological systems is also necessary.
Novel strategies for tumor-specific drug delivery
Tissue-targeted drug delivery remains one of the most important medical applications of nanotechnology. Nanomaterial-based vehicles could be designed to specifically recognize diseased tissues or cell types, protect the encapsulated cargo from macrophage or enzyme clearance, actively penetrate biological barriers, and “intelligently” release drug in response to particular microenvironmental signals [8, 9, 32]. While molecular recognition and stimuli-triggered drug release are still the most commonly used strategies for nanomaterial-based tumor targeting, new-generation intelligent nanocarriers have sought to integrate these simple activities rationally and hierarchically, in order to enable a more precise temporospatial control over their actions [33], [34], [35].
Assembly and disassembly of composite nanostructure plays a critical role in the cargo encapsulation and release behavior of intelligent drug carriers. Developing biomaterials with programmable supramolecular behavior would therefore greatly facilitate the quality control of nanotherapeutics as well as reduce non-specific drug release in vivo. Compared to other self-assembly platforms such as peptides, polymers, or liposomes, nucleic acid-based nanomaterials showed particular promise in this aspect, since their self-assembly process could be precisely manipulated via base pairing. It has long been reported that by rationally programming the nucleotide sequences of building blocks, DNA and RNA molecules could be organized into an infinite variety of 2D and 3D structures, with rigorous control over their size, shape and surface functionality [15, 35], [36], [37], [38], [39], [40], [41], [42], [43]. Using such bottom-up approach, therapeutic agents, targeting ligands and stimuli-sensitive switches can be integrated into the nanosystem with striking spatial and stoichiometric accuracy. Previous study has already demonstrated that DNA self-assembled nanostructures were able to safely deliver thrombin, a potent blood coagulator, through intravenous administration to the tumor site without causing non-specific peripheral thrombosis [44]. Notably, the application of DNA origami technique allows thrombin molecules to be precisely arranged at a series of fixed binding sites inside the tubular vehicle, therefore significantly minimizing non-specific exposure in bloodstream (Figure 2A). This advantage of programmable drug loading has been further exploited to deliver multiple agents at the same time. For example, a DNA-origami-delivered tumor nanovaccine was constructed by loading three active ingredients (a tumor antigen peptide plus two different nucleic acid adjuvants) together into a similar tubular “nanorobot” (Figure 2B) [45]. Being equipped with a set of pH-sensitive DNA locking strands, the nanostructure maintained its “closed” tubular form in neutral environment, but can be triggered to “open up” by the acidic pH in endosomes of dendritic cells (DCs) following cellular internalization. The two nucleic acid adjuvants, both toll-like receptor (TLR) agonists, simultaneously activated TLR3 and TLR9 pathways, respectively, thus promoting DC maturation and antigen presentation via a synergic mechanism. Although the potential of polymeric or liposomal nanoparticles in multidrug delivery has already been elaborately described, DNA self-assembled carriers hold unique advantage in that different drugs could be delivered with exact stoichiometric ratio, since the number/position of drug loading sites and conjugation linkers were accurately programmed via sequence encoding. Particularly, small nucleic acid drugs could be efficiently loaded into nucleic acid nanocarriers via hybridization and safely delivered avoiding enzyme degradation.
![Figure 2:
DNA nanotechnology-based intelligent nanomedicines for cancer therapy. (A) Schematic illustration and atomic force microscopy (AFM) characterization of a DNA origami nanorobot for tumor-specific delivery of thrombin. The drug-loaded tubular structure could selectively recognize nucleolin, a tumor blood vessel-specific biomarker. Reproduced with permission [44]. Copyright © 2018 Springer Nature. (B) Schematic illustration and AFM characterization of a pH-sensitive DNA origami nanorobot for tumor vaccination. Reproduced with permission [45]. Copyright © 2020 Springer Nature.](/document/doi/10.1515/mr-2023-0028/asset/graphic/j_mr-2023-0028_fig_002.jpg)
DNA nanotechnology-based intelligent nanomedicines for cancer therapy. (A) Schematic illustration and atomic force microscopy (AFM) characterization of a DNA origami nanorobot for tumor-specific delivery of thrombin. The drug-loaded tubular structure could selectively recognize nucleolin, a tumor blood vessel-specific biomarker. Reproduced with permission [44]. Copyright © 2018 Springer Nature. (B) Schematic illustration and AFM characterization of a pH-sensitive DNA origami nanorobot for tumor vaccination. Reproduced with permission [45]. Copyright © 2020 Springer Nature.
In vivo self-assembly of nanomaterials represents another emerging trend for intelligent drug delivery. While nanoparticle formulations may improve targeted delivery of antitumor therapeutic, their ability to penetrate into tumor tissue is limited compared to free drugs [8, 46]. This could be particularly problematic in solid tumors with highly fibrotic and dense stroma. Small-sized building blocks that can be administered in their free form but selectively aggregate/assemble into nanostructures at the tumor site were thus developed to address this challenge. So far, existing in vivo self-assembly platforms were mainly constructed with functional short peptides, which were hydrophilic in their original form yet became largely hydrophobic in response to tumor-related stimuli, therefore ending up in spontaneous aggregation [19, 47], [48], [49]. Cong et al. reported a polymer–peptide conjugate that underwent a sharp hydrophobicity increase when pH changed from 7.4 to 6.5, allowing the monomer to specifically self-assemble in the weakly acidic TME [46]. Through the incorporation of a cell penetrating peptide motif, the monomer was able to deeply penetrate into the tumor, facilitating the delivery of a conjugated cytotoxic peptide. Since such stimuli-triggered assembly typically occurs in a deeper region of tumor tissue, in situ constructed self-assembly often show remarkably prolonged tumor retention compared to intravenously administered nanoparticles, while off-target monomers could be quickly degraded to minimize side effects [49]. For example, a photosensitizer-conjugated peptide that can be specifically cleaved by the TME-overexpressed enzyme gelatinase has been employed for targeted photothermal therapy, with the enzyme-induced peptide aggregation significantly improving the intratumoral half-life of photosensitizer [50]. An et al. conjugated a caspase 3/7 cleavable peptide with a molecular recognition motif targeting X-linked inhibitor of apoptosis protein (XIAP), characteristically upregulated in many types of tumor cells [51]. The specific binding between XIAP and the peptide monomer induced cell uptake and activated intracellular caspase 3/7 enzyme, leading to the cleavage of the monomer, which subsequently self-assembled into fiber-like superstructures. When the peptide was further linked with doxorubicin, this in situ cascade activatable fibril formation significantly improve the accumulation and detention of the chemotherapeutic drug. Recently, it has also been demonstrated that certain peptide sequences acquire tendency of aggregation only after binding with proteins. An amyloid-like peptide GNNNQNY conjugated with integrin-targeted motif RGD was reported to show “intelligent” self-assembly behavior that was strictly integrin-dependent, leading to highly efficient targeted delivery based on tumor-related integrin overexpression [52].
Besides bottom-up approaches aiming at rigorously manipulate the intelligent behavior of nanosystem from the construction level, the use of naturally derived materials as functional elements has also witnessed notable progress. It has been suggested that naturally formed supramolecular systems, such as protein coronas [29, 53], plasma membranes [54], [55], [56], cell-secreted vesicles [13, 57, 58], often retain unique biomimetic activity that cannot be readily simulated by artificial nanomaterials. For instance, Zeng et al. reported that cisplatin, a platinum-containing first-line chemotherapeutic, spontaneously form Pt nanoparticles with diameter of 6–8 nm in human blood due to serum albumin-induced nucleation [59]. The product nanoparticles were coated with a serum protein corona and could selectively accumulate in tumor tissues via albumin-mediated tumor targeting. Particularly, the intratumoral half-life of Pt nanoparticles resulted from blood-triggered synthesis was remarkably longer compared to chemically synthesized Pt nanoparticles coated with albumin. The authors suggested that while single-composition albumin coating was vulnerable to hydrolyzation in vivo, while the complex corona formed in human blood could effectively protect the albumin molecules from rapid degradation.
Cell-membrane modified nanomaterials have attracted enormous attention in these years. Coating nanoparticles with plasma membrane isolated from living cells may allow us to exploit the complex membrane protein profile for tumor-targeting, biological barrier penetration, or immunotherapy, while avoiding complicated chemical modification [56, 60], [61], [62]. A biomimetic nanoparticle coated with tumor-associated macrophage (TAM) membrane has recently been developed to improve the strongly immunosuppressive TME after photothermal therapy treatment [63]. When injected intravenously, the nanoparticles showed chemotactic behavior similar to living TAMs, and actively targeted the post-treatment tumor tissue rich of inflammatory cytokines. Platelet membrane coated nanomaterials, on the other hand, were expected to show platelet-like wound-binding activity and tumor cell adhesion [64], [65], [66], thus have been exploited for targeted delivery to post-surgical tumor sites [67]. Loaded with an immune checkpoint inhibitor and an anti-angiogenesis drug, this wound-targeted nanosystem effectively suppressed the relapse rate after surgery in a hepatocellular carcinoma mouse model.
Nanovaccines and enhanced immunotherapy
In cancer nanomedicine, immunotherapy aims at stimulating the immune system to “automatically” monitor and deplete malignant cells, and is currently regarded as one of the most potent and promising therapeutic modalities against cancer [68, 69]. While impressive clinical success has already been achieved by various immunotherapeutic approaches, including immune checkpoint inhibition and adoptive T cell therapy, their response rates among patients are still largely unsatisfactory [16, 70]. Much effort has been dedicated to the development of novel strategies to enhance anti-tumor immunity, in which nanomaterial-based platforms are playing an increasingly crucial role [16], [17], [18, 71].
Similar to vaccines against infectious diseases, tumor vaccines present tumor-related biological information (e.g., antigens) to the immune system and elicit antigen-specific immune recognition and response [72]. Tumor vaccination is able to provoke long-term systematic immune memory against metastases and recurrence, and allows highly individualized treatment [21], [22], [23, 73]. Traditionally, peptides containing tumor antigen information were delivered to antigen presenting cells (APCs), especially DCs, whose maturation subsequently activated antigen-specific CD8+ T cell immunity [72]. However, the immunogenicity of such antigen peptides is typically weak, requiring co-delivery of adjuvants. Nanovaccines that are capable of delivering multiple components simultaneously into the same APC are therefore of great interest, and multifunctional strategies to actively target APCs, promote antigen cross-presentation, or facilitate customized treatment have been accordingly developed [23, 72, 74, 75].
Qin et al. proposed a two-step strategy for the targeted delivery of nanovaccines to lymph nodes (LNs) [76]. Compared to peripheral DCs, which are the conventional targets of subcutaneously or intramuscularly injected vaccines, LN-resident DCs can be more attractive targets for tumor antigen delivery, due to their dense local population and close proximity to T cells. A major obstacle of LN-targeted delivery, however, is the lack of convenient molecular target for ligand-receptor recognition. In this work, a PEGylated phospholipid precursor containing a terminal azide group was first injected subcutaneously. Through a previously known mechanism called albumin hitchhiking [77], the precursor efficiently accumulated in LNs after binding to intrinsic albumin. The lipid moiety could then readily insert into the plasma membrane of lymphatic endothelial cells, “labeling” LNs with the azide group. Tumor antigen and adjuvants were then delivered by a liposome-coated nanoparticle modified with dibenzocyclooctyne (DBCO) groups, which selectively react with azide in physiological conditions via bioorthogonal chemistry [78]. The authors stated that targeting LN-resident cells helped to circumvent the limited efficiency of DC migration from peripheral tissues, and significantly enhanced the antigen presentation process.
Yin et al. presented another strategy to improve the APC-mediated immune response of subcutaneous vaccines [79]. An intelligent transformable hydrogel was constructed with graphene oxide and low molecular weight polyethylenimine via electrostatic interaction. Once injected into living body, this hydrogel tended to slowly transform into small nanoparticles on the nano-bio interface. Encapsulated with an mRNA antigen and an adjuvant, the hydrogel could hence gradually produce antigen- and adjuvant-loaded nanoparticles in situ for over 30 days after subcutaneous administration, providing a much more persistent immunostimulatory effect compared to conventional nanovaccines. By continuously presenting the same antigen to the immune system, this transformable platform also effectively enhanced antigen-specific T cell response.
One of the predominant limitations of traditional tumor vaccines lies in the fact that tumor cells are highly heterogeneous among patients and it is impossible for a prescribed vaccine formulation to effectively treat all the cases [21, 22, 73, 80]. Developing nanoplatforms that can be easily customized case-by-case to deliver patient-specific information may thus help to solve this problem. A “plug-and-display” platform based on bacteria outer membrane vesicles (OMVs) has been developed to deliver customizable combination of multiple tumor antigens [81]. As bacteria-derived nanosized membrane vesicles, OMVs contain various pathogen-associated molecular patterns (PAMPs) and have intrinsic immunostimulatory effects [82], [83], [84]. Previously, OMVs have been used as anti-tumor immunotherapeutic [85] and to deliver checkpoint antibodies [86]. By genetically manipulate the source bacteria, two different protein catchers were expressed on the OMV surface, which could respectively bind specific protein tags through the peptide superglue technology, allowing different antigens linked to these tags to be rapidly and flexibly displayed (Figure 3) [81]. The OMV platform served as both vehicle and adjuvant. Such versatile platforms represent an emerging trend in tumor vaccine development and may facilitate the clinical translation of nanovaccines against cancer.
![Figure 3:
Schematic illustration of a customizable tumor antigen delivery platform enabled by biologically engineered bacteria outer membrane vesicles. Reproduced under the terms of the CC-BY Creative Commons Attribution 4.0 International License (http://creativecommons.org/licenses/by/4.0/) [81].](/document/doi/10.1515/mr-2023-0028/asset/graphic/j_mr-2023-0028_fig_003.jpg)
Schematic illustration of a customizable tumor antigen delivery platform enabled by biologically engineered bacteria outer membrane vesicles. Reproduced under the terms of the CC-BY Creative Commons Attribution 4.0 International License (http://creativecommons.org/licenses/by/4.0/) [81].
In addition to synthetic validated tumor antigens (e.g., peptides and mRNAs), the use of tumor cell-derived components (e.g., whole cells, cell lysates, cell-derived vesicles) as antigenic material has also been proposed [73, 87], [88], [89]. Tumor cell membranes, for example, contain a great variety of antigenic motifs, especially neoantigens and provide an excellent platform for personalized antigen delivery. Diverse strategies have demonstrated that co-delivery of tumor cell membrane fragments and adjuvants could elicit strong tumor-specific immunity in animal models [90], [91], [92], [93], [94]. Recently, Chen et al. reported a hybrid membrane nanovaccine formulation, constructed by fusing autologous tumor cell membrane with bacterial cytoplasmic membrane into a hybrid coating layer, which was further decorated on the surface of polymeric nanoparticles (Figure 4A) [25]. Similar to OMVs, bacterial cytoplasmic membrane could be used as naturally-derived adjuvant, with favorable biosafety due to the separation from toxic cell wall components such as lipopolysaccharide (LPS). In the hybrid product, the immunostimulatory properties of the bacteria-derived components enhanced APC uptake, and subsequently the processing and presentation of antigens on the tumor cell membrane surface. Such strategy was able to elicit anti-tumor immunity and immune memory without the need of antigen isolation and validation, while the use of autologous tumor cell membranes provided a possibility to dynamically represent the biological information of heterogeneous tumors on individual level. In addition, Zhang et al. used nanoparticles co-assembled by tumor cell membrane and E. coli-derived membrane to stimulate bone-marrow-derived cells (BMDCs). The pulsed BMDCs expressed tumor-specific molecules class I (MHC-I) antigen complexes, costimulatory molecules and lymphocyte homing molecules on their membrane, which could be extracted and decorated on dendritic nanostructures resulting in a dendritic cell-mimicking nanovaccine system (Figure 4B) [95].
![Figure 4:
Engineering autologous tumor cell membranes for enhanced and personalized tumor vaccination. (A) Schematic illustration of a hybrid membrane nanovaccine using bacterial cytoplasmic membranes as biomimetic adjuvant. Reproduced with permission [25]. Copyright © 2021 The American Association for the Advancement of Science. (B) Schematic illustration of a DC membrane nanovaccine presenting autologous tumor antigens. Reproduced with permission [95]. Copyright © 2022 Wiley-VCH GmbH. NPs, nanoparticles; nanoEBs, E. coli-B16-OVA hybrid nanostructures; nanoECs, E. coli-CT26 hybrid nanostructures; TAAs, tumor-associated antigens; BMDCs, bone-marrow-derived cells; DCs, dendritic cells; FFNPs, folic acid and ferrous ion self-assembled nanostructures.](/document/doi/10.1515/mr-2023-0028/asset/graphic/j_mr-2023-0028_fig_004.jpg)
Engineering autologous tumor cell membranes for enhanced and personalized tumor vaccination. (A) Schematic illustration of a hybrid membrane nanovaccine using bacterial cytoplasmic membranes as biomimetic adjuvant. Reproduced with permission [25]. Copyright © 2021 The American Association for the Advancement of Science. (B) Schematic illustration of a DC membrane nanovaccine presenting autologous tumor antigens. Reproduced with permission [95]. Copyright © 2022 Wiley-VCH GmbH. NPs, nanoparticles; nanoEBs, E. coli-B16-OVA hybrid nanostructures; nanoECs, E. coli-CT26 hybrid nanostructures; TAAs, tumor-associated antigens; BMDCs, bone-marrow-derived cells; DCs, dendritic cells; FFNPs, folic acid and ferrous ion self-assembled nanostructures.
Compared with other administration routes, oral administration of therapeutics has many advantages in terms of safety, patient compliance and cost [96, 97]. As the largest immune organ in human body, intestine contains over 70 % of the body’s immune cells, making orally administered tumor vaccines particularly attractive [98]. In order to circumvent the challenges posed by the complex gastrointestinal environment to vaccine delivery, Yue et al. developed a genetically engineered E. coli bacteria, which could be readily retained in the intestine after oral administration [99]. The bacteria started to produce tumor antigen-containing OMVs in situ only when induced by another orally taken substance, arabinose. These OMVs could effectively penetrate the intestinal epithelial barrier [100] to encounter the DCs in the lamina propria, leading to enhanced antigen presentation and DC maturation (Figure 5).
![Figure 5:
Schematic illustration of an orally administered tumor vaccine in situ produced by engineered intestinal bacteria. Reproduced under the terms of the CC-BY Creative Commons Attribution 4.0 International License (http://creativecommons.org/licenses/by/4.0/) [63]. DCs, dendritic cells; TCR, T cell receptor.](/document/doi/10.1515/mr-2023-0028/asset/graphic/j_mr-2023-0028_fig_005.jpg)
Schematic illustration of an orally administered tumor vaccine in situ produced by engineered intestinal bacteria. Reproduced under the terms of the CC-BY Creative Commons Attribution 4.0 International License (http://creativecommons.org/licenses/by/4.0/) [63]. DCs, dendritic cells; TCR, T cell receptor.
While current nanovaccines have been demonstrated as encouragingly powerful in eliciting anti-tumor immune response [72, 74], there are still many other factors that affect their effectiveness in clinic. In many solid tumors, a highly immunosuppressive microenvironment prevents the infiltration and activity of cytotoxic immune cells, resulting in reduced response rate and resistance [21, 70, 101]. A great number of intelligent nanomaterial-based strategies have been developed to remodel the TME and to sensitize such “cold” tumors to anti-tumor immunity [70, 102], most commonly via enhanced immune checkpoint inhibition [86, 103], [104], [105] and macrophage repolarization [106], [107], [108], [109]. For example, Lin et al. reported that the immunosuppressive TME could be effectively relieved by restoring the local expression of phosphatase and tensin homolog deleted on chromosome 10 (PTEN), a crucial tumor suppressor gene mutated frequently in a wide range of tumors. A nanocarrier encapsulating PTEN-encoding mRNA significantly improved the therapeutic efficacy of immune checkpoint inhibition treatment [110]. On the other hand, Feng et al. constructed an OMV-based intelligent nanosystem to re-educate tumor-associated macrophages (TAMs), the most abundant immune suppressive cells in TME [106]. A CD47-neutralizing nanobody was expressed on the OMV surface, which could block the CD47-dependent anti-phagocytosis signal on tumor cells, making them vulnerable to macrophage-mediated immunity. The whole system thus acted as a two-way adapter that simultaneously bound tumor cells via the nanobody and TLRs on TAMs via the immunogenic OMV vehicle, repolarizing TAMs and sensitizing tumor cells at the same time. The nanoplatform was further coated with a polymeric later containing diselenide bonds, which shielded its immunogenicity during intravenous delivery, but could selectively release the active OMV-nanobody composite upon localized radiation treatment.
Nanomaterial platforms for temporospatially specific tumor imaging
Techniques for high-resolution and real-time tissue imaging are of tremendous importance in cancer diagnosis and treatment, and have always been among the most active fields in nanobiotechnology [20]. Advanced imaging techniques allow clinicians to diagnose and stratify tumors in their early stage, perform image-guided surgery, monitor disease progression or patient response to treatment. Adequate bioanalytical methods are also critical in studying the molecular mechanisms involved in tumor development or the interactions between tumor cells and anti-cancer nanotherapeutics. In recent years, many intelligent strategies have been developed to improve the sensitivity, selectivity, and time or space resolution of nanomaterial-based imaging platforms, with notable progress being made in this aspect [3, 111].
In early-generation nanomaterial-based imaging strategies, small molecule dyes or nanoparticle probes were encapsulated in nanocarrier and delivered to the tumor site via molecular targeting or stimuli-triggered activation. Nonetheless, conventional nanocarriers underwent degradation in the tumor tissue in a relatively short time, and the small-sized imaging agents encapsulated were also rapidly cleared from the tumor. Novel delivery strategies based on assembly/aggregation-induced retention (AIR) effect of fluorescent probes have been reported to significantly amplify imaging signals by allowing the imaging agent to be stably retained in the target tissue for a longer time [112, 113]. Zhao et al. designed a functional peptide-conjugated near-infrared (NIR) dye that was responsive to fibroblast activation protein-α (FAP-α), therefore forming fiber-like β-sheet nanostructures selectively in the presence of cancer-associated fibroblasts, a major cellular component of TME (Figure 6A) [114]. Such TME-responsive aggregation behavior led to a 5.5-fold signal amplification 48 h post-injection. To improve the tumor specificity, this platform has been further modified to address the challenge of small-sized tumor monitoring. A tumor-targeting motif and a long-circulation motif were incorporated into the peptide-NIR dye conjugate, enabling it to actively target small orthotropic pancreatic tumors. The AIR effect further extended the imaging window to as long as 96 h post-injection. Since the fibroblast-triggered nanofiber formation selectively occurred at the boundary of pancreatic tumors, this AIR-based strategy may greatly facilitate boundary imaging and identification of small tumors during surgery. Zhou et al., on the other hand, constructed an iron oxide nanoparticle-based in situ self-assembly platform for the imaging of hypoxic tumors [115]. Ultrasmall iron oxide nanoparticles were modified with a thiol-containing crosslinking ligand and a nitroimidazole-based hypoxia-sensitive ligand, which remained inert in normal tissues but could irreversibly crosslink with thiol groups under hypoxia conditions in the presence of NADPH and reductase. When intravenously injected in vivo, this intelligent nanoprobe selectively aggregated in response to hypoxia regions of tumor, thus significantly amplifying the local T2-weighted magnetic resonance signal in a hypoxia-dependent manner (Figure 6B). Since hypoxia condition played an important role in solid tumor’s resistance to chemotherapy and radiotherapy, the author suggested that such stimuli-responsive assembly strategy might be potentially exploited for hypoxia-targeted theranostics.
![Figure 6:
In situ self-assembled nanosystems for intelligent tumor imaging. (A) Schematic illustration of tumor associated fibroblast-responsive self-assembly of rationally designed peptide probes. Reproduced with permission [114]. Copyright © 2019 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim. (B) Schematic illustration of hypoxia-sensitive in situ assembly of ultrasmall iron oxide nanoparticles for enhanced magnetic resonance imaging. Reproduced with permission [115]. CFAs, cancer associated fibroblasts; FAP-α, fibroblast activation protein-α.](/document/doi/10.1515/mr-2023-0028/asset/graphic/j_mr-2023-0028_fig_006.jpg)
In situ self-assembled nanosystems for intelligent tumor imaging. (A) Schematic illustration of tumor associated fibroblast-responsive self-assembly of rationally designed peptide probes. Reproduced with permission [114]. Copyright © 2019 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim. (B) Schematic illustration of hypoxia-sensitive in situ assembly of ultrasmall iron oxide nanoparticles for enhanced magnetic resonance imaging. Reproduced with permission [115]. CFAs, cancer associated fibroblasts; FAP-α, fibroblast activation protein-α.
Among existing stimuli-triggered imaging nanoplatforms, DNA nanodevices have showed noteworthy promise since their high programmability and sequence-specific recognition potential enabled extraordinary flexibility and precision in terms of stimuli-responsive design [111, 116, 117]. Particularly, a variety of DNA-based fluorescence nanoprobes have been constructed to achieve precise temporal and spatial control of signal activation, especially on the cellular and subcellular level. Shao et al. reported a DNA nanosensor that could selectively visualize enzyme activity in a specific subcellular organelle [118]. The nanosystem was consisted of an ultraviolet (UV) light-activatable DNA nanoprobe, an upconversion nanoparticle (NIR-to-UV transducer) and a mitochondria- or nucleus-targeting ligand. The conjugated targeting ligand helped to deliver the nanoprobe into interested organelles, and under light radiation, the otherwise inert DNA nanoprobe transformed to a molecular beacon structure that could be specifically cleaved by apurinic/apyrimidinic endonuclease 1 (APE1), an enzyme with important roles in tumor occurrence and metastasis. The coupling with upconversion nanoparticles allowed the nanoprobe to respond to NIR light that had superior tissue penetration capability. Another organelle-targeted nanosystem was similarly constructed for the detection of mitochondrial microRNAs (miRNAs) [119]. After mitochondria-targeted localization mediated by triphenylphosphonium (TPP), the nanosensor could be selectively “switched on” by NIR light signal, and react with specific local miRNA molecules via toehold-mediated strand displacement reactions to activate strong fluorescence (Figure 7A). Notably, by carefully programming the original DNA sequence, this system can be set to only respond the simultaneous presence of two different miRNAs, enabling logic-gated imaging. Xiang et al. also reported a multivariate-gated nanoprobe that responded to dual targets, but relying on an interesting peptide nucleic acid (PNA)-enabled system [120]. It has been demonstrated that by incorporating PNA/peptide copolymers into DNA nanostructures, the biological activity of functional DNA sequences could be manipulated via peptide-targeted approaches such as enzyme cleavage, significantly expanding the potential of stimuli-responsiveness. This nanosystem was constructed with a matrix metalloproteinase (MMP)-responsive PNA/peptide/PNA copolymer, the cleavage of which by MMP2/9 in TME consequently activated an ATP-sensitive DNA fluorescence nanoprobe. The whole system was further coupled with a tumor-targeting aptamer, leading to spatially specific correlated imaging of MMP2/9 and ATP, both characteristic biomarkers of tumor invasiveness (Figure 7B). By providing temporospatially specific methods to monitor the production, activity, subcellular localization of key proteins/signaling molecules involved in tumor progression, such new-generation intelligent nanoprobes might eventually promote our understanding of the heterogeneous and dynamic pathology of cancer.
![Figure 7:
Intelligent DNA nanobiosensors for multi-target tumor imaging. (A) Schematic illustration of temporospatially specific detection of multiple mitochondrial microRNAs via DNA technology. Reproduced with permission [119]. Copyright © 2021 Wiley-VCH GmbH. (B) Schematic illustration of correlated imaging of two TME biomarkers by a peptide nucleic acid nanosystem. Reproduced with permission [120]. Copyright © 2021 Wiley-VCH GmbH.](/document/doi/10.1515/mr-2023-0028/asset/graphic/j_mr-2023-0028_fig_007.jpg)
Intelligent DNA nanobiosensors for multi-target tumor imaging. (A) Schematic illustration of temporospatially specific detection of multiple mitochondrial microRNAs via DNA technology. Reproduced with permission [119]. Copyright © 2021 Wiley-VCH GmbH. (B) Schematic illustration of correlated imaging of two TME biomarkers by a peptide nucleic acid nanosystem. Reproduced with permission [120]. Copyright © 2021 Wiley-VCH GmbH.
Understanding nano-bio interactions
Despite of the continuous emergence of novel intelligent nanomaterials that were supposed to hold great promise in medical applications, the clinical translation of nanomaterial-based treatments remained notoriously troublesome [2, 69, 121, 122]. Among the primary obstacles hindering the translation of anti-cancer nanomaterials from benchside to bedside, it has long been noted that the additional complexity compared with traditional drug forms has made it much more difficult to accurately characterize and predict their in vivo behaviors [2, 8, 123, 124]. With the increased research interest dedicated to highly sophisticated, robot-like nanosystems that are expected to perform multiple in vivo actions, the complexity in their interactions with biological systems is becoming particularly problematic. Integrated methods and tools to systematically evaluate nano-bio interactions are urgently needed, and deep understanding of the in vivo fate of intelligent nanomaterials will be essential for the development of next-generation nanomedicines. Below we would like to briefly summarize several representative advances in this aspect.
Formation of protein corona at the nano-bio interface inevitably occurs to all the nanosystems working in biological environment, and is seen as an essential event for nanomaterials to acquire biological identity [29, 53]. Real-time, in situ methods to characterize the dynamic evolution of adsorbed proteins would be highly necessary for accurate prediction of nanomaterial behavior in complex application scenarios and have attracted much research interest [30, 125], [126], [127], [128]. In particular, the so-called “soft” corona, consisting of a superficial layer of relatively loosely attached proteins, is posing specific challenges for bioanalytical study, both due to their critical roles in nano-bio interaction as the outmost layer of nanomaterial-protein complexes, and their highly dynamic and unstable identity [126], [127], [128], [129]. Sanchez-Guzman et al. demonstrated that using cryo-transmission electron microscopy (cryoTEM) and in situ techniques such as synchrotron-radiation circular dichroism (SRCD), it was possible to trace the conformation changes of soft corona components caused by nanosurface-driven unfolding [127]. The authors suggested that such changes could not be detected using traditional methods that require nanomaterials being separated from the biological medium, since soft coronas were largely lost during centrifugation (Figure 8). In order to address this separation challenge, Baimanov et al. have recently reported a bio-layer interferometry (BLI)-based strategy for centrifugation-free analysis of soft corona proteins [129]. Immobilizing Cu2S nanoparticles on BLI biosensors enabled time-resolved in situ monitoring of corona formation, and the soft corona components could be selectively eluted for tandem mass spectrometry profiling (Figure 9).
![Figure 8:
Detection of soft coronas on nanoparticle surface. Classical isolation approaches typically resulted in the loss of soft coronas and conformation changes of corona proteins compared to their in situ states (upper panel). It is therefore important to develop analytical methods to capture the subtle conformation alterations of weakly bound proteins induced by the nano-bio interface (bottom panel). Reproduced with permission [127]. Copyright © 2020 American Chemical Society.](/document/doi/10.1515/mr-2023-0028/asset/graphic/j_mr-2023-0028_fig_008.jpg)
Detection of soft coronas on nanoparticle surface. Classical isolation approaches typically resulted in the loss of soft coronas and conformation changes of corona proteins compared to their in situ states (upper panel). It is therefore important to develop analytical methods to capture the subtle conformation alterations of weakly bound proteins induced by the nano-bio interface (bottom panel). Reproduced with permission [127]. Copyright © 2020 American Chemical Society.
![Figure 9:
Schematic illustration of bio-layer interferometry (BLI)-based separation of hard and soft coronas. (A) Using BLI biosensors to monitor the dynamic protein adsorption on the surface of immobilized nanoparticles. (B) Centrifuge-free separation of hard and soft coronas through washing buffer (WB). Reproduced under the terms of the CC-BY Creative Commons Attribution 4.0 International License (http://creativecommons.org/licenses/by/4.0/) [129]. NPs, nanoparticles; SC, soft corona; HC, hard corona.](/document/doi/10.1515/mr-2023-0028/asset/graphic/j_mr-2023-0028_fig_009.jpg)
Schematic illustration of bio-layer interferometry (BLI)-based separation of hard and soft coronas. (A) Using BLI biosensors to monitor the dynamic protein adsorption on the surface of immobilized nanoparticles. (B) Centrifuge-free separation of hard and soft coronas through washing buffer (WB). Reproduced under the terms of the CC-BY Creative Commons Attribution 4.0 International License (http://creativecommons.org/licenses/by/4.0/) [129]. NPs, nanoparticles; SC, soft corona; HC, hard corona.
Synchrotron radiation-based analytical techniques have undergone enormous advancement in recent years and have provided critical tools for study of nano-bio interactions on the cellular and tissue levels. Synchrotron light sources are able to produce radiation with wavelength tunable from infrared to X-ray region with extremely high brilliance, stability and level of polarization, compared to conventional sources. Coupling synchrotron radiation sources with X-ray based analytical techniques enabled non-invasive and in situ mapping of nanomaterials in living cells or tissues with ultrahigh spatial resolution and excellent penetration depth, and could be used to trace the dynamic change of nanomaterials during nano-bio interactions according to properties such as elemental composition, oxidation state, chemical coordination, crystalline phase, etc., as reviewed elsewhere [130], [131], [132]. Synchrotron X-ray based methods are particularly powerful for detection of interactions between nanomaterials and living cells due to their subcellular resolution and living cell imaging ability, and have been applied to characterize the intracellular fate and cell-level biological effects of a wide range of nanomaterials, including inorganic nanostructures, lipid-based nanoparticles [133] and metal-organic frameworks [134]. For example, the cell uptake process and subcellular localization of nanovaccine containing manganese nanoadjuvant during its interaction with DCs has been characterized using synchrotron radiation hard X-ray nano-computed tomography (nano-CT) [135]. In an integrated study [136], diverse synchrotron radiation-based tools were combined to trace the in vivo course of MoS2 nanomaterials on both cellular and tissue level. Soft X-ray nano-CT revealed the subcellular distribution of MoS2 nanodots in various blood cells, with synchrotron radiation microbeam X-ray fluorescence (SRXRF) being applied to map the precise distribution of molybdenum and other elements in liver and spleen. X-ray absorption near-edge spectroscopy (XANES) was further employed to study the metabolism of nanomaterial in liver according to the chemical form changes of molybdenum. It was demonstrated that via the mediation of surface adsorbed serum proteins, principally apolipoprotein E (ApoE), intravenously injected MoS2 nanomaterials significantly accumulated in liver sinusoid and splenic red pulp. Interestingly, biotransformation data indicated that hepatic metabolism of MoS2 transformed the nanomaterial into bioavailable molybdenum, which could be incorporated into molybdenum-dependent metabolic enzymes in liver, up-regulating their activity. The authors suggested that this study indicated the possibility of nanomaterials being actively reused by organisms and in turn having a long-term impact on liver function, implying a more complicated picture regarding the in vivo fate of nanomaterials compared to the traditional biodistribution-metabolism-excretion paradigm (Figure 10).
![Figure 10:
Precise understanding of nano-bio interactions enabled by advanced analytical techniques. Integrated application of synchrotron radiation-based methods revealed the complicated in vivo fate of MoS2 nanomaterials on cell, tissue and whole body levels. Reproduced with permission [136]. Copyright © 2021 Springer Nature.](/document/doi/10.1515/mr-2023-0028/asset/graphic/j_mr-2023-0028_fig_010.jpg)
Precise understanding of nano-bio interactions enabled by advanced analytical techniques. Integrated application of synchrotron radiation-based methods revealed the complicated in vivo fate of MoS2 nanomaterials on cell, tissue and whole body levels. Reproduced with permission [136]. Copyright © 2021 Springer Nature.
Multi-omics research is another analytical approach that have been demonstrated as particularly useful in nano-bio interface characterization, especially in systematically exploring the underlying molecular mechanisms of nanomaterial-associated biological effects [137, 138]. For example, through integrative application of proteomic, metabonomic and lipidomic methods, Cai et al. analyzed the impact of composition changes in gold nanoparticles’ protein corona during nanoparticle-cell interaction on cellular metabolism [139]. Proteomic profiling suggested that the protein composition of corona dramatically evolved across the different stages of nanoparticle-cell interactions (in the bloodstream, in the lysosomal compartment, and in the cytoplasm after lysosome escape), with a notable abundance of intracellular proteins, particularly pyruvate kinase M2 (PKM2) and chaperones, subsequently displacing the initial serum proteins after cellular uptake. Metabonomic and lipidomic data further indicated that this dynamic change in nanomaterial surface protein coating had significant influences on the intracellular metabolism, likely by interfering with corona protein-dependent processes such as glycolysis (PKM2-dependent) and chaperone-mediated autophagy.
Understanding the complex events at nano-bio interfaces as well as their principles also facilitate the development of novel anti-cancer treatments. For example, it has been reported that the composition of adsorbed serum proteins could be manipulated through nanomaterial surface properties and be exploited for tissue-targeted delivery [29, 53, 140]. Tuning the surface charge of liposomes could lead to differences in their reticuloendothelial system clearance profile and selectively increase their liver or spleen accumulation to facilitate organ-specific drug delivery [141]. During the interaction with single cells, it was demonstrated that the morphology of nanomaterials significantly affects their tendency to enter the cell or adhere to cell surface, allowing researchers to design nanocarriers with specific size/shape to maximize their extravasation or cell uptake [8, 124]. Liang et al. reported that surface rigidity of nanoparticles also influenced their tumor-targeting efficiency, presumably by modulating their contact with plasma membrane of endothelial cells and tumor cells [142].
It has long been demonstrated that certain nanomaterials have specific binding capability with functional proteins and could regulate their biological activity through specific molecular interactions. Proteomic analysis indicated that metallofullerenol nanoparticle Gd@C82(OH)22 had a strong binding affinity with complement 1q (C1q), a protein component of complement system [143]. Such binding spontaneously altered the protein’s secondary structure, making metallofullerenol-binding C1q capable of eliciting innate immune response, promoting the secretion of pro-inflammatory cytokines and activating macrophages. C1q-coated Gd@C82(OH)22 nanoparticles could therefore be potentially exploited for immunomodulation. In another study, it was reported that graphdiyne oxide (GDYO) nanosheets specifically interacted with signal transducer and activator of transcription 3 (STAT3), a key signaling molecule in the polarization of macrophages [144]. Molecular simulation suggested that the GDYO-STAT3 affinity was driven by multiple factors, including structure matching, hydrogen bonding and salt bridges. Importantly, when GDYO was administered to tumor-bearing mice and internalized by macrophages in the TME, this strong interaction between GDYO and STAT3 eventually resulted in the inhibition of immunosuppressive macrophage phenotype (M2) and the re-polarization of TAMs towards the anti-tumor phenotype (M1). These results implied that besides biological effects related to general nano properties such as shape, size, softness, surface charge, nanomaterials could also regulate cellular activity via specific molecular mechanisms. Exploring such mechanisms, especially those closely related to tumor-associated pathways and proteins, may thus provide invaluable information for drug discovery in the future.
So far, complicated biological effects remained one of the major obstacles in nanomedicine development and it is expected that further technical and methodological advances would be required to confront this challenge. Regarding characterization methods, current super-resolution microscopic techniques such as structured illumination microscopy (SIM) or photo-activated localization microscopy (PALM) allowed fluorescence imaging of cellular and subcellular structures with spatial resolution of less than 100 nm [145], and have already been combined with electron microscopic methods to study nanoscale targets (including fluorescent nanostructures) in an ultrastructural context [146]. Further integration with techniques capable of 3D scanning and reconstruction of cells or organoids such as lightsheet microscopy [147], serial block-face electron microscopy (SBEM) [148] and micro-computed tomography (micro-CT)/nano-CT [136, 148], and those for dynamic imaging of living tissues such as multiphoton microscopy [149], would allow us to correlate biological information over a wide range of scales (from molecular level to whole animal level), and in both spatially resolved and time-resolved manner. One of such multi-scale imaging platforms has already been successfully constructed to map the comprehensive mitochondrial and metabolic network in lung cancer mouse models [148], and these techniques could be extremely helpful in characterizing interactions between nanomaterials and living body via a holistic approach.
While proteome analysis has always played a critical role in nano-bio interface study, its potential has been limited by the fact that protein samples must be extracted from isolated and lysed cells. Recent advances in labeling and profiling techniques, for example proximity labeling, have enabled proteomic profiling of living cells, resulting in the possibility to perform proteome analysis in a highly spatially specific fashion [150, 151]. Xie et al., for instance, conducted a proximity labeling-based, centrosome-specific proteome study to identify the key centrosomal proteins in cancer cell mitosis, implying the potential application in mapping nanomaterial-induced changes on the organelle level. Moreover, Zhang et al. exploited this method for in situ labeling of nanoparticle protein coronas, which has enabled dynamic proteomic study of the rapidly formed corona at different stages of nanoparticle-cell interactions [152].
To further deepen our knowledge of nano-bio interactions, efficient experimental and theoretical models to predict nanomaterials’ in vivo behavior would be necessary. Animal-free tissue models, including microfluidic organ-on-chips and organoids, have showed great potential for high-throughput analysis of nano-bio interface in complex tissue microenvironment [153, 154]. In current nanomedicine, the application of these models was mainly limited to toxicity assessment, and how to construct animal-free organ models to study different types of nano-bio interactions remained largely unclear. Nonetheless, it has already been reported that vascular-on-chips could be used to mimic normal and tumor blood vessels, and to evaluate the extravasation ability of nanoparticles under different conditions [155, 156]. Similarly, blood-brain-barrier-on-a-chip models have been used to study the principles of nanoparticle transportation into brain [157]. Recently, an emerging type of animal-free tumor model, patient-derived tumor-like cell clusters (PTCs) have drawn much interest. Constructed with patient-derived tumor samples retaining much of the microenvironmental components, PTCs showed attractive performance in prediction of individual chemosensitivity. However, whether PTCs are also suitable for studying nanomaterial-tumor interactions, including targeting, stroma penetration, and stimuli-triggered drug release is still to be verified.
Statistical and computational tools have long been employed to predict bioactivity and biological effects of nanomaterials. For several types of nanomaterials, especially metal and metal oxide nanostructures, structure-activity relationship models (nano-SARs) have already been extensively constructed to predict toxicological properties [158]. Computational techniques like molecular dynamics have been proved particularly useful in simulating interactions involving biological components, for example biomaterial self-assembly, protein binding, cellular uptake, and penetration across endothelial barriers [159, 160]. Predictive models for highly complicated nanostructures, however, are still much needed, especially for those containing “soft” building blocks such as biopolymers, lipids, peptides and nucleic acids. In addition, while a considerable number of models have been developed to predict biomedical properties of nanomaterials, such as size, surface charge, drug encapsulation efficiency, or cellular uptake [160, 161], it is often unclear how should these properties be correlated to the in vivo delivery and therapeutic efficiency of the nanosystem. In recent years, many have suggested that the rapid development of machine learning algorithms could remarkably improve our capability of mass data interpretation and exploitation, and eventually lead to the accurate understanding of complicated nano-bio interfaces [160, 162]. Sengottiyan et al., for instance, demonstrated that through the application of genetic algorithm-based machine learning techniques, it has become possible to include protein corona fingerprint information into the descriptors of structure-property relationship models, which could be an attractive progress in bridging the gap between properties of nanomaterials and their eventual biological effects [163].
Clinically-oriented development of intelligent nanomedicine
Since the approval of the first nanomaterial-based anti-cancer drug, liposomal doxorubicin (Doxil) in United States, 1995, many nanomaterials have been approved for clinical usage or entered clinical trials [6, 7, 121, 164]. The majority of these early-generation nanotherapeutics, however, were based on relatively simple re-formulation of existing anti-cancer drugs and were therefore difficult to achieve significantly improved efficacy. In the recent years, it has been recognized that to facilitate the translation of nanotherapeutics, intelligent nanosystems should be rationally designed in a clinically-oriented manner [69, 165]. This implied developing therapies not according to currently available drugs, but to the specific needs presented by individual cases of disease and patient, in order to optimize the therapeutic response and minimize non-specific effects.
With the increased understanding of cancer pathology and the continuous advancement of diagnostic and prognostic methods, the convergence with precise patient stratification is regarded as one of the golden opportunities for nanomedicine [3, 69]. In clinical trials and bedside application, patients that are more likely to respond to a certain nanotherapy could be pre-selected, so that the therapeutic potential of treatment could be fully demonstrated and exploited. Moreover, due to the great heterogeneity of cancer, it is also necessary to develop specific nanotherapies for each distinct subgroup of patients. For example, in patients with pancreatic cancer, approximately 5 % carry mutation in genes encoding the BRCA proteins, which regularly participate in cellular DNA repair via the homologous recombination pathway [166]. Pancreatic cancer with BRCA mutations is therefore vulnerable to poly-ADP-ribose polymerase (PARP) inhibitors, since they could cause DNA double-strand breaks and lead to synthetic lethality in the absence of BRCA-mediated repair mechanism. Du et al. showed that a targeted nanoparticle that co-deliver PARP inhibitor olaparib, and a first-line chemotherapeutic for pancreatic cancer, gemcitabine, could significantly improve the therapeutic efficacy of both drugs in BRCA-mutated pancreatic tumors and effectively suppressed tumor growth [166]. Similarly, Ebeid et al. reported that in cancer cells with loss-of-function p53 mutation, the combination of paclitaxel and tyrosine kinase inhibitor delivered by nanoparticles specifically induce synthetic lethality due to the abrogation of G2/M checkpoint [167]. By carefully selecting the targeted cell genotype, such “customized” nanotherapies also had the potential advantage of improved tumor specificity since normal cells without the targeted gene mutations would not be affected by the treatment.
While patient pre-selection has facilitated the development of subgroup-specific treatments, fully individualized, patient-specific nanotherapy is still the ultimate goal of precision nanomedicine. In this aspect, immunotherapeutic approaches have probably made the most encouraging progress, given that many types of tumor vaccines could be personally prepared using autologous material, including tumor tissue sample and tumor cell-derived extracellular vesicles, as previously discussed [21, 73]. Another example of therapeutic nanoplatform with potential of individualized medicine was the use of tumor cell membrane-decorated nanosystems [168], [169], [170], [171] for drug delivery via the so-called homotypic targeting effect [56, 172]. In a previous study, Ren et al. suggested that patient-derived proteins could also be exploited as individualized surface modification of therapeutic nanomaterials [143]. This conceptual strategy was inspired by the fact that certain disease-related proteins could be specifically enriched in the “corona” formed at the nano-bio interface and this enrichment might lead to particular biological response mediated by nanomaterial-protein interactions. On the other hand, versatile platforms to encapsulate and deliver personalized therapeutic agents with high efficiency are also of great importance. For example, nucleic acid-based therapeutics have huge potential in precision medicine since their sequence could be easily altered to encode patient-specific biological information [173, 174]. An OMV-based anti-tumor nanovaccine was hence developed for rapid and universal loading of mRNA antigens [27]. OMVs were engineered to express an RNA binding protein and a lysosomal escape protein on the membrane, enabling the membrane surface to quickly adsorb any mRNA antigen labeled with a binding sequence that could be specifically recognized by the RNA binding element.
In clinical practice, cancer patients are typically treated with a combination of multiple drugs or therapeutic modalities. Although early-generation nanotherapeutics were mainly single-drug formulations, there is an increasing consensus that improved combination therapy represents a major advantage of nanomedicine and should be exploited maximally [4, 69, 175]. Nanomaterial carriers are able to deliver multiple drugs ratiometrically to the same tissue/cell at the same time, optimizing their synergistic effect. Furthermore, the multifunctional potential of nanocarriers also enable them to release the loaded drug in a temporospatially controlled manner, delivering different drugs to diverse targets or sequentially at a fixed order. Intelligent nanoplatforms should therefore be specifically developed to assist the rational design of new combined treatments that could synthetically deal with the heterogeneous and complicated tumor biology. For instance, although nanomaterial-based delivery has facilitated the application of procoagulant agents in tumor vasculature occlusion therapy, their efficacy could be limited by possible recurrence due to the uneven distribution of blood vessels within the tumor tissue. Co-delivery with chemotherapeutic drug was reported to effectively reduce the recurrence rate in animal models after such blood vessel occlusion and might be preferable to single-modality treatment [176]. In another study, a tumor-targeted nanocarrier was constructed for combined administration of olaparib and JQ1, a drug that could disrupt homologous recombination pathway-dependent DNA repair, against non BRCA-mutant pancreatic cancer [177]. While pancreatic tumors without BRCA mutations were insensitive to PARP inhibitors, the co-application of JQ1 could create DNA repair deficiency in tumor cells to sensitize them towards olaparib. The use of multi-drug loaded nanocarrier significantly enhanced their synergistic effect and reduced the toxicity, which were difficult to be achieved by simultaneous administration of free drugs due to their different pharmacokinetic profiles and non-coordinated biodistribution.
In summary, disease-oriented rational design of nanomedicine for patient-specific and synergistic combinatory therapy has given rise to the major opportunities of next-generation anticancer nanomaterials. Many challenges, however, are still lying ahead on the road of their eventual clinical translation. The enormous design flexibility of intelligent nanomaterials inevitably complicated their industrial manufacturing (cost, scalability, reproducibility, processability) and pharmaceutical properties (in vivo dynamic transformation, interactions with serum proteins, pharmacokinetics, immunogenity, metabolism and clearance). Moreover, while the multifunctionality potential of nanoformulations has rendered them particularly attractive for precision healthcare, such customizability also made it extremely difficult to develop standardized quality control and evaluation protocols (both preclinical and clinical) for highly complex nanosystems. To properly address these challenges, a deeply interdisciplinary endeavor is probably needed, so that both industrial and clinical concerns, including scalable and biosafe fabrication, transportation and storage, quality control procedures, as well as real world scenarios of usage, could be fully incorporated into nanomaterial rational design. On the other hand, since intelligent nanomaterials are often designed to have a more sophisticated therapeutic mechanism compared to conventional drugs, it is also important to develop specific evaluation methods and clinical trial formulas in accord with the system design. In current cancer nanomedicine, patient stratification techniques are mainly applied to identify cases that are more sensitive to the delivered therapeutic agents. It has already been suggested that developing biomarkers to predict the patient-specific functions of nanocarriers, such as extravasation and tumor penetration, would greatly facilitate clinical trials and real-world application of nanoformulations [2].
Perspectives
Intelligent nanomaterials have demonstrated important potential in a wide range of theranostic anti-cancer applications. Through the rapid advancement of material and biological engineering technologies, nanosystems could be designed to perform complex actions in a temporospatially controlled manner, and to encode personalized biological information for patient-specific treatment. Based on the recent advances and emerging trends discussed in previous sections, we suggest that further research endeavor directed at rational design of combination therapies, convergence with patient stratification and individualized drug delivery, as well as precise knowledge of nanomaterials’ in vivo behavior will be highly desirable for the future of cancer nanomedicine. In the end of this review, we would also like to suggest several issues and prospects that can be particularly important for the field of cancer nanomedicine in the next few years.
Programmable synthesis of nanomaterials
Many of the existing problems in clinical translation of nanomedicines derived from our limited ability to construct a complicated nanosystem that precisely correspond to the conceptual design. Developing protocols and techniques to synthesis and fabricate nanomaterials in a programmable and reproducible fashion would greatly facilitate their quality control, characterization and biosafety management, as already preliminarily demonstrated by the promising anti-cancer activities of DNA origami nanostructures. At present, it has become possible to synthesize many simple nanostructures with tunable properties, yet for most kinds of nanomaterials the programmable assembly of multiple functional units into ordered 2D or 3D structure with high spatial addressability remained a challenge. Full control over the synthesis and assembly of nanostructures, moreover, would also benefit the incorporation of many stimuli-responsive functions such as in situ aggregation and time-resolved drug release. In these years, encouraging progresses have been made in nanomaterial synthesis guided by computational simulation and artificial intelligence. Machine learning techniques have been applied to screen optimal protocols for nanomaterial preparation, while molecular simulation tools have enabled us to predict assembly behaviors of macrobiomolecules [159, 162]. These methods can also be used to predict important biomedical properties of new types of nanomaterials with different design and composition [159, 161, 162, 178]. A more profound convergence with in silico techniques would be very helpful in identifying novel strategies to manipulate the synthesis of complex nanosystems.
Nanovaccines
As we have discussed in previous sections, nanovaccines represented one of the most important opportunities for anti-cancer nanomaterials. While we believe that clinically-oriented development should be the universal principle of future nanomedicine, it seems to be particularly relevant in the case of nanovaccines since they are currently among the most promising nanoplatforms in terms of clinical translation [69]. We suggest that in the next few years, research effort should be specifically dedicated to the translational issues of nanomaterials for tumor vaccination, e.g., evaluation methods (preclinical models, characterization standards, patient stratification strategies) particularly suitable for nanomaterial-based vaccines and adjuvants; efficient and safe protocols for acquirement, processing and storage of patient-derived materials; therapeutic regimens for combination with other clinically available immunotherapies or other therapeutic modalities. The success of mRNA vaccines against epidemic diseases could provide important reference and insight on the manufacturing and real-world application of nanoplatform-delivered vaccines.
Regulations and ethical considerations
It has now been vastly recognized that the challenges created by the increasing complexity of nanomaterials are not restricted to scientific contexts. Notably, the expansion of intelligent nanomedicine has introduced an unprecedented number of novel drug types and forms, resulting in an unprecedented need of specific regulations (e.g., industry guidelines, standardized assays, data requirements, and specific approval pathways) that so far has remained largely unmet. Multiple reports have indicated that regulatory uncertainties strongly affected the innovation process in nanomedicine research and development [179, 180]. Particularly, it has been suggested that the lack of data collection protocols, safety and ethical guidelines, and appropriate approval pathways caused researchers and industrial sponsors to heavily rely on clinically approved drugs and technologies, which in turn limited the possibility of disease-oriented innovation [180]. To address this challenge, it seems to us that a collaborative participation of both innovators and regulatory bodies should be firmly pursued. While more scientific effort should be dedicated to the establishment of practical assays and techniques for preclinical and clinical evaluation of anti-cancer nanomaterials, the regulatory bodies need to collaborate with the scientific community and industry to convert the frontier knowledge into detailed and definite protocols or guidelines.
In summary, despite of many encouraging progresses made by state-of-the-art nanotechnology in cancer theranostics, the highly heterogeneous nature of cancers and the complicated nano-bio interactions would always remain extremely demanding challenges for nanoscientists. Only a profound cooperation with clinical, industrial and regulatory forces could further expand the boundaries of nanomedicine, and eventually translate the bright promise of intelligent nanomaterials into concrete clinical success against cancer.
-
Informed consent: Not applicable.
-
Ethical approval: Not applicable.
-
Author contributions: All authors have accepted responsibility for the entire content of this manuscript and approved its submission.
-
Competing interests: Authors declare no conflict of interest.
-
Research funding: This work was supported by the National Basic Research Plan of China (2018YFE0205300), the National Basic Science Center Project (T2288102), and the Key Area Research and Development Program of Guangdong Province (2020B0101020004).
References
1. Wang, J, Li, Y, Nie, G. Multifunctional biomolecule nanostructures for cancer therapy. Nat Rev Mater 2021;6:766–83. https://doi.org/10.1038/s41578-021-00315-x.Search in Google Scholar PubMed PubMed Central
2. Shi, J, Kantoff, PW, Wooster, R, Farokhzad, OC. Cancer nanomedicine: progress, challenges and opportunities. Nat Rev Cancer 2016;17:20. https://doi.org/10.1038/nrc.2016.108.Search in Google Scholar PubMed PubMed Central
3. Chen, H, Zhang, W, Zhu, G, Xie, J, Chen, X. Rethinking cancer nanotheranostics. Nat Rev Mater 2017;2:17024. https://doi.org/10.1038/natrevmats.2017.24.Search in Google Scholar PubMed PubMed Central
4. Nam, J, Son, S, Park, KS, Zou, W, Shea, LD, Moon, JJ. Cancer nanomedicine for combination cancer immunotherapy. Nat Rev Mater 2019;4:398–414. https://doi.org/10.1038/s41578-019-0108-1.Search in Google Scholar
5. Schroeder, A, Heller, DA, Winslow, MM, Dahlman, JE, Pratt, GW, Langer, R, et al.. Treating metastatic cancer with nanotechnology. Nat Rev Cancer 2012;12:39–50. https://doi.org/10.1038/nrc3180.Search in Google Scholar PubMed
6. Anselmo, AC, Mitragotri, S. Nanoparticles in the clinic. Bioeng Transl Med 2016;1:10–29. https://doi.org/10.1002/btm2.10003.Search in Google Scholar PubMed PubMed Central
7. Anselmo, AC, Mitragotri, S. Nanoparticles in the clinic: an update. Bioeng Transl Med 2019;4:e10143. https://doi.org/10.1002/btm2.10143.Search in Google Scholar PubMed PubMed Central
8. Blanco, E, Shen, H, Ferrari, M. Principles of nanoparticle design for overcoming biological barriers to drug delivery. Nat Biotechnol 2015;33:941. https://doi.org/10.1038/nbt.3330.Search in Google Scholar PubMed PubMed Central
9. Zhao, Z, Ukidve, A, Kim, J, Mitragotri, S. Targeting strategies for tissue-specific drug delivery. Cell 2020;181:151–67. https://doi.org/10.1016/j.cell.2020.02.001.Search in Google Scholar PubMed
10. Petersen, GH, Alzghari, SK, Chee, W, Sankari, SS, La-Beck, NM. Meta-analysis of clinical and preclinical studies comparing the anticancer efficacy of liposomal versus conventional non-liposomal doxorubicin. J Contr Release 2016;232:255–64. https://doi.org/10.1016/j.jconrel.2016.04.028.Search in Google Scholar PubMed
11. Wilhelm, S, Tavares, AJ, Dai, Q, Ohta, S, Audet, J, Dvorak, HF, et al.. Analysis of nanoparticle delivery to tumours. Nat Rev Mater 2016;1:16014. https://doi.org/10.1038/natrevmats.2016.14.Search in Google Scholar
12. Agrahari, V, Agrahari, V, Chou, ML, Chew, CH, Noll, J, Burnouf, T. Intelligent micro-/nanorobots as drug and cell carrier devices for biomedical therapeutic advancement: promising development opportunities and translational challenges. Biomaterials 2020;260:120163. https://doi.org/10.1016/j.biomaterials.2020.120163.Search in Google Scholar PubMed
13. Armstrong, JPK, Holme, MN, Stevens, MM. Re-engineering extracellular vesicles as smart nanoscale therapeutics. ACS Nano 2017;11:69–83. https://doi.org/10.1021/acsnano.6b07607.Search in Google Scholar PubMed PubMed Central
14. Habibi, N, Kamaly, N, Memic, A, Shafiee, H. Self-assembled peptide-based nanostructures: smart nanomaterials toward targeted drug delivery. Nano Today 2016;11:41–60. https://doi.org/10.1016/j.nantod.2016.02.004.Search in Google Scholar PubMed PubMed Central
15. Yang, C, Wu, X, Liu, J, Ding, B. Stimuli-responsive nucleic acid nanostructures for efficient drug delivery. Nanoscale 2022;14:17862–70. https://doi.org/10.1039/d2nr05316k.Search in Google Scholar PubMed
16. Goldberg, MS. Improving cancer immunotherapy through nanotechnology. Nat Rev Cancer 2019;19:587–602. https://doi.org/10.1038/s41568-019-0186-9.Search in Google Scholar PubMed
17. Gong, N, Sheppard, NC, Billingsley, MM, June, CH, Mitchell, MJ. Nanomaterials for T-cell cancer immunotherapy. Nat Nanotechnol 2021;16:25–36. https://doi.org/10.1038/s41565-020-00822-y.Search in Google Scholar PubMed
18. Zhuang, J, Holay, M, Park, JH, Fang, RH, Zhang, J, Zhang, LF. Nanoparticle delivery of immunostimulatory agents for cancer immunotherapy. Theranostics 2019;9:7826–48. https://doi.org/10.7150/thno.37216.Search in Google Scholar PubMed PubMed Central
19. Deng, Y, Zhan, W, Liang, G. Intracellular self-assembly of peptide conjugates for tumor imaging and therapy. Adv Healthcare Mater 2021;10:2001211. https://doi.org/10.1002/adhm.202001211.Search in Google Scholar PubMed
20. Li, P, Wang, D, Hu, J, Yang, X. The role of imaging in targeted delivery of nanomedicine for cancer therapy. Adv Drug Deliv Rev 2022;189:114447. https://doi.org/10.1016/j.addr.2022.114447.Search in Google Scholar PubMed
21. Hu, Z, Ott, PA, Wu, CJ. Towards personalized, tumour-specific, therapeutic vaccines for cancer. Nat Rev Immunol 2018;18:168–82. https://doi.org/10.1038/nri.2017.131.Search in Google Scholar PubMed PubMed Central
22. Sahin, U, Türeci, Ö. Personalized vaccines for cancer immunotherapy. Science 2018;359:1355–60. https://doi.org/10.1126/science.aar7112.Search in Google Scholar PubMed
23. Zhou, J, Kroll, AV, Holay, M, Fang, RH, Zhang, L. Biomimetic nanotechnology toward personalized vaccines. Adv Mater 2020;32:1901255. https://doi.org/10.1002/adma.201901255.Search in Google Scholar PubMed PubMed Central
24. Lan, T, Que, H, Luo, M, Zhao, X, Wei, X. Genome editing via non-viral delivery platforms: current progress in personalized cancer therapy. Mol Cancer 2022;21:71. https://doi.org/10.1186/s12943-022-01550-8.Search in Google Scholar PubMed PubMed Central
25. Chen, L, Qin, H, Zhao, R, Zhao, X, Lin, L, Chen, Y, et al.. Bacterial cytoplasmic membranes synergistically enhance the antitumor activity of autologous cancer vaccines. Sci Transl Med 2021;13:eabc2816. https://doi.org/10.1126/scitranslmed.abc2816.Search in Google Scholar PubMed
26. Foy, SP, Jacoby, K, Bota, DA, Hunter, T, Pan, Z, Stawiski, E, et al.. Non-viral precision T cell receptor replacement for personalized cell therapy. Nature 2023;615:687–96. https://doi.org/10.1038/s41586-022-05531-1.Search in Google Scholar PubMed PubMed Central
27. Li, Y, Ma, X, Yue, Y, Zhang, K, Cheng, K, Feng, Q, et al.. Rapid surface display of mRNA antigens by bacteria-derived outer membrane vesicles for a personalized tumor vaccine. Adv Mater 2022;34:2109984. https://doi.org/10.1002/adma.202109984.Search in Google Scholar PubMed
28. Surana, S, Shenoy, AR, Krishnan, Y. Designing DNA nanodevices for compatibility with the immune system of higher organisms. Nat Nanotechnol 2015;10:741–7. https://doi.org/10.1038/nnano.2015.180.Search in Google Scholar PubMed PubMed Central
29. Cai, R, Chen, C. The crown and the scepter: roles of the protein corona in nanomedicine. Adv Mater 2019;31:1805740. https://doi.org/10.1002/adma.201805740.Search in Google Scholar PubMed
30. Yu, Y, Luan, Y, Dai, W. Dynamic process, mechanisms, influencing factors and study methods of protein corona formation. Int J Biol Macromol 2022;205:731–9. https://doi.org/10.1016/j.ijbiomac.2022.03.105.Search in Google Scholar PubMed
31. Lee, D, Huntoon, K, Lux, J, Kim, BYS, Jiang, W. Engineering nanomaterial physical characteristics for cancer immunotherapy. Nat Rev Bioeng 2023;1:499–517. https://doi.org/10.1038/s44222-023-00047-3.Search in Google Scholar
32. Qin, H, Ding, Y, Mujeeb, A, Zhao, Y, Nie, G. Tumor microenvironment targeting and responsive peptide-based nanoformulations for improved tumor therapy. Mol Pharm 2017;92:219–31. https://doi.org/10.1124/mol.116.108084.Search in Google Scholar PubMed
33. Zhang, Y, Li, J, Pu, K. Recent advances in dual- and multi-responsive nanomedicines for precision cancer therapy. Biomaterials 2022;291:121906. https://doi.org/10.1016/j.biomaterials.2022.121906.Search in Google Scholar PubMed
34. Ussia, M, Pumera, M. Towards micromachine intelligence: potential of polymers. Chem Soc Rev 2022;51:1558–72. https://doi.org/10.1039/d1cs00587a.Search in Google Scholar PubMed
35. Guo, Y, Tang, J, Yao, C, Yang, D. Multimodules integrated functional DNA nanomaterials for intelligent drug delivery. Wiley Interdiscip Rev Nanomed Nanobiotechnol 2022;14:e1753. https://doi.org/10.1002/wnan.1753.Search in Google Scholar PubMed
36. Wiraja, C, Zhu, Y, Lio, DCS, Yeo, DC, Xie, M, Fang, WN, et al.. Framework nucleic acids as programmable carrier for transdermal drug delivery. Nat Commun 2019;10:1147. https://doi.org/10.1038/s41467-019-09029-9.Search in Google Scholar PubMed PubMed Central
37. Lee, H, Lytton-Jean, AK, Chen, Y, Love, KT, Park, AI, Karagiannis, ED, et al.. Molecularly self-assembled nucleic acid nanoparticles for targeted in vivo siRNA delivery. Nat Nanotechnol 2012;7:389–93. https://doi.org/10.1038/nnano.2012.73.Search in Google Scholar PubMed PubMed Central
38. Johnson, MB, Halman, JR, Satterwhite, E, Zakharov, AV, Bui, MN, Benkato, K, et al.. Programmable nucleic acid based polygons with controlled neuroimmunomodulatory properties for predictive QSAR modeling. Small 2017;13:1701255. https://doi.org/10.1002/smll.201701255.Search in Google Scholar PubMed PubMed Central
39. Zhao, Y, Dai, X, Wang, F, Zhang, X, Fan, C, Liu, X. Nanofabrication based on DNA nanotechnology. Nano Today 2019;26:123–48. https://doi.org/10.1016/j.nantod.2019.03.004.Search in Google Scholar
40. Gerling, T, Kube, M, Kick, B, Dietz, H. Sequence-programmable covalent bonding of designed DNA assemblies. Sci Adv 2018;4:eaau1157. https://doi.org/10.1126/sciadv.aau1157.Search in Google Scholar PubMed PubMed Central
41. Seeman, NC, Sleiman, HF. DNA nanotechnology. Nat Rev Mater 2018;3:17068. https://doi.org/10.1038/natrevmats.2017.68.Search in Google Scholar
42. Dey, S, Fan, C, Gothelf, KV, Li, J, Lin, C, Liu, L, et al.. DNA origami. Nat Rev Methods Primers 2021;1:13. https://doi.org/10.1038/s43586-020-00009-8.Search in Google Scholar
43. He, L, Mu, J, Gang, O, Chen, X. Rationally programming nanomaterials with DNA for biomedical applications. Adv Sci 2021;8:2003775. https://doi.org/10.1002/advs.202003775.Search in Google Scholar PubMed PubMed Central
44. Li, S, Jiang, Q, Liu, S, Zhang, Y, Tian, Y, Song, C, et al.. A DNA nanorobot functions as a cancer therapeutic in response to a molecular trigger in vivo. Nat Biotechnol 2018;36:258–64. https://doi.org/10.1038/nbt.4071.Search in Google Scholar PubMed
45. Liu, S, Jiang, Q, Zhao, X, Zhao, R, Wang, Y, Wang, Y, et al.. A DNA nanodevice-based vaccine for cancer immunotherapy. Nat Mater 2021;20:421–30. https://doi.org/10.1038/s41563-020-0793-6.Search in Google Scholar PubMed
46. Cong, Y, Ji, L, Gao, YJ, Liu, FH, Cheng, DB, Hu, Z, et al.. Microenvironment-induced in situ self-assembly of polymer–peptide conjugates that attack solid tumors deeply. Angew Chem Int Ed 2019;58:4632–7. https://doi.org/10.1002/anie.201900135.Search in Google Scholar PubMed
47. Gao, J, Zhan, J, Yang, Z. Enzyme-instructed self-assembly (EISA) and hydrogelation of peptides. Adv Mater 2020;32:1805798. https://doi.org/10.1002/adma.201805798.Search in Google Scholar PubMed
48. Yang, J, An, HW, Wang, H. Self-assembled peptide drug delivery systems. ACS Appl Bio Mater 2021;4:24–46. https://doi.org/10.1021/acsabm.0c00707.Search in Google Scholar PubMed
49. Qi, GB, Gao, YJ, Wang, L, Wang, H. Self-assembled peptide-based nanomaterials for biomedical imaging and therapy. Adv Mater 2018;30:1703444. https://doi.org/10.1002/adma.201703444.Search in Google Scholar PubMed
50. Zhang, D, Qi, G-B, Zhao, YX, Qiao, SL, Yang, C, Wang, H. In situ formation of nanofibers from purpurin18-peptide conjugates and the assembly induced retention effect in tumor sites. Adv Mater 2015;27:6125–30. https://doi.org/10.1002/adma.201502598.Search in Google Scholar PubMed
51. An, HW, Li, LL, Wang, Y, Wang, Z, Hou, D, Lin, YX, et al.. A tumour-selective cascade activatable self-detained system for drug delivery and cancer imaging. Nat Commun 2019;10:4861. https://doi.org/10.1038/s41467-019-12848-5.Search in Google Scholar PubMed PubMed Central
52. Wang, MD, Hou, DY, Lv, GT, Li, RX, Hu, XJ, Wang, ZJ, et al.. Targeted in situ self-assembly augments peptide drug conjugate cell-entry efficiency. Biomaterials 2021;278:121139. https://doi.org/10.1016/j.biomaterials.2021.121139.Search in Google Scholar PubMed
53. Hamad-Schifferli, K. Exploiting the novel properties of protein coronas: emerging applications in nanomedicine. Nanomedicine 2015;10:1663–74. https://doi.org/10.2217/nnm.15.6.Search in Google Scholar PubMed
54. Zhen, X, Cheng, P, Pu, K. Recent advances in cell membrane-camouflaged nanoparticles for cancer phototherapy. Small 2019;15:1804105. https://doi.org/10.1002/smll.201804105.Search in Google Scholar PubMed
55. Kroll, AV, Fang, RH, Zhang, L. Biointerfacing and applications of cell membrane-coated nanoparticles. Bioconjugate Chem 2017;28:23–32. https://doi.org/10.1021/acs.bioconjchem.6b00569.Search in Google Scholar PubMed PubMed Central
56. Luk, BT, Zhang, L. Cell membrane-camouflaged nanoparticles for drug delivery. J Contr Release 2015;220:600–7. https://doi.org/10.1016/j.jconrel.2015.07.019.Search in Google Scholar PubMed PubMed Central
57. Agarwal, S, Agarwal, V, Agarwal, M, Singh, M. Exosomes: structure, biogenesis, types and application in diagnosis, and gene and drug delivery. Curr Gene Ther 2020;20:195–206. https://doi.org/10.2174/1566523220999200731011702.Search in Google Scholar PubMed
58. Vader, P, Mol, EA, Pasterkamp, G, Schiffelers, RM. Extracellular vesicles for drug delivery. Adv Drug Deliv Rev 2016;106:148–56. https://doi.org/10.1016/j.addr.2016.02.006.Search in Google Scholar PubMed
59. Zeng, X, Sun, J, Li, S, Shi, J, Gao, H, Sun Leong, W, et al.. Blood-triggered generation of platinum nanoparticle functions as an anti-cancer agent. Nat Commun 2020;11:567. https://doi.org/10.1038/s41467-019-14131-z.Search in Google Scholar PubMed PubMed Central
60. Hu, CM, Zhang, L, Aryal, S, Cheung, C, Fang, RH, Zhang, L. Erythrocyte membrane-camouflaged polymeric nanoparticles as a biomimetic delivery platform. Proc Natl Acad Sci U S A 2011;108:10980. https://doi.org/10.1073/pnas.1106634108.Search in Google Scholar PubMed PubMed Central
61. Wang, J, Zhu, M, Nie, G. Biomembrane-based nanostructures for cancer targeting and therapy: from synthetic liposomes to natural biomembranes and membrane-vesicles. Adv Drug Deliv Rev 2021;178:113974. https://doi.org/10.1016/j.addr.2021.113974.Search in Google Scholar PubMed
62. Fang, RH, Gao, W, Zhang, L. Targeting drugs to tumours using cell membrane-coated nanoparticles. Nat Rev Clin Oncol 2023;20:33–48. https://doi.org/10.1038/s41571-022-00699-x.Search in Google Scholar PubMed
63. Yue, Y, Li, F, Li, Y, Wang, Y, Guo, X, Cheng, Z, et al.. Biomimetic nanoparticles carrying a repolarization agent of tumor-associated macrophages for remodeling of the inflammatory microenvironment following photothermal therapy. ACS Nano 2021;15:15166–79. https://doi.org/10.1021/acsnano.1c05618.Search in Google Scholar PubMed
64. Hu, CM, Fang, RH, Wang, KC, Luk, BT, Thamphiwatana, S, Dehaini, D, et al.. Nanoparticle biointerfacing by platelet membrane cloaking. Nature 2015;526:118. https://doi.org/10.1038/nature15373.Search in Google Scholar PubMed PubMed Central
65. Hu, Q, Qian, C, Sun, W, Wang, J, Chen, Z, Bomba, HN, et al.. Engineered nanoplatelets for enhanced treatment of multiple myeloma and thrombus. Adv Mater 2016;28:9573–80. https://doi.org/10.1002/adma.201603463.Search in Google Scholar PubMed PubMed Central
66. Ying, M, Zhuang, J, Wei, X, Zhang, X, Zhang, Y, Jiang, Y, et al.. Remote-loaded platelet vesicles for disease-targeted delivery of therapeutics. Adv Funct Mater 2018;28:1801032. https://doi.org/10.1002/adfm.201801032.Search in Google Scholar PubMed PubMed Central
67. Wang, C, Sun, W, Ye, Y, Hu, Q, Bomba, HN, Gu, Z. In situ activation of platelets with checkpoint inhibitors for post-surgical cancer immunotherapy. Nat Biomed Eng 2017;1:0011. https://doi.org/10.1038/s41551-016-0011.Search in Google Scholar
68. Waldman, AD, Fritz, JM, Lenardo, MJ. A guide to cancer immunotherapy: from T cell basic science to clinical practice. Nat Rev Immunol 2020;20:651–68. https://doi.org/10.1038/s41577-020-0306-5.Search in Google Scholar PubMed PubMed Central
69. van der Meel, R, Sulheim, E, Shi, Y, Kiessling, F, Mulder, WJM, Lammers, T. Smart cancer nanomedicine. Nat Nanotechnol 2019;14:1007–17. https://doi.org/10.1038/s41565-019-0567-y.Search in Google Scholar PubMed PubMed Central
70. Sanmamed, MF, Chen, L. A paradigm shift in cancer immunotherapy: from enhancement to normalization. Cell 2018;175:313–26. https://doi.org/10.1016/j.cell.2018.09.035.Search in Google Scholar PubMed PubMed Central
71. Nam, J, Son, S, Park, KS, Zou, W, Shea, LD, Moon, JJ. Cancer nanomedicine for combination cancer immunotherapy. Nat Rev Mater 2019;4:398–414. https://doi.org/10.1038/s41578-019-0108-1.Search in Google Scholar
72. Liu, G, Zhu, M, Zhao, X, Nie, G. Nanotechnology-empowered vaccine delivery for enhancing CD8+ T cells-mediated cellular immunity. Adv Drug Deliv Rev 2021;176:113889. https://doi.org/10.1016/j.addr.2021.113889.Search in Google Scholar PubMed
73. Chiang, CL, Coukos, G, Kandalaft, LE. Whole tumor antigen vaccines: where are we? Vaccines 2015;3:344–72. https://doi.org/10.3390/vaccines3020344.Search in Google Scholar PubMed PubMed Central
74. Aikins, ME, Xu, C, Moon, JJ. Engineered nanoparticles for cancer vaccination and immunotherapy. Acc Chem Res 2020;53:2094–105. https://doi.org/10.1021/acs.accounts.0c00456.Search in Google Scholar PubMed PubMed Central
75. Qin, L, Zhang, H, Zhou, Y, Umeshappa, CS, Gao, H. Nanovaccine-based strategies to overcome challenges in the whole vaccination cascade for tumor immunotherapy. Small 2021;17:2006000. https://doi.org/10.1002/smll.202006000.Search in Google Scholar PubMed
76. Qin, H, Zhao, R, Qin, Y, Zhu, J, Chen, L, Di, C, et al.. Development of a cancer vaccine using in vivo click-chemistry-mediated active lymph node accumulation for improved immunotherapy. Adv Mater 2021;33:2006007. https://doi.org/10.1002/adma.202006007.Search in Google Scholar PubMed
77. Abdallah, M, Müllertz, OO, Styles, IK, Mörsdorf, A, Quinn, JF, Whittaker, MR, et al.. Lymphatic targeting by albumin-hitchhiking: applications and optimisation. J Contr Release 2020;327:117–28. https://doi.org/10.1016/j.jconrel.2020.07.046.Search in Google Scholar PubMed
78. Karver, MR, Weissleder, R, Hilderbrand, SA. Bioorthogonal reaction pairs enable simultaneous, selective, multi-target imaging. Angew Chem Int Ed 2012;51:920–2. https://doi.org/10.1002/ange.201104389.Search in Google Scholar
79. Yin, Y, Li, X, Ma, H, Zhang, J, Yu, D, Zhao, R, et al.. In situ transforming RNA nanovaccines from polyethylenimine functionalized graphene oxide hydrogel for durable cancer immunotherapy. Nano Lett 2021;21:2224–31. https://doi.org/10.1021/acs.nanolett.0c05039.Search in Google Scholar PubMed
80. Peng, M, Mo, Y, Wang, Y, Wu, P, Zhang, Y, Xiong, F, et al.. Neoantigen vaccine: an emerging tumor immunotherapy. Mol Cancer 2019;18:128. https://doi.org/10.1186/s12943-019-1055-6.Search in Google Scholar PubMed PubMed Central
81. Cheng, K, Zhao, R, Li, Y, Qi, Y, Wang, Y, Zhang, Y, et al.. Bioengineered bacteria-derived outer membrane vesicles as a versatile antigen display platform for tumor vaccination via Plug-and-Display technology. Nat Commun 2021;12:2041. https://doi.org/10.1038/s41467-021-22308-8.Search in Google Scholar PubMed PubMed Central
82. Kaparakis-Liaskos, M, Ferrero, RL. Immune modulation by bacterial outer membrane vesicles. Nat Rev Immunol 2015;15:375–87. https://doi.org/10.1038/nri3837.Search in Google Scholar PubMed
83. Caruana, JC, Walper, SA. Bacterial membrane vesicles as mediators of microbe – microbe and microbe – host community interactions. Front Microbiol 2020;11:432. https://doi.org/10.3389/fmicb.2020.00432.Search in Google Scholar PubMed PubMed Central
84. Krishnan, N, Kubiatowicz, LJ, Holay, M, Zhou, J, Fang, RH, Zhang, L. Bacterial membrane vesicles for vaccine applications. Adv Drug Deliv Rev 2022;185:114294. https://doi.org/10.1016/j.addr.2022.114294.Search in Google Scholar PubMed
85. Kim, OY, Park, HT, Dinh, NTH, Choi, SJ, Lee, J, Kim, JH, et al.. Bacterial outer membrane vesicles suppress tumor by interferon-γ-mediated antitumor response. Nat Commun 2017;8:626. https://doi.org/10.1038/s41467-017-00729-8.Search in Google Scholar PubMed PubMed Central
86. Li, Y, Zhao, R, Cheng, K, Zhang, K, Wang, Y, Zhang, Y, et al.. Bacterial outer membrane vesicles presenting programmed death 1 for improved cancer immunotherapy via immune activation and checkpoint inhibition. ACS Nano 2020;14:16698–711. https://doi.org/10.1021/acsnano.0c03776.Search in Google Scholar PubMed
87. Schnurr, M, Galambos, P, Scholz, C, Then, F, Dauer, M, Endres, S, et al.. Tumor cell lysate-pulsed human dendritic cells induce a T-cell response against pancreatic carcinoma cells: an in vitro model for the assessment of tumor vaccines. Cancer Res 2001;61:6445–50.Search in Google Scholar
88. André, F, Schartz, NEC, Chaput, N, Flament, C, Raposo, G, Amigorena, S, et al.. Tumor-derived exosomes: a new source of tumor rejection antigens. Vaccine 2002;20:A28–31. https://doi.org/10.1016/s0264-410x(02)00384-5.Search in Google Scholar PubMed
89. Cheung, AS, Koshy, ST, Stafford, AG, Bastings, MMC, Mooney, DJ. Adjuvant-loaded subcellular vesicles derived from disrupted cancer cells for cancer vaccination. Small 2016;12:2321–33. https://doi.org/10.1002/smll.201600061.Search in Google Scholar PubMed PubMed Central
90. Fang, RH, Hu, C-MJ, Luk, BT, Gao, W, Copp, JA, Tai, Y, et al.. Cancer cell membrane-coated nanoparticles for anticancer vaccination and drug delivery. Nano Lett 2014;14:2181–8. https://doi.org/10.1021/nl500618u.Search in Google Scholar PubMed PubMed Central
91. Kroll, AV, Fang, RH, Jiang, Y, Zhou, J, Wei, X, Yu, CL, et al.. Nanoparticulate delivery of cancer cell membrane elicits multiantigenic antitumor immunity. Adv Mater 2017;29:1703969. https://doi.org/10.1002/adma.201703969.Search in Google Scholar PubMed PubMed Central
92. Yang, R, Xu, J, Xu, L, Sun, X, Chen, Q, Zhao, Y, et al.. Cancer cell membrane-coated adjuvant nanoparticles with mannose modification for effective anticancer vaccination. ACS Nano 2018;12:5121–9. https://doi.org/10.1021/acsnano.7b09041.Search in Google Scholar PubMed
93. Ochyl, LJ, Bazzill, JD, Park, C, Xu, Y, Kuai, R, Moon, JJ. PEGylated tumor cell membrane vesicles as a new vaccine platform for cancer immunotherapy. Biomaterials 2018;182:157–66. https://doi.org/10.1016/j.biomaterials.2018.08.016.Search in Google Scholar PubMed PubMed Central
94. Liu, WL, Zou, MZ, Liu, T, Zeng, JY, Li, X, Yu, WY, et al.. Cytomembrane nanovaccines show therapeutic effects by mimicking tumor cells and antigen presenting cells. Nat Commun 2019;10:3199. https://doi.org/10.1038/s41467-019-11157-1.Search in Google Scholar PubMed PubMed Central
95. Zhang, J, Fan, B, Cao, G, Huang, W, Jia, F, Nie, G, et al.. Direct presentation of tumor-associated antigens to induce adaptive immunity by personalized dendritic cell-mimicking nanovaccines. Adv Mater 2022;34:2205950. https://doi.org/10.1002/adma.202205950.Search in Google Scholar PubMed
96. Vela Ramirez, JE, Sharpe, LA, Peppas, NA. Current state and challenges in developing oral vaccines. Adv Drug Deliv Rev 2017;114:116–31. https://doi.org/10.1016/j.addr.2017.04.008.Search in Google Scholar PubMed PubMed Central
97. Parodi, A, Buzaeva, P, Nigovora, D, Baldin, A, Kostyushev, D, Chulanov, V, et al.. Nanomedicine for increasing the oral bioavailability of cancer treatments. J Nanobiotechnol 2021;19:354. https://doi.org/10.1186/s12951-021-01100-2.Search in Google Scholar PubMed PubMed Central
98. Zimmermann, P, Curtis, N. The influence of the intestinal microbiome on vaccine responses. Vaccine 2018;36:4433–9. https://doi.org/10.1016/j.vaccine.2018.04.066.Search in Google Scholar PubMed
99. Yue, Y, Xu, J, Li, Y, Cheng, K, Feng, Q, Ma, X, et al.. Antigen-bearing outer membrane vesicles as tumour vaccines produced in situ by ingested genetically engineered bacteria. Nat Biomed Eng 2022;6:898–909. https://doi.org/10.1038/s41551-022-00886-2.Search in Google Scholar PubMed
100. Stentz, R, Carvalho, AL, Jones, EJ, Carding, SR. Fantastic voyage: the journey of intestinal microbiota-derived microvesicles through the body. Biochem Soc Trans 2018;46:1021–7. https://doi.org/10.1042/bst20180114.Search in Google Scholar
101. Joyce, JA, Fearon, DT. T cell exclusion, immune privilege, and the tumor microenvironment. Science 2015;348:74–80. https://doi.org/10.1126/science.aaa6204.Search in Google Scholar PubMed
102. Liu, J, Chen, Q, Feng, L, Liu, Z. Nanomedicine for tumor microenvironment modulation and cancer treatment enhancement. Nano Today 2018;21:55–73. https://doi.org/10.1016/j.nantod.2018.06.008.Search in Google Scholar
103. Wang, C, Sun, W, Wright, G, Wang, AZ, Gu, Z. Inflammation-triggered cancer immunotherapy by programmed delivery of CpG and anti-PD1 antibody. Adv Mater 2016;28:8912–20. https://doi.org/10.1002/adma.201506312.Search in Google Scholar PubMed PubMed Central
104. Han, X, Cheng, K, Xu, Y, Wang, Y, Min, H, Zhang, Y, et al.. Modularly designed peptide nanoprodrug augments antitumor immunity of PD-L1 checkpoint blockade by targeting indoleamine 2,3-dioxygenase. J Am Chem Soc 2020;142:2490–6. https://doi.org/10.1021/jacs.9b12232.Search in Google Scholar PubMed
105. Cheng, K, Ding, Y, Zhao, Y, Ye, S, Zhao, X, Zhang, Y, et al.. Sequentially responsive therapeutic peptide assembling nanoparticles for dual-targeted cancer immunotherapy. Nano Lett 2018;18:3250–8. https://doi.org/10.1021/acs.nanolett.8b01071.Search in Google Scholar PubMed
106. Feng, Q, Ma, X, Cheng, K, Liu, G, Li, Y, Yue, Y, et al.. Engineered bacterial outer membrane vesicles as controllable two-way adaptors to activate macrophage phagocytosis for improved tumor immunotherapy. Adv Mater 2022;34:2206200. https://doi.org/10.1002/adma.202206200.Search in Google Scholar PubMed
107. Ngambenjawong, C, Gustafson, HH, Pun, SH. Progress in tumor-associated macrophage (TAM)-targeted therapeutics. Adv Drug Deliv Rev 2017;114:206–21. https://doi.org/10.1016/j.addr.2017.04.010.Search in Google Scholar PubMed PubMed Central
108. Shields, CW, Evans, MA, Wang, LL-W, Baugh, N, Iyer, S, Wu, D, et al.. Cellular backpacks for macrophage immunotherapy. Sci Adv 2020;6:eaaz6579. https://doi.org/10.1126/sciadv.aaz6579.Search in Google Scholar PubMed PubMed Central
109. Yu, GT, Rao, L, Wu, H, Yang, LL, Bu, LL, Deng, WW, et al.. Myeloid-derived suppressor cell membrane-coated magnetic nanoparticles for cancer theranostics by inducing macrophage polarization and synergizing immunogenic cell death. Adv Funct Mater 2018;28:1801389. https://doi.org/10.1002/adfm.201801389.Search in Google Scholar
110. Lin, YX, Wang, Y, Ding, J, Jiang, A, Wang, J, Yu, M, et al.. Reactivation of the tumor suppressor PTEN by mRNA nanoparticles enhances antitumor immunity in preclinical models. Sci Transl Med 2021;13:eaba9772. https://doi.org/10.1126/scitranslmed.aba9772.Search in Google Scholar PubMed PubMed Central
111. Chen, J, Fu, S, Zhang, C, Liu, H, Su, X. DNA logic circuits for cancer theranostics. Small 2022;18:2108008. https://doi.org/10.1002/smll.202108008.Search in Google Scholar PubMed
112. Shi, T, Huang, C, Li, Y, Huang, F, Yin, S. NIR-II phototherapy agents with aggregation-induced emission characteristics for tumor imaging and therapy. Biomaterials 2022;285:121535. https://doi.org/10.1016/j.biomaterials.2022.121535.Search in Google Scholar PubMed
113. Zhang, NY, Hu, XJ, An, HW, Liang, JX, Wang, H. Programmable design and self assembly of peptide conjugated AIEgens for biomedical applications. Biomaterials 2022;287:121655. https://doi.org/10.1016/j.biomaterials.2022.121655.Search in Google Scholar PubMed
114. Zhao, XX, Li, LL, Zhao, Y, An, HW, Cai, Q, Lang, JY, et al.. In situ self-assembled nanofibers precisely target cancer-associated fibroblasts for improved tumor imaging. Angew Chem Int Ed 2019;58:15287–94. https://doi.org/10.1002/ange.201908185.Search in Google Scholar
115. Zhou, H, Guo, M, Li, J, Qin, F, Wang, Y, Liu, T, et al.. Hypoxia-triggered self-assembly of ultrasmall iron oxide nanoparticles to amplify the imaging signal of a tumor. J Am Chem Soc 2021;143:1846–53. https://doi.org/10.1021/jacs.0c10245.Search in Google Scholar PubMed
116. Li, L, Xing, H, Zhang, J, Lu, Y. Functional DNA molecules enable selective and stimuli-responsive nanoparticles for biomedical applications. Acc Chem Res 2019;52:2415–26. https://doi.org/10.1021/acs.accounts.9b00167.Search in Google Scholar PubMed PubMed Central
117. Xu, X, Shang, Y, Liu, F, Jiang, Q, Ding, B. Logic devices based on nucleic acid self-assembly. InfoMat 2021;3:1070–82. https://doi.org/10.1002/inf2.12240.Search in Google Scholar
118. Shao, Y, Zhao, J, Yuan, J, Zhao, Y, Li, L. Organelle-specific photoactivation of DNA nanosensors for precise profiling of subcellular enzymatic activity. Angew Chem Int Ed 2021;60:8923–31. https://doi.org/10.1002/anie.202016738.Search in Google Scholar PubMed
119. Zhao, J, Li, Z, Shao, Y, Hu, W, Li, L. Spatially selective imaging of mitochondrial micrornas via optically programmable strand displacement reactions. Angew Chem Int Ed 2021;60:17937–41. https://doi.org/10.1002/ange.202105696.Search in Google Scholar
120. Xiang, Z, Zhao, J, Qu, J, Song, J, Li, L. A multivariate-gated DNA nanodevice for spatioselective imaging of pro-metastatic targets in extracellular microenvironment. Angew Chem Int Ed 2022;61:e202111836. https://doi.org/10.1002/ange.202111836.Search in Google Scholar
121. Bobo, D, Robinson, KJ, Islam, J, Thurecht, KJ, Corrie, SR. Nanoparticle-based medicines: a review of FDA-approved materials and clinical trials to date. Pharm Res 2016;33:2373–87. https://doi.org/10.1007/s11095-016-1958-5.Search in Google Scholar PubMed
122. Metselaar, JM, Lammers, T. Challenges in nanomedicine clinical translation. Drug Deliv Transl Res 2020;10:721–5. https://doi.org/10.1007/s13346-020-00740-5.Search in Google Scholar PubMed PubMed Central
123. Nel, AE, Mädler, L, Velegol, D, Xia, T, Hoek, EMV, Somasundaran, P, et al.. Understanding biophysicochemical interactions at the nano–bio interface. Nat Mater 2009;8:543–57. https://doi.org/10.1038/nmat2442.Search in Google Scholar PubMed
124. Zhu, M, Nie, G, Meng, H, Xia, T, Nel, A, Zhao, Y. Physicochemical properties determine nanomaterial cellular uptake, transport, and fate. Acc Chem Res 2013;46:622–31. https://doi.org/10.1021/ar300031y.Search in Google Scholar PubMed PubMed Central
125. Carril, M, Padro, D, del Pino, P, Carrillo-Carrion, C, Gallego, M, Parak, WJ. In situ detection of the protein corona in complex environments. Nat Commun 2017;8:1542. https://doi.org/10.1038/s41467-017-01826-4.Search in Google Scholar PubMed PubMed Central
126. Kari, OK, Ndika, J, Parkkila, P, Louna, A, Lajunen, T, Puustinen, A, et al.. In situ analysis of liposome hard and soft protein corona structure and composition in a single label-free workflow. Nanoscale 2020;12:1728–41. https://doi.org/10.1039/c9nr08186k.Search in Google Scholar PubMed
127. Sanchez-Guzman, D, Giraudon-Colas, G, Marichal, L, Boulard, Y, Wien, F, Degrouard, J, et al.. In situ analysis of weakly bound proteins reveals molecular basis of soft corona formation. ACS Nano 2020;14:9073–88. https://doi.org/10.1021/acsnano.0c04165.Search in Google Scholar PubMed
128. Weiss, ACG, Krüger, K, Besford, QA, Schlenk, M, Kempe, K, Förster, S, et al.. In situ characterization of protein corona formation on silica microparticles using confocal laser scanning microscopy combined with microfluidics. ACS Appl Mater Interfaces 2019;11:2459–69. https://doi.org/10.1021/acsami.8b14307.Search in Google Scholar PubMed
129. Baimanov, D, Wang, J, Zhang, J, Liu, K, Cong, Y, Shi, X, et al.. In situ analysis of nanoparticle soft corona and dynamic evolution. Nat Commun 2022;13:5389. https://doi.org/10.1038/s41467-022-33044-y.Search in Google Scholar PubMed PubMed Central
130. Sanchez-Cano, C, Alvarez-Puebla, RA, Abendroth, JM, Beck, T, Blick, R, Cao, Y, et al.. X-ray-based techniques to study the nano–bio interface. ACS Nano 2021;15:3754–807. https://doi.org/10.1021/acsnano.0c09563.Search in Google Scholar PubMed PubMed Central
131. Wang, B, Feng, W, Chai, Z, Zhao, Y. Probing the interaction at nano-bio interface using synchrotron radiation-based analytical techniques. Sci China Chem 2015;58:768–79. https://doi.org/10.1007/s11426-015-5394-x.Search in Google Scholar
132. Wang, L, Zhang, T, Li, P, Huang, W, Tang, J, Wang, P, et al.. Use of synchrotron radiation-analytical techniques to reveal chemical origin of silver-nanoparticle cytotoxicity. ACS Nano 2015;9:6532–47. https://doi.org/10.1021/acsnano.5b02483.Search in Google Scholar PubMed
133. Lam, YY, Hawley, A, Tan, A, Boyd, BJ. Coupling in vitro cell culture with synchrotron SAXS to understand the bio-interaction of lipid-based liquid crystalline nanoparticles with vascular endothelial cells. Drug Deliv Transl Res 2020;10:610–20. https://doi.org/10.1007/s13346-020-00718-3.Search in Google Scholar PubMed
134. Liang, K, Richardson, JJ, Cui, J, Caruso, F, Doonan, CJ, Falcaro, P. Metal–organic framework coatings as cytoprotective exoskeletons for living cells. Adv Mater 2016;28:7910–4. https://doi.org/10.1002/adma.201602335.Search in Google Scholar PubMed
135. Wang, Y, Xie, Y, Luo, J, Guo, M, Hu, X, Chen, X, et al.. Engineering a self-navigated MnARK nanovaccine for inducing potent protective immunity against novel coronavirus. Nano Today 2021;38:101139. https://doi.org/10.1016/j.nantod.2021.101139.Search in Google Scholar PubMed PubMed Central
136. Cao, M, Cai, R, Zhao, L, Guo, M, Wang, L, Wang, Y, et al.. Molybdenum derived from nanomaterials incorporates into molybdenum enzymes and affects their activities in vivo. Nat Nanotechnol 2021;16:708–16. https://doi.org/10.1038/s41565-021-00856-w.Search in Google Scholar PubMed
137. Majumdar, S, Keller, AA. Omics to address the opportunities and challenges of nanotechnology in agriculture. Crit Rev Environ Sci Technol 2021;51:2595–636. https://doi.org/10.1080/10643389.2020.1785264.Search in Google Scholar
138. Shin, TH, Nithiyanandam, S, Lee, DY, Kwon, DH, Hwang, JS, Kim, SG, et al.. Analysis of nanotoxicity with integrated omics and mechanobiology. Nanomaterials 2021;11:2385. https://doi.org/10.3390/nano11092385.Search in Google Scholar PubMed PubMed Central
139. Cai, R, Ren, J, Guo, M, Wei, T, Liu, Y, Xie, C, et al.. Dynamic intracellular exchange of nanomaterials’ protein corona perturbs proteostasis and remodels cell metabolism. Proc Natl Acad Sci U S A 2022;119:e2200363119. https://doi.org/10.1073/pnas.2200363119.Search in Google Scholar PubMed PubMed Central
140. Zhang, Z, Guan, J, Jiang, Z, Yang, Y, Liu, J, Hua, W, et al.. Brain-targeted drug delivery by manipulating protein corona functions. Nat Commun 2019;10:3561. https://doi.org/10.1038/s41467-019-11593-z.Search in Google Scholar PubMed PubMed Central
141. Cheng, Q, Wei, T, Farbiak, L, Johnson, LT, Dilliard, SA, Siegwart, DJ. Selective organ targeting (SORT) nanoparticles for tissue-specific mRNA delivery and CRISPR–Cas gene editing. Nat Nanotechnol 2020;15:313–20. https://doi.org/10.1038/s41565-020-0669-6.Search in Google Scholar PubMed PubMed Central
142. Liang, Q, Bie, N, Yong, T, Tang, K, Shi, X, Wei, Z, et al.. The softness of tumour-cell-derived microparticles regulates their drug-delivery efficiency. Nat Biomed Eng 2019;3:729–40. https://doi.org/10.1038/s41551-019-0405-4.Search in Google Scholar PubMed
143. Ren, J, Cai, R, Wang, J, Daniyal, M, Baimanov, D, Liu, Y, et al.. Precision nanomedicine development based on specific opsonization of human cancer patient-personalized protein coronas. Nano Lett 2019;19:4692–701. https://doi.org/10.1021/acs.nanolett.9b01774.Search in Google Scholar PubMed
144. Guo, M, Zhao, L, Liu, J, Wang, X, Yao, H, Chang, X, et al.. The underlying function and structural organization of the intracellular protein corona on graphdiyne oxide nanosheet for local immunomodulation. Nano Lett 2021;21:6005–13. https://doi.org/10.1021/acs.nanolett.1c01048.Search in Google Scholar PubMed
145. Prakash, K, Diederich, B, Heintzmann, R, Schermelleh, L. Super-resolution microscopy: a brief history and new avenues. Philos Trans Royal Soc A 2022;380:20210110. https://doi.org/10.1098/rsta.2021.0110.Search in Google Scholar PubMed PubMed Central
146. Bond, C, Santiago-Ruiz, AN, Tang, Q, Lakadamyali, M. Technological advances in super-resolution microscopy to study cellular processes. Mol Cell 2022;82:315–32. https://doi.org/10.1016/j.molcel.2021.12.022.Search in Google Scholar PubMed PubMed Central
147. Verveer, PJ, Swoger, J, Pampaloni, F, Greger, K, Marcello, M, Stelzer, EHK. High-resolution three-dimensional imaging of large specimens with light sheet–based microscopy. Nat Methods 2007;4:311–3. https://doi.org/10.1038/nmeth1017.Search in Google Scholar PubMed
148. Han, M, Bushong, EA, Segawa, M, Tiard, A, Wong, A, Brady, MR, et al.. Spatial mapping of mitochondrial networks and bioenergetics in lung cancer. Nature 2023;615:712–9. https://doi.org/10.1038/s41586-023-05793-3.Search in Google Scholar PubMed PubMed Central
149. Zong, W, Wu, R, Li, M, Hu, Y, Li, Y, Li, J, et al.. Fast high-resolution miniature two-photon microscopy for brain imaging in freely behaving mice. Nat Methods 2017;14:713–9. https://doi.org/10.1038/nmeth.4305.Search in Google Scholar PubMed
150. Xie, B, Pu, Y, Yang, F, Chen, W, Yue, W, Ma, J, et al.. Proteomic mapping and targeting of mitotic pericentriolar material in tumors bearing centrosome amplification. Cancer Res 2022;82:2576–92. https://doi.org/10.1158/0008-5472.can-22-0225.Search in Google Scholar PubMed
151. Bosch, JA, Chen, CL, Perrimon, N. Proximity-dependent labeling methods for proteomic profiling in living cells: an update. Wiley Interdiscip Rev Dev Biol 2021;10:e392. https://doi.org/10.1002/wdev.392.Search in Google Scholar PubMed PubMed Central
152. Zhang, Z, Ren, J, Dai, W, Zhang, H, Wang, X, He, B, et al.. Fast and dynamic mapping of the protein corona on nanoparticle surfaces by photocatalytic proximity labeling. Adv Mater 2023;35:2206636. https://doi.org/10.1002/adma.202206636.Search in Google Scholar PubMed
153. Stavrou, M, Phung, N, Grimm, J, Andreou, C. Organ-on-chip systems as a model for nanomedicine. Nanoscale 2023;15:9927–40. https://doi.org/10.1039/d3nr01661g.Search in Google Scholar PubMed PubMed Central
154. Nabi, SU, Ali, SI, Rather, MA, Sheikh, WM, Altaf, M, Singh, H, et al.. Organoids: a new approach in toxicity testing of nanotherapeutics. J Appl Toxicol 2022;42:52–72. https://doi.org/10.1002/jat.4206.Search in Google Scholar PubMed
155. Chen, X, Zhang, YS, Zhang, X, Liu, C. Organ-on-a-chip platforms for accelerating the evaluation of nanomedicine. Bioact Mater 2021;6:1012–27. https://doi.org/10.1016/j.bioactmat.2020.09.022.Search in Google Scholar PubMed PubMed Central
156. Wang, HF, Ran, R, Liu, Y, Hui, Y, Zeng, B, Chen, D, et al.. Tumor-vasculature-on-a-chip for investigating nanoparticle extravasation and tumor accumulation. ACS Nano 2018;12:11600–9. https://doi.org/10.1021/acsnano.8b06846.Search in Google Scholar PubMed
157. Zhang, W, Mehta, A, Tong, Z, Esser, L, Voelcker, NH. Development of polymeric nanoparticles for blood–brain barrier transfer—strategies and challenges. Adv Sci 2021;8:2003937. https://doi.org/10.1002/advs.202003937.Search in Google Scholar PubMed PubMed Central
158. Gakis, GP, Aviziotis, IG, Charitidis, CA. Metal and metal oxide nanoparticle toxicity: moving towards a more holistic structure-activity approach. Environ Sci Nano 2023;10:761–80. https://doi.org/10.1039/d2en00897a.Search in Google Scholar
159. Xu, X, Liu, A, Liu, S, Ma, Y, Zhang, X, Zhang, M, et al.. Application of molecular dynamics simulation in self-assembled cancer nanomedicine. Biomater Res 2023;27:39. https://doi.org/10.1186/s40824-023-00386-7.Search in Google Scholar PubMed PubMed Central
160. Adir, O, Poley, M, Chen, G, Froim, S, Krinsky, N, Shklover, J, et al.. Integrating artificial intelligence and nanotechnology for precision cancer medicine. Adv Mater 2020;32:1901989. https://doi.org/10.1002/adma.201901989.Search in Google Scholar PubMed PubMed Central
161. Yang, RX, McCandler, CA, Andriuc, O, Siron, M, Woods-Robinson, R, Horton, MK, et al.. Big data in a nano world: a review on computational, data-driven design of nanomaterials structures, properties, and synthesis. ACS Nano 2022;16:19873–91. https://doi.org/10.1021/acsnano.2c08411.Search in Google Scholar PubMed PubMed Central
162. Tao, H, Wu, T, Aldeghi, M, Wu, TC, Aspuru-Guzik, A, Kumacheva, E. Nanoparticle synthesis assisted by machine learning. Nat Rev Mater 2021;6:701–16. https://doi.org/10.1038/s41578-021-00337-5.Search in Google Scholar
163. Sengottiyan, S, Mikolajczyk, A, Jagiełło, K, Swirog, M, Puzyn, T. Core, coating, or corona? The importance of considering protein coronas in nano-QSPR modeling of zeta potential. ACS Nano 2023;17:1989–97. https://doi.org/10.1021/acsnano.2c06977.Search in Google Scholar PubMed PubMed Central
164. Bulbake, U, Doppalapudi, S, Kommineni, N, Khan, W. Liposomal formulations in clinical use: an updated review. Pharmaceutics 2017;9:12. https://doi.org/10.3390/pharmaceutics9020012.Search in Google Scholar PubMed PubMed Central
165. Hare, JI, Lammers, T, Ashford, MB, Puri, S, Storm, G, Barry, ST. Challenges and strategies in anti-cancer nanomedicine development: an industry perspective. Adv Drug Deliv Rev 2017;108:25–38. https://doi.org/10.1016/j.addr.2016.04.025.Search in Google Scholar PubMed
166. Du, C, Qi, Y, Zhang, Y, Wang, Y, Zhao, X, Min, H, et al.. Epidermal growth factor receptor-targeting peptide nanoparticles simultaneously deliver gemcitabine and olaparib to treat pancreatic cancer with breast cancer 2 (BRCA2) mutation. ACS Nano 2018;12:10785–96. https://doi.org/10.1021/acsnano.8b01573.Search in Google Scholar PubMed
167. Ebeid, K, Meng, X, Thiel, KW, Do, AV, Geary, SM, Morris, AS, et al.. Synthetically lethal nanoparticles for treatment of endometrial cancer. Nat Nanotechnol 2018;13:72–81. https://doi.org/10.1038/s41565-017-0009-7.Search in Google Scholar PubMed PubMed Central
168. Chen, Z, Zhao, P, Luo, Z, Zheng, M, Tian, H, Gong, P, et al.. Cancer cell membrane–biomimetic nanoparticles for homologous-targeting dual-modal imaging and photothermal therapy. ACS Nano 2016;10:10049–57. https://doi.org/10.1021/acsnano.6b04695.Search in Google Scholar PubMed
169. Sun, H, Su, J, Meng, Q, Yin, Q, Chen, L, Gu, W, et al.. Cancer-cell-biomimetic nanoparticles for targeted therapy of homotypic tumors. Adv Mater 2016;28:9581–8. https://doi.org/10.1002/adma.201602173.Search in Google Scholar PubMed
170. Min, H, Wang, J, Qi, Y, Zhang, Y, Han, X, Xu, Y, et al.. Biomimetic metal–organic framework nanoparticles for cooperative combination of antiangiogenesis and photodynamic therapy for enhanced efficacy. Adv Mater 2019;31:1808200. https://doi.org/10.1002/adma.201808200.Search in Google Scholar PubMed
171. Zhang, JJ, Lin, Y, Zhou, H, He, H, Ma, JJ, Luo, MY, et al.. Cell membrane-camouflaged NIR II fluorescent Ag2Te quantum dots-based nanobioprobes for enhanced in vivo homotypic tumor imaging. Adv Healthcare Mater 2019;8:1900341. https://doi.org/10.1002/adhm.201900341.Search in Google Scholar PubMed
172. Khaldoyanidi, SK, Glinsky, VV, Sikora, L, Glinskii, AB, Mossine, VV, Quinn, TP, et al.. MDA-MB-435 human breast carcinoma cell homo- and heterotypic adhesion under flow conditions is mediated in part by Thomsen-Friedenreich antigen-galectin-3 interactions. J Biol Chem 2003;278:4127–34. https://doi.org/10.1074/jbc.m209590200.Search in Google Scholar
173. Bulaklak, K, Gersbach, CA. The once and future gene therapy. Nat Commun 2020;11:5820. https://doi.org/10.1038/s41467-020-19505-2.Search in Google Scholar PubMed PubMed Central
174. Manzari, MT, Shamay, Y, Kiguchi, H, Rosen, N, Scaltriti, M, Heller, DA. Targeted drug delivery strategies for precision medicines. Nat Rev Mater 2021;6:351–70. https://doi.org/10.1038/s41578-020-00269-6.Search in Google Scholar PubMed PubMed Central
175. Lammers, T, Kiessling, F, Ashford, M, Hennink, W, Crommelin, D, Storm, G. Cancer nanomedicine: is targeting our target? Nat Rev Mater 2016;1:16069. https://doi.org/10.1038/natrevmats.2016.69.Search in Google Scholar PubMed PubMed Central
176. Li, S, Zhang, Y, Ho, SH, Li, B, Wang, M, Deng, X, et al.. Combination of tumour-infarction therapy and chemotherapy via the co-delivery of doxorubicin and thrombin encapsulated in tumour-targeted nanoparticles. Nat Biomed Eng 2020;4:732–42. https://doi.org/10.1038/s41551-020-0573-2.Search in Google Scholar PubMed
177. Wang, Y, Du, C, Zhao, Y, Nie, G, Yang, Y. Trap and kill strategy for non-BRCA mutant pancreatic cancer by co-delivery of olaparib and JQ1 with plectin-1 targeting peptide nanoparticles. Nano Today 2020;33:100877. https://doi.org/10.1016/j.nantod.2020.100877.Search in Google Scholar
178. Jones, DE, Ghandehari, H, Facelli, JC. A review of the applications of data mining and machine learning for the prediction of biomedical properties of nanoparticles. Comput Methods Progr Biomed 2016;132:93–103. https://doi.org/10.1016/j.cmpb.2016.04.025.Search in Google Scholar PubMed PubMed Central
179. Kwon, S, Youtie, J, Porter, A, Newman, N. How does regulatory uncertainty shape the innovation process? Evidence from the case of nanomedicine. J Technol Tran 2022. https://doi.org/10.1007/s10961-022-09980-8.Search in Google Scholar
180. Zingg, R, Fischer, M. The consolidation of nanomedicine. Wiley Interdiscip Rev Nanomed Nanobiotechnol 2019;11:e1569. https://doi.org/10.1002/wnan.1569.Search in Google Scholar PubMed PubMed Central
© 2023 the author(s), published by De Gruyter, Berlin/Boston
This work is licensed under the Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 International License.
Articles in the same Issue
- Frontmatter
- Reviews
- Current state of stem cell research in non-human primates: an overview
- Current status and future perspective of natural killer cell therapy for cancer
- Intelligent nanomaterials for cancer therapy: recent progresses and future possibilities
- Perspectives
- Stem cell-derived embryo models: a frontier of human embryology
- Human stem cell modeling of neuropsychiatric disorders: from polygenicity to convergence
- Optimizing health-span: advances in stem cell medicine and longevity research
- Enrichment of nano delivery platforms for mRNA-based nanotherapeutics
Articles in the same Issue
- Frontmatter
- Reviews
- Current state of stem cell research in non-human primates: an overview
- Current status and future perspective of natural killer cell therapy for cancer
- Intelligent nanomaterials for cancer therapy: recent progresses and future possibilities
- Perspectives
- Stem cell-derived embryo models: a frontier of human embryology
- Human stem cell modeling of neuropsychiatric disorders: from polygenicity to convergence
- Optimizing health-span: advances in stem cell medicine and longevity research
- Enrichment of nano delivery platforms for mRNA-based nanotherapeutics