Next Article in Journal
Enhancing Bioactivity and Conjugation in Green Coffee Bean (Coffea arabica) Extract through Cold Plasma Treatment: Insights into Antioxidant Activity and Phenolic–Protein Conjugates
Next Article in Special Issue
Multistage Self-Assembled Nanomaterials for Cancer Immunotherapy
Previous Article in Journal
The Hybrid Nano-Biointerface between Proteins/Peptides and Two-Dimensional Nanomaterials
Previous Article in Special Issue
The Anti-Arthritic Activity of Diclofenac Lipid-Core Nanocapsules: Stereological Analysis Showing More Protection of Deep Joint Components
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Advance Progress in Assembly Mechanisms of Carrier-Free Nanodrugs for Cancer Treatment

1
State Key Laboratory of Organic-Inorganic Composites, Bei**g Laboratory of Biomedical Materials, Bei**g University of Chemical Technology, Bei**g 100029, China
2
Department of Hepatology, Tongliao Infectious Disease Hospital, Tongliao 028000, China
3
Department of Interventional Ultrasound, PLA Medical College & Fifth Medical Center of Chinese PLA General Hospital, Bei**g 100039, China
*
Authors to whom correspondence should be addressed.
Molecules 2023, 28(20), 7065; https://doi.org/10.3390/molecules28207065
Submission received: 21 August 2023 / Revised: 29 September 2023 / Accepted: 10 October 2023 / Published: 13 October 2023
(This article belongs to the Special Issue Targeted Nanomedicine Design)

Abstract

:
Nanocarriers have been widely studied and applied in the field of cancer treatment. However, conventional nanocarriers still suffer from complicated preparation processes, low drug loading, and potential toxicity of carriers themselves. To tackle the hindrance, carrier-free nanodrugs with biological activity have received increasing attention in cancer therapy. Extensive efforts have been made to exploit new self-assembly methods and mechanisms to expand the scope of carrier-free nanodrugs with enhanced therapeutic performance. In this review, we summarize the advanced progress and applications of carrier-free nanodrugs based on different types of assembly mechanisms and strategies, which involved noncovalent interactions, a combination of covalent bonds and noncovalent interactions, and metal ions-coordinated self-assembly. These carrier-free nanodrugs are introduced in detail according to their assembly and antitumor applications. Finally, the prospects and existing challenges of carrier-free nanodrugs in future development and clinical application are discussed. We hope that this comprehensive review will provide new insights into the rational design of more effective carrier-free nanodrug systems and advancing clinical cancer and other diseases (e.g., bacterial infections) infection treatment.

Graphical Abstract

1. Introduction

Cancer possesses has a high global incidence rate and mortality, which is extremely harmful to human life and health [1]. Traditional chemotherapy (CT) relying on anticancer drugs is the most utilized clinical therapeutic modality [2,3]. However, anticancer drugs can only achieve limited success, they still face challenges including poor solubility, low bioavailability, severely toxic side effects, insufficient stability, and rapid blood clearance, which pose grievous barriers to cancer treatment [4]. To address the aforementioned challenges, numerous innovative and elaborately designed therapeutic drug delivery technologies have been devoted to improving cancer therapy efficiency.
With the development of nanotechnology, nanocarrier-based drug delivery systems (DDS) offer advanced opportunities for cancer diagnosis and therapy [5,6]. In general, traditional nanocarriers include polymer micelles [7], dendrimers [8], liposomes [9] and inorganic nanoparticles [10]. These nanocarriers can significantly increase stability, improve biocompatibility, prolong blood circulation, and enhance tumor accumulation and the therapeutic efficacy of the encapsulated active drugs [11,12], which distinctly facilitate the development of nanomedicines. Nevertheless, the relevant nanocarrier-based DDS still face some common challenges, such as complicated preparation or purification processes, lack of active functionality, low drug loading capacity and incomplete drug release, as well as the potential systemic toxicity of carriers-themselves, which hinder the scaled-up production and practical use of nanocarrier-based nanomedicines [13,14]. Rational designing advanced nano-DDSs for efficient cancer and other related disease therapy still remains challenging.
Carrier-free nanodrugs emerge as smart nanodrug delivery platforms for cancer treatment. Commonly, carrier-free nanodrugs are fabricated via the self-assembly or co-assembly of several drug active ingredients without or with little use of inert materials, in which the active ingredients are mainly drugs with different therapeutic mechanisms [15], including chemotherapeutic drugs [16], photosensitizers [17], targeting agents [18], antibodies to certain proteins [19], genes [20], and other functional biomolecules. This endows carrier-free nanodrugs with desirable advantages over nanocarrier-based DDS: (ⅰ) simple and green assembly process; (ⅱ) integration of nano-based diagnosis and therapy; (ⅲ) inhibition of multidrug resistance (MDR); (ⅳ) synergistic therapeutic effects superior to multidrug combinations with equal-dose; (ⅴ) high drug loading capacity; (ⅵ) low toxicity and good pharmacokinetics [21,22,23]. Therefore, carrier-free nanodrugs can compensate for the shortcomings of traditional nanocarriers, achieving efficient drug utilization and synergistic therapy effects.
Based on the above distinct benefits, a series of assembly mechanisms and therapeutic strategies have been proposed to achieve better therapeutic outcomes. Currently, the assembly mechanisms and strategies during the preparation of carrier-free nanodrugs are mainly classified into the following categories. First, drug molecules or functional biomolecules with special physical properties, chemical structure, and aggregation behavior automatically organize themselves into structurally well-defined nanodrugs driven by noncovalent interactions. These noncovalent interactions consist of hydrophobic interaction, hydrogen bonding interaction, electrostatic interaction, and π-π stacking interaction, etc. [24,25,26]. During the assembly process, the driving force is a key factor affecting the success of carrier-free nanodrugs. Second, covalent bonds are employed to combine one or multiple drug molecules or drug molecules with functional biomolecules to form prodrugs, which are subsequently self-assembled to construct carrier-free nanodrugs through noncovalent interactions [27]. The prodrugs usually display certain special responsive structures, which can achieve controlled release of nanodrugs at specific sites. For example, pH-responsive imine (Schiff-base) bonds [28], reduction-responsive disulfide bonds [29], easily hydrolyzable ester bonds [16], and enzyme-cleavable specific peptide linkers [30]. Third, carrier-free nanodrugs can be also designed directly by coordination-driven self-assembly of metal ions with drug molecules or functional biomolecules [31,32]. Coordination covalent bonds exhibit stable and dynamic behavior in complex environments, and the strength is between weak noncovalent interactions and strong covalent bonds [33].
Nowadays, many excellent published reviews have summarized the preparation methods, assembled ingredients, physicochemical properties, and external modifications [11,21,23,34,35,36]. The self-assembly mechanisms and strategies involved in their preparation are few systematically discussed. In this review, based on the assembly mechanism and strategy involved during carrier-free nanodrugs preparation, we summarize and discuss the research progress on the construction of carrier-free nanodrugs via noncovalent interactions, the combination of covalent bonds and noncovalent interactions, along with metal ions coordination-driven assembly (Scheme 1). Simultaneously, the relevant applications of various carrier-free nanodrugs for antitumor applications are also presented. Finally, we discuss the future prospects and challenges of carrier-free nanodrugs in clinical applications.

2. Noncovalent Interactions-Driven Carrier-Free Nanodrugs

Noncovalent forces play an important role in constructing carrier-free nanodrugs [37]. The special functional groups of active agents induce intermolecular noncovalent interactions between them or active agents and water molecules [38]. For example, aliphatic groups can cause hydrophobic interactions [39], hydroxyl/carboxyl groups interact via hydrogen bonds [40], aromatic groups can induce π-π stacking interactions [41], and ionic groups show electrostatic interactions [42]. Drug molecules and functional biomolecules can directly form carrier-free nanodrugs through these noncovalent interactions in the assembly medium solution [43,44,45], which effectively alleviates the problems of poor drug solubility, low anticancer efficiency, and serious side effects.

2.1. Hydrogen Bonding Interaction

Hydrogen bonds are an important noncovalent driving force for the self-assembly of carrier-free nanodrugs, which mainly exist between H atoms and electronegative atoms (e.g., O, F, and N) and are a relatively weak force [46]. Typically, a large number of drugs and functional biomolecules (e.g., anticancer drugs, photosensitizers, nucleic acids, etc.) with multiple hydrogen bonding sites, such as amide groups, amino groups, hydroxyl groups, or carboxyl groups can promote hydrogen bond-driven assembled nanostructures [47]. The assembly principle of hydrogen-bonded nanomedicine typically depends on the use of hydrogen bond and hydrophobic interactions working together, in which hydrogen bond-directed recognition occurs shielding the hydrophobic microenvironment in the medium solution [40]. Thus, one or more drugs and functional biomolecules can co-assemble to form carrier-free nanodrugs by hydrogen bonding and hydrophobic interactions during a single ordered process.
Natural honokiol (HK) is a Magnolia bar-derived anticancer drug for clinical trials in advanced non-small cell lung cancer [48]. However, HK shows hydrophobicity, low bioavailability and rapid metabolism, which greatly limit its application. Ji et al. employed HK to self-assemble into carrier-free nanoparticles (SA) via intermolecular hydrogen bonds and hydrophobic interactions (Figure 1A) [49]. During the assemble process, hydrogen bonding interactions drove 1D complex units’ formation, while hydrophobic units promoted further 3D self-assembly, thereby forming SA with hydrophilic outward and hydrophobic inward, respectively. SA exhibited good solubility and stability, excellent tumor-targeting ability and antitumor immunity (Figure 1B), which enhanced the targeted therapy of HK. This work indicates that hydrogen bonding and hydrophobic interactions-driven self-assembly offer a safe and “molecularly economical” strategy to rationally design clinical agents, which fully exploits the assembly concept of green chemistry.
Some natural small molecules (NSMs) usually contain chemical structures with hydrophobic skeletons and active groups, which can offer hydrogen bond sites. Therefore, the driving forces between two NSMs co-assembly mainly involve hydrogen bonding and hydrophobic interactions. Wang’s group has developed a variety of carrier-free nanodrugs using pure NSMs by hydrogen bonding and hydrophobic interactions [50]. The co-assembled NPs demonstrated strong synergistic antitumor effects and reliable biological safety. Dihydroartemisinin (DHA) is a plant-derived natural drug for effective cancer treatment. Unfortunately, the poor solubility and bioavailability of DHA seriously hamper its clinical application. Thus, Li et al. prepared carrier-free nanoparticles (DHA NPs) based on the hydrogen bonding and hydrophobic interactions between DHA molecules [51]. Compared with pure DHA molecules, DHA NPs not only overcame the poor water solubility of DHA, but also revealed favorable drug-loading, stability, acid-responsive drug release, and antitumor efficacy. All these studies demonstrate the importance of hydrogen bonding in the assembly process of carrier-free nanodrugs.

2.2. π-π Stacking Interaction

The π-π stacking interaction is a weak noncovalent interaction in chemistry, which usually coexists with hydrophobic interaction, particularly in cases involving aromatic groups with bonded structures [52,53]. This π-π stacking interaction typically preserves the functional attributes of the compounds and serves as a primary driving force for self-assembling drugs or functional biomolecules with high π-electron densities into carrier-free nanodrugs, showing potentially wide-ranging applications [54,55].
Most photosensitizers have many aromatic π-conjugated structures with C–C double bonds and electron-rich heterocycles [56]. These structures play a pivotal role in facilitating the co-assembly of photosensitizers with active drug molecules via π-π interactions, which is beneficial for enhancing the cellular internalization and photodynamic efficiency of photosensitizers. Photosensitizer Chlorin e6 (Ce6) with a hydrophobic conjugated skeleton can interact with various hydrophobic chemotherapeutic agents [57]. Generally, photosensitizer loading in dense carrier materials possesses low loading capacity and quenching effect due to aggregation [58]. For example, Wang et al. constructed a Ce6-loaded ZIF-8@ssPDA nanosystem, which improved the bioavailability of Ce6 but still had the risk of Ce6 quenching due to the strong sealing of the carrier [59]. The carrier-free nanodrug delivery systems can better avoid this problem. To achieve synergistic effects with low toxicity, Mai and colleagues designed a carrier-free immunotherapeutic booster (C9SN) for the co-delivery of the glutaminase inhibitor compound 968 (C968) and Ce6 to tumor sites. This design harnessed dual synergistic mechanisms involving both π-π stacking and hydrophobic interactions between C968 and Ce6 molecules (Figure 2A) [60]. Ce6 is uniformly distributed inside C9SN to avoid the aggregation quenching effect. C9SN displayed high drug loading capacity enhanced cellular uptake, and improved tumor accumulation. This carrier-free C9SN triggered tumor immunogenicity and collaboratively reshaped the immunosuppressive tumor microenvironment under laser irradiation, which recruited and activated the cytotoxic T-lymphocytes, enhancing the systemic antitumor immunotherapy effect (Figure 2B).
Recently, the drug-drug delivery systems have been designed by employing π-π stacking interaction and hydrophobic interaction between two molecules. Wang et al. reported carrier-free nanoparticles (DiR-DOX NPs) co-assembled by doxorubicin (DOX) and the near-infrared photosensitizer DiR, further modified with a small amount of PEG (DiR-DOX-PEG NPs) [61]. The hydrophobic interaction of the alkyl chains and π-π stacking interaction between the conjugated structures of the two molecules drove the co-assembly process. Based on the combined action of chemotherapy and photothermal therapy, DiR-DOX-PEG NPs effectively eliminated tumors.
Multicomponent carrier-free nanodrugs involving multiple drugs or functional biomolecules can be also assembled by intermolecular π-π stacking and hydrophobic interactions. Qin’s team co-assembled Ce6, celecoxib (cyclooxygenase-2 inhibitor), and 6-thio-2′-deoxyguanosine (6-thio-dG, telomere-targeting drug) to form multicomponent carrier-free photodynamic nanodrugs (CC-6td NPs) through π-π stacking and hydrophobic interactions [62]. The CC-6td NPs could regulate dendritic cells (DCs) for robust immune responses, enhancing immunotherapeutic effects on tumors and inhibiting primary/abdominal tumor growth and postoperative recurrence. Furthermore, Wang et al. combined CDK4/6 protein hydrolysis-targeted chimeras (PROTAC) with Ce6 to form carrier-free nanoparticles via π-π stacking and hydrophobic interactions [63]. The dual-drugs-based nanoparticles improved antitumor efficacy and readjusted the immune-suppressing tumor microenvironment to enhance anti-tumor immunity.

2.3. Electrostatic Interaction

Electrostatic interaction, one noncovalent bond arising from opposite charges, mainly relies on the charge ratio, pH level, concentration, and ionic strength [64]. The strength of electrostatic interaction is notably influenced by the distance between charge centers and charge number, which directly impacts the interaction’s potency [65]. Therefore, drugs or functional biomolecules with charged groups, such as NH3+ and COO can self-assemble to form carrier-free nanodrugs via electrostatic forces.
Creating ternary-component self-delivery nanodrugs that consider dosage, morphology, stability, and multi-synergistic mechanisms poses challenges. A combination of hydrophobic and electrostatic interactions between molecules can enhance drug physicochemical attributes during self-assembly. Zhao and coworkers prepared a carrier-free photodynamic immunostimulant (BVC) by integrating the Ce6, the ASCT2 inhibitor (V9302), as well as the PD1/PDL1 blocker (BMS-1) [66]. Based on electrostatic interactions, the NH2 on V9302 and the pyrrole of Ce6 interacted with the were assembled by electrostatic interaction with the COOH on Ce6 and BMS. Also, hydrophobic interactions between aromatic and porphyrin moieties can facilitate assembly (Figure 3A). The BVC showed good drug delivery efficiency, robust photodynamic therapy (PDT) effect, and excellent ability to reprogram tumor metabolism. Therefore, BVC enhances immune recognition and reduces tumor cell immune escape to eradicate metastatic tumors (Figure 3B).
Similarly, Le and colleagues reported carrier-free NPs (IRE-NPs) using rapamycin (RAPA, chemotherapeutic drug), indocyanine green (ICG, photosensitizer) and protective polyphenols (EGCG) for synergistic chemo-photothermal therapy [67]. The hydrophobic and electrostatic interactions among RAPA, ICG, and EGCG molecules ensured the assembly of IRE-NPs. Compared with RAPA or ICG alone, IRE NPs demonstrated remarkable stability, photothermal effect, tumor accumulation ability, and anticancer efficacy in MCF-7 cells, HepG2 cells, and HeLa cells. The electrostatic and hydrophobic interactions can efficiently facilitate the assembly of carrier-free nanodrugs in water solution. Zhou and coworkers prepared a carrier-free self-delivery photothermal converter (CypCel) comprised of the photothermal agent Cypate (Cyp) and the anti-inflammatory drug Celecoxib (Cel) [68]. The Cyp and Cel with nonpolar groups could aggregate in aqueous solutions via Hydrophobic interactions. Besides, their polar groups mainly contacted water via electrostatic interactions, further promoting the assembly of CypCel. The CypCel could enhance tumor accumulation for superior photothermal therapy (PTT), and inhibit the inflammatory cascade against tumor cells.

2.4. Multiple Noncovalent Interactions

Since there are weak and highly unstable noncovalent interactions when acting alone, the aforementioned carrier-free nanodrugs mainly involve the co-driven assembly of hydrogen bonding, π-π stacking, electrostatic and hydrophobic interactions, respectively. These studies broaden the avenues to explore the interactions among noncovalent bonds. Moreover, the self-assembly process of carrier-free nanodrugs is also driven by triple-noncovalent bonds, which may play a role in advanced biological structures [55].
As an illustration, Li et al. devised a hybrid carrier-free nanoassembly incorporating etoposide (VP-16, a topoisomerase II inhibitor), pyropheophorbide a (PPa, a photosensitizer), and a small amount of DSPE-PEG2k (called VP-16@PPa PEG2K NPs, Figure 4A) [69]. Macromolecular docking simulations showed that synergistic hydrophobic, π-π stacking, and hydrogen bonding interactions, facilitated the precise co-assembly of VP-16 and PPa (Figure 4B). The resulting nanoassembly showed high drug loading efficiency, prolonged blood circulation time, targeted accumulation within tumors, as well as a favorable antitumor effect in 4T1 breast cancer through multimodal DNA damage (Figure 4).
Carrier-free nanodrugs, directly co-assembled by chemotherapy drugs and NIR probes through multiple intermolecular noncovalent interactions, are promising alternatives to cancer hyperthermia. Chen et al. co-assembled hydrophobic zinc phthalocyanine (ZNPC, photosensitizer) and ICG to construct carrier-free nanodrugs (ZNPC-ICG) via triple noncovalent interactions. The ZNPC-ICG were then camouflaged with erythrocyte membrane to obtain ZNPC-ICG@RBC nanoprobes [70]. These carrier-free nanoprobes selectively accumulated in tumor cells and produced both ROS and hyperthermia under laser irradiation, thereby achieving remarkable tumor ablation without any regeneration by synergistic PDT and PTT. In general, the use of multifunctional antitumor agents as endogenous synergists to enhance chemotherapy-PDT is still a challenge in the design of carrier-free nanodrugs. Lan et al. developed GA-Ce6-FA NPs composed of gambogic acid (GA), Ce6, and folic acid (FA) via electrostatic, hydrophobic, and π-π stacking interactions [71]. The carrier-free NPs revealed performance in fluorescence imaging, specific tumor-targeting and chemotherapy-PDT synergistic antitumor efficacy.
Although all of the above carrier-free nanodrugs are self-assembled from non-covalent interactions, structural stability can be maintained during in vivo circulation. When nanoparticles enter the physiological environment, they may interact with large proteins, and their surface will be covered by a layer of protein, forming what is known as the protein “corona” [72]. For many NPs, the corona formation occurs so quickly that provides a steric stabilizing layer against aggregation by particle contact [73]. For example, citrate-stabilized gold NPs aggregate immediately in PBS, but are stable in plasma [74]. In addition, non-covalent forces within the nanoparticles can maintain the structural integrity of the NPs to some extent. This conclusion was proved by the researchers’ nanomedicine stability experiments in 10% fetal bovine serum (FBS) containing plasma proteins, peptides, fats, carbo-hydrates, growth factors, hormones, and inorganic substances [49,60].
To date, with the development of nanotechnology, surface modification of carrier-free nanodrugs has been accepted as a primary strategy to improve stability and reduce the non-specific adsorption of proteins in the blood. For example, the imparting of a hydrophilic polymer coating (e.g., PEG, proteins, DNA, bio-coatings, etc.) and biomimicry camouflage (e.g., red blood cells membrane) on the surface of carrier-free nanodrugs can reduce non-specific binding in vivo [67,75].

3. Covalent Bonds-Driven Carrier-Free Nanodrugs

Prodrugs, referring to inactive conjugates of drugs, possess biologically reversible modifications to achieve ideal physicochemical properties and are easily decomposed in vivo to release active drugs. [76,77]. Prodrugs can usually offer several virtues, such as improved drug solubility, high drug loading ability, prolonged drug circulation, reduced adverse effects, and tumor-specific drug release in response to stimulation [78,79]. However, due to their small molecular nature, prodrugs are still facing rapid degradation, premature activation, renal clearance, and hepatic metabolism in vivo [23]. To address these problems, prodrugs-based carrier-free nanoplatforms, integrating the prodrugs strategy and nanotechnology into one system, have been viewed as promising candidates for clinical translation [80].
Idealized prodrugs can be activated by specific stimuli in the tumor microenvironment (TME), such as acidity, high glutathione (GSH) or reactive oxygen species (ROS), and specific enzymes, for drug release on demand [81,82]. Therefore, rational design of prodrugs combined with nanotechnology for constructing carrier-free nanodrugs can significantly improve drug stability and pharmacokinetics, reduce renal clearance or degradation, extend drug circulation time, and enhance tumor accumulation via EPR effect and therapeutic efficacy [83,84,85].
Typically, amphiphilic prodrugs can be constructed by covalent bonds between hydrophobic and hydrophilic drugs, or drugs and functional biomolecules [86]. Taking advantage of intermolecular noncovalent interactions, these prodrugs are easily self-assembled into carrier-free nanodrugs. Importantly, they can co-assemble with or co-encapsulate additional drugs, biomolecules, and small prodrugs into carrier-free nanodrugs [27]. Based on the presence of many cleavable chemical bonds (e.g., acid-responsive imine /Schiff base bonds [28], reduction-responsive disulfide bonds [29], easily hydrolyzable ester bonds [16], and enzyme-cleavable specifically peptide linkers [30]), these prodrug-based carrier-free nanodrugs can realize controlled drug release in the special TME [87].

3.1. Imine Bond

Compared with the physiological pH 7.4, the tumor’s extracellular pH averages between pH 6.8 and pH 7.0, and even can get as low as 5.7 [88]. The transmembrane pH gradient varies in different subcellular components, such as mitochondria for pH 8.0, endosomes for pH 5.6–6.0, and lysosomes for pH 4.5–5.0 [89,90]. These microenvironmental differences provide prerequisites for regional management and targeting of pH-responsive carrier-free nanodrugs [91].
The imine (Schiff base) bond is a typical pH-responsive dynamic covalent chemical linkage, which is formed by the condensation reaction between aldehyde or ketone and amine [92]. The imine (Schiff base) bond is relatively stable at pH 7.4 but can be rapidly hydrolyzed under acidic conditions [93]. Thus, the imine (Schiff base) bond is a promising linker for designing pH-sensitive carrier-free nanodrugs due to its simple synthesis process and acid-labile properties.
This imine bond can be used to construct carrier-free nano-prodrugs derived from anticancer twin drugs. Wang et al. designed a hydrophobic tryptamine (Try, an analog of serotonin)-cinnamaldehyde (CA, promoting tumoral ROS-mediated apoptosis) twin drug by aldimine condensation, which further formed pH-triggered small molecule carrier-free nano-prodrugs (denoted as Try-CA-NPs (Figure 5) [94]. Try-CA-NPs showed circulation stability, as well as charge reversal, large-to-small size transition, and drug release under intracellular endosomal pH trigger. Furthermore, liposoluble Try-CA-NPs easily crossed the blood-brain barrier (BBB) and targeted glioma cells through Try-mediated cellular uptake, finally completely destroying SH-SY5Y multicellular spheroids.
Cytotoxic drugs (e.g., hydrophobic DOX) are commonly used in clinical cancer therapy at the maximum tolerated dose. Nevertheless, high doses of drugs often induce serious toxicity and drug resistance [95]. To achieve enhanced efficacy, hydrophobic drugs can be conjugated to other drugs or functional biomolecules via bioenvironment-responsive bonds, and then realize drug-controlled release at target sites. Fan et al. reported an amphiphilic drug-drug conjugate (LMWH-DOX) using hydrophilic low molecular weight heparin (LMWH) and DOX via imine bonds [96]. The LMWH-DOX conjugate is then self-assembled into carrier-free LD NDs, which enhances the tumor accumulation and releases the drugs under an acidic stimulus. Therefore, LD NDs exhibited significant inhibitory effects on tumor growth and metastasis with good biocompatibility.
Utilizing the acidic condition of the TME, Wu et al. first obtained fluorinated prodrugs DOX and melittin (MPI, cytolytic peptide) via Schiff base (imine) reaction, which denoted as FDOX (FD) and FMPI (FM), respectively [97]. Next, the FD, FM, and thymocyte selection-associated high mobility group box protein (TOX) were co-assembled into carrier-free FD/FM@siTOX NPs. These NPs showed potent pH-responsive drug release, tumor accumulation, anti-liver cancer immune response and metastasis.

3.2. Disulfide Bond

Tumor cells show higher oxidative and reducing stress states due to intracellularly overproduced ROS and GSH than normal cells, thus it is difficult to release parent drugs into the tumor [98]. Taking advantage of the redox gradient between tumor and normal cells, some specific chemical linkages can be utilized to design prodrugs which provide a prerequisite for constructing redox-responsive carrier-free nanodrugs [99]. Inspired by the chemical properties and functional effect, disulfide bonds are widely used as reduction-sensitive linkers in the design of antitumor prodrugs [85,100].
When the disulfide bond is triggered by GSH, it is easy to convert into two -SH groups through a reduction reaction [101]. Connecting different drugs or functional biomolecules with disulfide bonds can fabricate multifunctional antitumor prodrugs-based carrier-free nanodrugs, achieving tumor-specific drug release. Pei and coworkers prepared a range of dimeric prodrugs by coupling two paclitaxel (PTX) molecules with different chain-length dicarboxylic acids via disulfide bonds. Each of these PTX dimers was able to self-assemble into NPs [102]. Furthermore, Kang et al. employed disulfide bonds to connect the indomethacin (IND, nonsteroidal anti-inflammatory drug) and PTX to obtain the drug conjugate, which self-assembled into carrier-free nanodrugs (IND-S-S-PTX NPs) by van der Waals force-driven self-assembly (Figure 6) [103]. The IND-S-S-PTX NPs realized PTX-controlled in response to high-level GSH in TME. The IC50 value of IND-S-S-PTX NPs was only about 1/3 of that of PTX, but the apoptotic ratio of IND-S-S-PTX NPs was almost five times that of PTX. Besides, the NPs could induce apoptosis towards A549/taxol and reverse multi-drug resistance by down-regulating MRP1 protein expression. However, when PTX was loaded into nanocluster carrier material, its anticancer activity was not significantly improved, which was almost the same as that of free PTX [104]. These studies show that connecting PTX with disulfide bonds and assembling prodrug-based carrier-free nanodrugs may significantly enhance the drug activity and antitumor potential than carrier delivery.
Apart from being widely used in the design of reduction-responsive carrier-free nanodrugs, the disulfide bond can also serve as an oxidation-responsive linker, because it is oxidized to sulfoxide or sulphone under oxidative stimulation [99]. Sun et al. reported three paclitaxel-citronellol (PTX-CIT) prodrugs linked by disulfide bonds-containing carbon chains with different lengths, which can be assembled into uniform NPs [101]. The disulfide bond-based prodrug NPs exhibited unique redox dual-responsiveness when exposed to co-stimuli of oxidation and reduction. The H2O2 oxidized disulfide bonds to hydrophilic sulfoxides or sulfones, facilitating the oxidation-responsive drug release and showing potent antitumor activity. This work provides the possibility for constructing redox dual-responsive carrier-free nanodrugs via disulfide bonds for cancer therapy.
The GSH is much higher in cancer cells than that in normal cells [105], thus carrier-free nanodrugs with disulfide bonds can add additional selective targeting for cancer treatment. Feng’s group developed a multifunctional carrier-free nano-photosensitizer (NPS) using one single amphiphilic glycosylated BODIPY (BSL) molecule containing two disulfide bonds [106]. For BSL NPS, the lactose moieties were hydrophilic targeting agents, while borondipyrromethenes (BODIPYs) were used as hydrophobic dual phototherapy (DPT) photosensitizers. BSL NPS showed superior biocompatibility, tumor-targetability, GSH-responsiveness, and precise release of BODIPY motif in TME, and simultaneously produced ROS and heat, which enhanced antitumor effect through synergistic PDT and PTT.
Additionally, disulfide bond-mediated prodrug nanoplatforms can be responsive to exogenous oxidants (single-linear oxygen, 1O2) induced by photosensitizers [107]. Yang et al. reported a prodrug nanoplatform encapsulating pyropheophorbide a (PPa) based on the disulfide bond bridged PTX and oleic acid (OA) [108]. The obtained prodrug nanosystem could respond to both redox heterogeneity in tumors and exogenous oxidants (singlet oxygen) produced by photosensitizers, then undergo the oxidative self-strengthened process and promoted PTX cascade release, realizing obvious chemo-photodynamic therapy against tumor.

3.3. Ester Bond

Ester bonds are commonly used for self-assemble of prodrugs because they can be decomposed by enzyme catalysis or spontaneous hydrolysis for drug release without specific stimulus [109]. The carboxylic ester linkage is an appealing option since many drugs or functional biomolecules contain carboxylic acid or alcohol functional groups [82]. Yang et al. developed an amphiphilic drug-drug conjugate (CP-DDA) using derivatives of dasatinib (DAS, hydrophobic drug) and cisplatin (CP, hydrophilic anti-tumor drug), which could further self-assemble into stable carrier-free nanoparticles (CP-DDA NPs, Figure 7) [110]. The DAS modified with succinic anhydride was bound to the CP derivative (DH-CP) via ester bonds (Figure 7A). After cellular uptake, CP-DDA NPs precisely released DAS and Pt(II) due to the hydrolysis of ester bonds and redox degradation of GSH in the tumor, thereby displaying potent synergistic antitumor activity and reduced toxicities of free drugs (Figure 7B).
In a different report, Li et al. used a succinic acid (SA) linker to connect the anticancer drug Ptx with the tumor-specific peptide RGD to form a Ptx-SA-RGD conjugate. It can be self-assembled into carrier-free nanofibres (P/T-NFs) with loading with tetrandrine (Tet, potential antitumor effects) [111]. These P/T-NFs improved the apoptosis-inducing effect via the ROS-initiated mitochondrial apoptotic pathway and showed a stronger antitumor effect.
Several ester bond-linked prodrugs carrier-free nanodrugs can exhibit pH-sensitive drug release characteristics. Fan et al. constructed an MTX-MAN conjugate by linking double-acting methotrexate (MTX, chemo-drug and FA receptors targeting ligand) and mannose (MAN, targeting ligand and chemo-drug for tumor therapy) via hydrolyzable ester bonds, which can self-assemble to from carrier-free MTX-MAN NPs [112]. MTX-MAN NPs could release drugs through double stimulus of lysosome acidity and esterase, and doubly recognized by tumor, exhibiting excellent tumor accumulation and synergistic chemotherapy effects.

3.4. Specifically Peptide Linker

Several enzymes are overexpressed during tumorigenesis and tumor metastasis, such as cathepsin B [113], and matrix metalloproteinases [114], which are important targets for DDS and therapeutics. Accordingly, these tumor-associated enzyme dysregulations could be exploited as endogenous triggers for designing enzyme-responsive prodrug carrier-free nanodrugs to achieve cancer diagnosis and treatment [79,88,115].
Peptide-drug conjugates (PDCs), one type of prodrug, are drug molecules or functional biomolecules that covalently attach to specific peptide sequences through enzymatically cleavable peptide linkers [116]. PDCs can increase the solubility of hydrophobic drugs, reduce toxicity, actively target specific receptors of tumor cells, and show enzyme-sensitive drug release and therapeutic properties of specific peptides [117]. Furthermore, PDCs can assemble to form carrier-free nanodrugs spontaneously or under certain stimuli [118]. Therefore, PDCs-based carrier-free nanodrugs can be employed for releasing therapeutic agents into tumors related to specific enzymes or enzymatic activity.
Cathepsin B is commonly overproduced in breast cancer, colorectal cancer, and prostate cancer. The targeted bio-enzyme of cathepsin B can serve as a tumor-specific biomarker, whilst it is also a protease that activates nano-prodrugs in tumors [119,120]. Using the pro-apoptotic peptide drug (SMAC), cathepsin B-cleavable peptide (FRRG), and DOX, Shim’s group developed a SMAC-FRRG-DOX conjugate, which self-assembled into DD-NPs [121]. The DD-NPs could be cleaved to release drugs by cathepsin B, displaying dual pro-apoptotic activity and chemotherapy for overcoming drug resistance. Subsequently, Shim et al. identified the most suitable peptide linker for cathepsin B-responsive DOX prodrug NPs [122]. Five peptide-DOX prodrugs were obtained by chemically coupling different cathepsin B cleavable peptides, Phe-Arg-Arg-Gly (FRRG), Phe-Arg-Arg-Leu (FRRL), Phe-Arg-Arg-Gly (FRRLG), Phe-Leu-Arg-Leu-Gly (FLRRG), and Phe-Leu-Leu-Arg-Arg-Gly (FLRRG), to DOX, all then self-assembled into carrier-free NPs (Figure 8A). Comparative studies revealed that FRRL-DOX nanoparticles showed uniform size, good stability, high cathepsin B-specificity, and efficient tumor cellular uptake. Besides, FRRL-DOX nanoparticles significantly reduced toxicity to normal tissues and induced tumor apoptosis (Figure 8B).
Matrix metalloproteinase-2 (MMP-2) is up-regulated in a wide range of tumors, so it can be used to design PDCs-based carrier-free nanodrugs [114]. Hu et al. prepared self-delivery MA-pepA-Ce6 NPs via bonding metformin (MET, PD-L1 inhibitor) to Ce6 through MMP-2 cleavable peptide (pepA) [123]. These MA-pepA-Ce6 NPs could realize effective drug accumulation under dual TME-specific stimuli sensitivity, demonstrating obvious antitumor immune response through the synergistic effect of PDT and PD-L1 inhibitor.

4. Metal Ions-Driven Carrier-Free Nanodrugs

The construction of carrier-free nanodrugs using noncovalent interactions (e.g., hydrogen bonding forces, hydrophobic and electrostatic interactions) is a facile synthesis process [124]. Nevertheless, the complex in vivo environment may lead to the decomposition of carrier-free nanodrugs according to the unstable nature of the noncovalent interactions, reducing drug efficacy and damaging normal tissues [125]. Although covalent linkages based-carrier-free nanodrugs are used during the systematic circulation, nanodrugs with excessively high stability may suffer from hindering drug release at the target site [33]. Consequently, it is necessary to develop therapeutic carrier-free nanodrugs through a simple, smart and versatile approach.
Recently, designing carrier-free nanodrugs by metal-coordinated self-assembly between metal ions and drug or functional biomolecules has gradually attracted great attention [22,126]. The coordinated covalent bonds exhibit stable and dynamic behavior in complicated environments since their strength is stronger than weak noncovalent interactions but weaker than the strong covalent bonds [127].
Metal ions show specific electronic, optical, radioactive, magnetic, and catalytic properties, and can integrate tumor diagnosis and therapy to achieve tremendous effect [128]. For example, Fe2+ and Cu+ can reduce H2O2 to produce highly toxic ROS for chemodynamic therapy (CDT) [129,130]. Mn2+ and Gd3+ can act as magnetic resonance imaging (MRI) contrast agents owing to their remarkable magnetic properties [131,132]. Therefore, metal ions can coordinate with various anticancer drugs, photosensitizers, or functional biomolecules to form carrier-free nanodrugs with versatile functionalities and synergetic therapy, which will provide more possibilities for effective tumor diagnosis and treatment. In this part, the metal ions-coordinated carrier-free nanodrugs will be discussed based on the extensive multifunction and application of metal ions in tumor diagnosis and treatment.

4.1. Fe2+/3+ Ions Coordination

Iron ions are the most commonly used coordination ions due to their excellent physicochemical and biological properties [133]. As an important active transitional metal element, ions usually participate in diverse intracellular activities, especially regulating the intracellular redox state. Fe2+/3+ ions induced Fenton reaction in CDT have been broadly investigated, which can exert Fenton activity under reducing and acidic TME, catalyzing H2O2 to produce ROS for killing tumor cells [33,134]. Apart from potential antitumor applications, compounds with ions can be also used as MRI contrast agents [135]. Moreover, Fe2+/3+ complexes have been shown to be excellent candidates for PTT [136]. Hence, Fe2+/3+-coordinated carrier-free nanodrugs are promising functional candidates for cancer therapeutics.
Iron ions can coordinate with natural polyphenols with multiple phenolic hydroxyl groups to form 3D metal polyphenol networks (MPN), which possess high specific surface areas and provide many adsorption sites for drug molecules [137]. Shang et al. reported Fe2+-coordinated carrier-free nanodrugs (p-QDF@M, Figure 9A) consisting of DOX, natural quercetin (enhancing pro-apoptotic signal), and RBC membranes for triple-negative breast cancer (TNBC) treatment [138]. The phenolic hydroxyl groups of quercetin could be coordinated with Fe2+ to form MPN. Then, DOX was effectively embedded into a 3D network of MPN via π-π stacking. After tumor cell internalization, the Fe2+ coordinated bonds in p-QDF@M were readily dissociated in lysosomal acidic conditions, thus releasing DOX and quercetin. Meanwhile, the released Fe2+ up-regulated the ROS level through the Fenton reaction, inducing mitochondrial damage and cancer cell apoptosis. Based on the synergistic effect of DOX, quercetin and Fe2+, p-QDF@M significantly enhanced the therapeutic effect against TNBC and inhibited pulmonary metastasis (Figure 9B).
The coordination forces between ions and multiple drugs can maintain stability, avoid premature drug leakage, and enhance therapeutic outcomes. As a typical example, Fan et al. designed carrier-free Trojan-horse metal-organic nanotheranostics with reducible diameters by coordinating-assembly of pemetrexed (PEM), FeIII ions, and antiangiogenesis pseudolaric acid B (PAB) [139]. These nanotheranostics were stable in physiological conditions but transformed into ultra-small-size nanotheranostics under moderately acidic TME, which were self-target internalized into the tumor. Additionally, these nanodiagnostics achieved visualization of self-targeting ability through computed tomography/magnetic resonance imaging, displaying excellent antitumor and antiangiogenic effects.
ROS, generated from the Fe2+-induced Fenton reaction, can also react with other substances, producing more toxic components for killing tumor cells. Chen and coworkers reported a redox-triggered C-centered free radicals nanogenerator, denoted as Fe(III)-ART NPs, via coordination self-assembly of Fe3+ and hydrolyzed artemisinin (ART) [140]. In a high GSH environment, the carrier-free Fe(III)-ART NPs released ART and Fe3+. Subsequently, Fe3+ was further reduced to Fe2+, which catalyzes the endoperoxide of ART to generate CDT. The Fe(III)-ART NPs not only realized self-enhanced magnetic resonance imaging but also presented excellent CDT effects.

4.2. Cu2+ Ions Coordination

Apart from Fe2+/3+ ions, various other transition metal ions can also exhibit analogous Fenton-like reactions to kill tumor cells. Copper, as an essential trace element in humans, has been demonstrated to be another primarily dominant coordinated ion, because its catalytic reactivity is more energetically competitive than that of other transition metal ions [141]. Cu2+ ions can be transformed into Cu+ ions through the interaction with intracellular GSH within the TME. Concurrently, the GSH is converted to GSSH, which protects the generated ROS and induces apoptosis of tumor cells [142]. Meanwhile, Cu+ catalyzes an efficient Fenton-like reaction that spontaneously converts H2O2 to highly toxic ·OH and other reactive oxygen species in weak acidic and neutral conditions [143]. Copper ions with the ability to deplete GSH and produce ·OH have become promising agents for tumor therapy in CDT [144,145].
Recently, Jiang et al. developed a carrier-free platform (Figure 10A) by assembling Cu2+, the chemotherapeutic cisplatin (CDDP), and multidrug resistance protein 1 (MDR1) siRNA [130]. The Cu-siMDR-CDDP system could respond to release CDDP, Cu2+, and siRNA in the acidic TME. This triggered cascade bioreactions for producing H2O2. Subsequently, Cu2+-induced Fenton-like reactions converted H2O2 into HO with GSH depletion, thereby disrupting the redox adaptation mechanisms in drug-resistant tumors. Additionally, the MDR1 siRNA inhibited the expression of P-glycoprotein (P-gp) and CDDP efflux (Figure 10B). Such Cu-siMDR-CDDP showed good therapy in inhibiting tumor growth and overcoming drug resistance.
Cu2+ ions can enhance the photothermal conversion efficiency and ROS productivity of a single photosensitizer. In view of the superior coordination capacity between two sulfonate anions of ICG and Cu2+ ions, Chen et al. designed the carrier-free PhA@NanoICG nanotheranostic through the coordination between Cu2+ and ICG, whilst encapsulating photodynamic-agent pheophorbide A (PhA) [146]. After PhA@NanoICG is internalized inside the tumor, Cu2+ ions could improve the photothermal conversion efficiency of ICG, induce GSH depletion, and destroy the cellular antioxidant defense system, thereby avoiding ROS consumption and improving the phototherapeutic outcomes.

4.3. Mn2+ Ions Coordination

Mn ion is not only a trace metal element closely related to human physiological function but also a common paramagnetic metal ion, thereby regularly serving as an MRI contrast agent for guiding cancer treatment [147]. Mn2+ ions display superior GSH removal ability to break the redox homeostasis of TME, this is conducive for ROS-based PDT and CDT [148]. Also, Mn2+ can catalyze the decomposition of endogenous H2O2 into O2, thus carrier-free nanodrugs formed by Mn2+ coordination can be used for cancer PDT ignoring the limitation of hypoxic TME [149]. Based on these findings, Mn2+ can coordinate with various drugs or functional biomolecules (e.g., anticancer drugs and photosensitizers [150] or photothermal agents [151]) to engineer carrier-free nanodrugs for imaging-guided tumor therapy.
For instance, Geng and coworkers revealed a core-shell nanoplatform by Mn2+-coordinated self-assembly (Figure 11A) [152]. Mn2+ coordinated with DOX and Ce6 to generate the core (chemotherapy) and the shell (PDT), respectively. Upon cloaking with RBC membranes, the “multi-in-one” nanodrugs exhibited improved IC50 value (0.116 μg/mL), prolonged circulation within the blood and enhanced accumulation at tumor sites. Owing to the Mn2+ coordination, these Mn2+-drug nanoparticles showed limited drug leakage in normal tissues but rapidly decomposed to release the drug in the acidic TME, thus demonstrating MRI-guided enhanced chemical-PDT therapy (Figure 11B). In a typical carrier delivery comparison example, Zheng et al. used hierarchical core-shell nanoparticles to encapsulate DOX (ZnO-DOX@ZIF-8), and the drug loading capacity was only 11.2%. In vitro, antitumor results showed that the cell survival rate only decreased to about 50% at a concentration of 3.125 μg/mL (0.4 μg/mL for free DOX) [153]. In comparison, the Mn2+ coordinated carrier-free assembly strategy effectively enhanced the DOX drug activity.
In a recent report, Liu’s group designed dual-photosensitizers co-loading nanodrugs (IMCP) using the MnII/III, ICG, and Ce6 via MnII-coordination co-assembly [131]. The MnII presented in the IMCP exhibited the ability to deplete GSH within tumor cells, thus facilitating ROS-based PDT therapy with good biosafety in the dark. Additionally, IMCP integrated the inherent fluorescence, the photoacoustic imaging capability of photosensitizers, and the MRI contrast potential of MnII/III. This amalgamation conferred remarkable imaging abilities for tumor diagnostics, achieving excellent tumor accumulation, self-enhanced phototherapy, and the guidance of multi-modal imaging for PDT/PTT-mediated tumor eradication.
Furthermore, the coordination between Mn2+ and the porphyrin ring within the TME following MnO2 degradation is also advantageous for PDT. Chu and colleagues introduced an innovative O2 self-sufficient nanodrug delivery platform, defined as MnO2/DVDMS. This platform was reduced by GSH and H2O2 to produce Mn2+ and O2. Subsequently, the Mn2+ could coordinate sinoporphyrin sodium (DVDMS, photosensitizer) to form nanotheranostics (nanoDVD) in situ, which improved the PDT efficacy. The elevated stacking of porphyrins within the nanoDVD enhanced the PTT effect [154]. This tumor environment-triggered NanoDVD enabled multimodal magnetic resonance, fluorescence, and photoacoustic imaging guidance, including for effective cancer treatment.

4.4. Gd3+ Ions Coordination

Magnetic gadolinium (Gd) is a widely used MRI contrast agent approved by the U.S. Food and Drug Administration (FDA), and it readily coordinates with O or N to form chelates [155]. Moreover, as the coordination number decreases, the longitudinal relaxation (r1) of Gd3+ ions will increase, this can improve its imaging ability in TME [156].
In this aspect, Wen et al. successfully developed an ICG-loaded carrier-free MPN theranostic agent through the coordination between curcumin and Gd3+, denoted as ICG@cur-Gd NPs (Figure 12A) [132]. Compared with the curcumin loaded in the spherical polymer nanostructures (2.755% of drug loading capacity) for combinatorial tumor therapy [157], the ICG@cur-Gd NPs assembled by the coordination of curcumin and Gd3+ significantly improve the bioavailability and drug loading of curcumin. The tumor acidic microenvironment induced the decomposition of the theranostic agent to specifically release ICG, curcumin, and Gd3+, thereby greatly enhancing the chemo-/photodynamic therapeutic efficacy with good systemic biocompatibility. Dual magnetic resonance (MR)/fluorescence imaging of ICG@cur-Gd NPs presented precise tumor visualization for guiding synergistic chemo-photodynamic cancer treatment (Figure 12B).
In another study, Fan and coworkers used methotrexate (MTX, anticancer and targeting agent), Gd, and artesunate (ASA, CDT-like agent) to coordinately assemble into carrier-free ASA-MTX-GdIII NPs [158]. Following self-targeting to cancer cells through the FA receptor, this nanoplatform underwent decomposition triggered by lysosomal acidity, releasing GdIII ions, ASA, and MTX. Then, the endogenous FeII ions, overproduced specifically from the tumor, reacted with ASA to generate ROS that acted synergistically with MTX. With MRI guidance and self-targeting, ASA-MTX-GdIII NPs showed coincident and synergetic chemodynamic-chemotherapy, resulting in ROS amplification and the near complete eradication of the tumor.
Moreover, Gd, as a high-Z element, holds promising for sensitizing radiation therapy (RT) [159]. Based on this characteristic, Huang and associates engineered carrier-free nanorods (ZGd-NR) through Gd3+ coordinating with zoledronic acid (Zol) [160]. The ZGd-NR utilized deposited X-rays to generate enough ·OH, which induced robust immunogenic cell death (ICD). This process also involved the consumption of tumor-associated macrophages and suppression of regulatory cytokines. Meanwhile, the released Zol caused apoptotic cell death of tumor-based macrophages by inhibiting the mevalonate pathway and reprograming TME. Therefore, ZGd-NR-sensitized RT not only promoted immune activation but also improved the efficacy of checkpoint blockade immunotherapy.

4.5. Zn2+ Ions Coordination

As another essential element, Zn shows excellent biocompatibility within the human body and holds a vital role in ensuring specific functions in the body [33]. Currently, the carrier-free nanostructures based on Zn coordination primarily involve the coordinated interaction between Zn and N atoms on imidazole and porphyrin rings [161], as well as amino acid/peptide/protein [162]. Hence, the assembly driven by Zn2+ ion coordination is a promising direction in the integration of drugs or functional biomolecules to construct innovative carrier-free tumor theranostic nanosystems.
As a typical example, Zhong’s group developed a zinc-coordinated carrier-free nanodrug (PZB NP) via the coordination among protoporphyrin IX (PpIX), Zn2+, GSH inhibitor L-butyrate sulfoximine (BSO) [161]. The multifunctional PZB NP could reduce existing intracellular GSH before PDT, and further block the production of antioxidant enzymes after PDT, thus amplifying ROS-induced tumor cell damage in a two-pronged manner and enhancing the antitumor PDT efficacy.
During the self-assembly process of carrier-free nano-prodrug, coupling hydrophobic drugs with natural amino acids and utilizing Zn2+ coordination-driven multimolecular interactions is an effective strategy. Xu et al. reported a carrier-free nano-prodrug (AHZ) by the coordination interaction between hydrophobic artesunate-histidine conjugate (AH) and Zn2+, in which histidine possessed strong coordination affinity with Zn2+ [31]. This AHZ nano-prodrug was quickly disassembled in the TME with weak acidity and high GSH levels and showed better anticancer performance than direct administration, which is suitable for cancer targeting.
Zn2+ ions can participate as cofactors in the coordination assembly of carrier-free nanodrugs. Shi’s team prepared amphiphilic DNAzyme-TBD conjugate by grafting AIE photosensitizer (TBD-Br) with a phosphorothioated DNAzyme skeleton, then self-assembled into nanoparticles, accompanied by the surface phosphorothioate groups coordinated with Zn2+ to obtain hybrid carrier-free DNAzyme NPs [163]. Because the surface Zn2+ provided sufficient cofactor supply, and TBD-Br triggered 1O2 production upon light, the obtained DNAzyme NPs effectively induced tumor cell apoptosis. Additionally, they also down-regulated early growth response factor-1 protein (EGR-1), leading to the inhibition of tumor cell proliferation.

4.6. Other Metal Ions Coordination

Aside from the abovementioned metal ions, several other metal ions are gradually used to construct carrier-free nanodrugs for cancer theragnostic applications. For instance, gold (Au)-based carrier-free nanodrugs have attracted increasing interest in the field of targeted delivery, bioimaging, theranostic and biosensing. Wang et al. synthesized the Au(III)-tetra-(4-pyridyl)porphyrin (AuTPyP) through the coordination of Au(III) and porphyrin, the AuTPyP monomers were further self-assembled into carrier-free Au-based porphyrin nanorods (AuPNSs) with heat/acid dual stimulation-responsiveness [164]. The resultant AuPNSs possessed excellent photo- and heat-conversion efficiencies for PTT, facilitating hydrophobic pyridine group protonation and subsequently drug release in acidic TME. With cRGD modification, cRGD-AuPNSs significantly enhanced cellular uptake and synergistic chemo-photothermal therapy efficiency.
To date, many significant achievements have been reported by the coordinated-driven self-assembly of Pt metallacycles and metallacages [165,166,167,168,169]. Platinum ions can form coordination bonds with numerous ligands, enabling the preparation of carrier-free nanodrugs for cancer theranostics [167]. As a typical example, Fu et al. fabricated carrier-free platinum (IV) methylene blue (MB) coordination nanotheranostics, which further encapsulated 10-hydroxycamptothecin (CPT) and represented as CPM [168]. The porous shuttle-shape CPM effectively catalyzed endogenous H2O2 into O2 for PDT enhancement. Furthermore, the red-light irradiation combined lysosomal acidity co-triggered the CPT release from CPM, enhancing localized production of mitochondrial ROS. This synergistic strategy achieved effective tumor ablation through phototherapy. In another work, ** a hopeful strategy for the designation of carrier-free nanodrugs with accurate control of drug proportion is crucial.
(2)
Insufficient targeting efficiency. Most carrier-free nanodrugs enter the tumor mainly through the EPR effect, which is not enough to achieve high tumor selectivity. Recently, multifunctional nanodrugs with surface-modified targeting ligands (e.g., antibody, aptamer, peptide) have shown outstanding specific targeting ability toward cancer cells. Nevertheless, the targeting capacity is achieved through incorporating additional active targeted agents. Accordingly, exploring carrier-free nanodrugs with exceptional tumor-specific self-targeting deserves further studies.
(3)
The lack of a good understanding of pharmacokinetics in vivo. It is essential to have in-depth knowledge of the pharmacokinetic behaviors and destiny of carrier-free nanodrugs in the human body before clinical application. Although numerous reports have shown that carrier-free nanodrugs perform better in prolonging circulation time, enhancing cellar uptake and tumor accumulation, and avoiding rapid clearance, the mechanism of drug metabolism and pharmacological action in vivo should be deeply investigated to promote their clinical application.
(4)
The rationality of drug compatibility has not been well illustrated. Generally, there is antagonism, no interaction, additivity, or synergism between drug molecules. How to select the best optimal drug combination according to the synergistic effect and intermolecular interactions between drug molecules still lacks reliable research [187,188]. Thus, it is urgent to design carrier-free nanodrugs through reasonable drug compatibility for exerting the prime synergistic therapeutic efficiency, reducing potential metabolic risks and toxic-side effects.
(5)
The limited therapeutic applications. To date, although carrier-free nanodrugs can act as excellent nano-candidates, their applications mainly involve cancer therapy. There are few carrier-free nanodrugs designed for the treatment of bacterial infections or other diseases. We look forward to these advanced performances of carrier-free nanodrugs in cancer treatment, including PDT, PTT, CDT, RT, fluorescence imaging, and MRI, can be extended to antibacterial fields, which will undoubtedly bring innovative prospects for antibacterial [189,190] anti-inflammatory [191,192], antiviral [193], wound healing [194] and other fields.
(6)
The lack of deeper research. The exploration and study of carrier-free nanodrugs still stand in their infancy. The stability of long-term preservation needs to be fully considered. The potential toxicity to organs as well as the body’s immune response triggered by carrier-free nanodrugs have received little attention. Furthermore, large-scale industrial production and application in clinical practice are still big challenges to all the carrier-free nanodrugs. It needs to make more great efforts to conduct intensive research on carrier-free nanodrugs for clinical applications.
Overall, carrier-free nanodrugs have shown significant advantages in cancer therapy and present an expansive perspective for drug delivery systems, which offer enormous potential for advantageous clinical antitumor outcomes. Despite this, there is a long way for carrier-free nanodrugs to transfer from laboratory investigation to practical clinical translation, people are firmly convinced that these carrier-free nanodrugs will open up an innovative and wide avenue for the diagnosis and therapy of clinical diseases.

Author Contributions

X.Z.: investigation, conceptualization, and writing—original draft; S.H., L.H., X.C. and X.W. (**n Wang): investigation; Y.-n.F.: conceptualization, supervision, writing—review and editing; H.S. and G.L.: supervision; X.W. (**ng Wang): conceptualization, supervision, writing—review and editing, and funding acquisition. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Natural Science Foundation of China (52273118), and the Key Program of Bei**g Natural Science Foundation (Z200025).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

This study did not report any data.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Not applicable.

References

  1. Siegel, R.L.; Miller, K.D.; Wagle, N.S.; Jemal, A. Cancer statistics, 2023. CA Cancer J. Clin. 2023, 73, 17–48. [Google Scholar] [CrossRef]
  2. Mao, J.J.; Pillai, G.G.; Andrade, C.J.; Ligibel, J.A.; Basu, P.; Cohen, L.; Khan, I.A.; Mustian, K.M.; Puthiyedath, R.; Dhiman, K.S. Integrative oncology: Addressing the global challenges of cancer prevention and treatment. CA Cancer J. Clin. 2022, 72, 144–164. [Google Scholar] [CrossRef]
  3. Rivers, D. Lifestyle interventions for cancer survivors. Nat. Rev. Mater. 2022, 22, 130. [Google Scholar] [CrossRef]
  4. Alamzadeh, Z.; Beik, J.; Mahabadi, V.P.; Ardekani, A.A.; Ghader, A.; Kamrava, S.K.; Dezfuli, A.S.; Ghaznavi, H.; Shakeri-Zadeh, A. Ultrastructural and optical characteristics of cancer cells treated by a nanotechnology based chemo-photothermal therapy method. J. Photochem. Photobiol. B 2019, 192, 19–25. [Google Scholar] [CrossRef]
  5. Mura, S.; Nicolas, J.; Couvreur, P. Stimuli-responsive nanocarriers for drug delivery. Nat. Mater. 2013, 12, 991–1003. [Google Scholar] [CrossRef]
  6. Izci, M.; Maksoudian, C.; Manshian, B.B.; Soenen, S.J. The use of alternative strategies for enhanced nanoparticle delivery to solid tumors. Chem. Rev. 2021, 121, 1746–1803. [Google Scholar] [CrossRef]
  7. Liu, Y.; Hui, Y.; Ran, R.; Yang, G.Z.; Wibowo, D.; Wang, H.F.; Middelberg, A.P.J.; Zhao, C.X. Synergetic Combinations of Dual-Targeting Ligands for Enhanced In Vitro and In Vivo Tumor Targeting. Adv. Healthc. Mater. 2018, 7, e1800106. [Google Scholar] [CrossRef]
  8. Mansuri, S.; Kesharwani, P.; Tekade, R.K.; Jain, N.K. Lyophilized mucoadhesive-dendrimer enclosed matrix tablet for extended oral delivery of albendazole. Eur. J. Pharm. Biopharm. 2016, 102, 202–213. [Google Scholar] [CrossRef]
  9. Liu, C.; Zhao, Z.; Gao, R.; Zhang, X.; Sun, Y.; Wu, J.; Liu, J.; Chen, C. Matrix Metalloproteinase-2-Responsive Surface-Changeable Liposomes Decorated by Multifunctional Peptides to Overcome the Drug Resistance of Triple-Negative Breast Cancer through Enhanced Targeting and Penetrability. ACS Biomater. Sci. Eng. 2022, 8, 2979–2994. [Google Scholar] [CrossRef]
  10. Kundu, P.K.; Samanta, D.; Leizrowice, R.; Margulis, B.; Zhao, H.; Borner, M.; Udayabhaskararao, T.; Manna, D.; Klajn, R. Light-controlled self-assembly of non-photoresponsive nanoparticles. Nat. Chem. 2015, 7, 646–652. [Google Scholar] [CrossRef]
  11. Huang, L.; Zhao, S.; Fang, F.; Xu, T.; Zhang, J. Advances and perspectives in carrier-free nanodrugs for cancer chemo-monotherapy and combination therapy. Biomaterials 2021, 268, 120557. [Google Scholar] [CrossRef]
  12. Peer, D.; Karp, J.M.; Hong, S.; Farokhzad, O.C.; Margalit, R.; Langer, R. Nanocarriers as an emerging platform for cancer therapy. Nat. Nanotechnol. 2007, 2, 751–760. [Google Scholar] [CrossRef]
  13. Mei, H.; Cai, S.; Huang, D.; Gao, H.; Cao, J.; He, B. Carrier-free nanodrugs with efficient drug delivery and release for cancer therapy: From intrinsic physicochemical properties to external modification. Bioact. Mater. 2022, 8, 220–240. [Google Scholar] [CrossRef]
  14. Qin, S.-Y.; Zhang, A.-Q.; Cheng, S.-X.; Rong, L.; Zhang, X.-Z. Drug self-delivery systems for cancer therapy. Biomaterials 2017, 112, 234–247. [Google Scholar] [CrossRef]
  15. Xu, Y.; Huang, Y.; Zhang, X.; Lu, W.; Yu, J.; Liu, S. Carrier-free Janus nano-prodrug based on camptothecin and gemcitabine: Reduction-triggered drug release and synergistic in vitro antiproliferative effect in multiple cancer cells. Int. J. Pharm. 2018, 550, 45–56. [Google Scholar] [CrossRef]
  16. Huang, P.; Wang, D.; Su, Y.; Huang, W.; Zhou, Y.; Cui, D.; Zhu, X.; Yan, D. Combination of Small Molecule Prodrug and Nanodrug Delivery: Amphiphilic Drug-Drug Conjugate for Cancer Therapy. J. Am. Chem. Soc. 2014, 136, 11748–11756. [Google Scholar] [CrossRef]
  17. Li, X.; Yu, S.; Lee, D.; Kim, G.; Lee, B.; Cho, Y.; Zheng, B.-Y.; Ke, M.-R.; Huang, J.-D.; Nam, K.T.; et al. Facile Supramolecular Approach to Nucleic-Acid-Driven Activatable Nanotheranostics that Overcome Drawbacks of Photodynamic Therapy. ACS Nano 2017, 12, 681–688. [Google Scholar] [CrossRef]
  18. Dong, C.; Jiang, Q.; Qian, X.; Wu, W.; Wang, W.; Yu, L.; Chen, Y. A self-assembled carrier-free nanosonosensitizer for photoacoustic imaging-guided synergistic chemo-sonodynamic cancer therapy. Nanoscale 2020, 12, 5587–5600. [Google Scholar] [CrossRef]
  19. Kung Sutherland, M.S.; Walter, R.B.; Jeffrey, S.C.; Burke, P.J.; Yu, C.; Kostner, H.; Stone, I.; Ryan, M.C.; Sussman, D.; Lyon, R.P. SGN-CD33A: A novel CD33-targeting antibody-drug conjugate using a pyrrolobenzodiazepine dimer is active in models of drug-resistant AML. Blood 2013, 122, 1455–1463. [Google Scholar] [CrossRef]
  20. Tan, X.; Li, B.B.; Lu, X.; Jia, F.; Santori, C.; Menon, P.; Li, H.; Zhang, B.; Zhao, J.J.; Zhang, K. Light-Triggered, Self-Immolative Nucleic Acid-Drug Nanostructures. J. Am. Chem. Soc. 2015, 137, 6112–6115. [Google Scholar] [CrossRef]
  21. Huang, L.; Hu, S.; Fu, Y.-N.; Wan, Y.; Li, G.; Wang, X. Multicomponent carrier-free nanodrugs for cancer treatment. J. Mater. Chem. B 2022, 10, 9735–9754. [Google Scholar] [CrossRef]
  22. Zhong, Y.T.; Cen, Y.; Xu, L.; Li, S.Y.; Cheng, H. Recent Progress in Carrier-Free Nanomedicine for Tumor Phototherapy. Adv. Healthc. Mater. 2023, 12, 2202307. [Google Scholar] [CrossRef] [PubMed]
  23. Liu, L.-H.; Zhang, X.-Z. Carrier-free nanomedicines for cancer treatment. Prog. Mater. Sci. 2022, 125, 100919. [Google Scholar] [CrossRef]
  24. Chen, M.; Hou, Y.; Chen, N.; Yang, E.; Sun, Z.; Wu, H.; Xu, X.; Yang, J.; Ma, G.; Huo, X. Co-assemblies based on natural Hemslecin A and β-sitosterol as a new sight for synergistic anti-gastric cancer efficacy in TCM. Colloids Interface Sci. Commun. 2022, 49, 100629. [Google Scholar] [CrossRef]
  25. Liu, J.; Peng, F.; Kang, Y.; Gong, D.; Fan, J.; Zhang, W.; Qiu, F. High-Loading Self-Assembling Peptide Nanoparticles as a Lipid-Free Carrier for Hydrophobic General Anesthetics. Int. J. Nanomed. 2021, 16, 5317–5331. [Google Scholar] [CrossRef]
  26. Zhao, L.P.; Zheng, R.R.; Huang, J.Q.; Chen, X.Y.; Deng, F.A.; Liu, Y.B.; Huang, C.Y.; Yu, X.Y.; Cheng, H.; Li, S.Y. Self-Delivery Photo-Immune Stimulators for Photodynamic Sensitized Tumor Immunotherapy. ACS Nano 2020, 14, 17100–17113. [Google Scholar] [CrossRef]
  27. Li, H.; Zang, W.; Mi, Z.; Li, J.; Wang, L.; **e, D.; Zhao, L.; Wang, D. Tailoring carrier-free nanocombo of small-molecule prodrug for combinational cancer therapy. J. Control. Release 2022, 352, 256–275. [Google Scholar] [CrossRef]
  28. Sun, N.; Zhao, C.; Cheng, R.; Liu, Z.; Li, X.; Lu, A.; Tian, Z.; Yang, Z. Cargo-Free Nanomedicine with pH Sensitivity for Codelivery of DOX Conjugated Prodrug with SN38 To Synergistically Eradicate Breast Cancer Stem Cells. Mol. Pharm. 2018, 15, 3343–3355. [Google Scholar] [CrossRef]
  29. Hou, M.; Xue, P.; Gao, Y.-E.; Ma, X.; Bai, S.; Kang, Y.; Xu, Z. Gemcitabine-camptothecin conjugates: A hybrid prodrug for controlled drug release and synergistic therapeutics. Biomater. Sci. 2017, 5, 1889–1897. [Google Scholar] [CrossRef]
  30. Peng, M.; Qin, S.; Jia, H.; Zheng, D.; Rong, L.; Zhang, X. Self-delivery of a peptide-based prodrug for tumor-targeting therapy. Nano Res. 2015, 9, 663–673. [Google Scholar] [CrossRef]
  31. Xu, P.; Wang, X.; Li, T.; Li, L.; Wu, H.; Tu, J.; Zhang, R.; Zhang, L.; Guo, Z.; Chen, Q. Bioinspired Microenvironment Responsive Nanoprodrug as an Efficient Hydrophobic Drug Self-Delivery System for Cancer Therapy. ACS Appl. Mater. Interfaces 2021, 13, 33926–33936. [Google Scholar] [CrossRef] [PubMed]
  32. Zhang, S.; Wang, Z.; Kong, Z.; Wang, Y.; Zhang, X.; Sun, B.; Zhang, H.; Kan, Q.; He, Z.; Luo, C.; et al. Photosensitizer-driven nanoassemblies of homodimeric prodrug for self-enhancing activation and synergistic chemo-photodynamic therapy. Theranostics 2021, 11, 6019–6032. [Google Scholar] [CrossRef] [PubMed]
  33. Xu, J.; Wang, J.; Ye, J.; Jiao, J.; Liu, Z.; Zhao, C.; Li, B.; Fu, Y. Metal-coordinated supramolecular self-assemblies for cancer theranostics. Adv. Sci. 2021, 8, e2101101. [Google Scholar] [CrossRef]
  34. Zhang, X.; Li, N.; Zhang, S.; Sun, B.; Chen, Q.; He, Z.; Luo, C.; Sun, J. Emerging carrier-free nanosystems based on molecular self-assembly of pure drugs for cancer therapy. Med. Res. Rev. 2020, 40, 1754–1775. [Google Scholar] [CrossRef] [PubMed]
  35. Yang, M.-Y.; Zhao, R.-R.; Fang, Y.-F.; Jiang, J.-L.; Yuan, X.-T.; Shao, J.-W. Carrier-free nanodrug: A novel strategy of cancer diagnosis and synergistic therapy. Int. J. Pharm. 2019, 570, 118663. [Google Scholar] [CrossRef] [PubMed]
  36. Wang, Y.; Yang, P.; Zhao, X.; Gao, D.; Sun, N.; Tian, Z.; Ma, T.; Yang, Z. Multifunctional cargo-free nanomedicine for cancer therapy. Int. J. Mol. Sci. 2018, 19, 2963. [Google Scholar] [CrossRef] [PubMed]
  37. Webber, M.J.; Appel, E.A.; Meijer, E.; Langer, R. Supramolecular biomaterials. Nat. Mater. 2016, 15, 13–26. [Google Scholar] [CrossRef]
  38. Fu, S.; Li, G.; Zang, W.; Zhou, X.; Shi, K.; Zhai, Y. Pure drug nano-assemblies: A facile carrier-free nanoplatform for efficient cancer therapy. Acta Pharm. Sin. B 2022, 12, 92–106. [Google Scholar] [CrossRef]
  39. Chandler, D. Interfaces and the driving force of hydrophobic assembly. Nature 2005, 437, 640–647. [Google Scholar] [CrossRef]
  40. Chen, S.; Wu, Y.; Lortie, F.; Bernard, J.; Binder, W.H.; Zhu, J. Hydrogen-Bonds-Mediated Nanomedicine: Design, Synthesis, and Applications. Macromol. Rapid Commun. 2022, 43, e2200168. [Google Scholar] [CrossRef]
  41. Niu, D.; Jiang, Y.; Ji, L.; Ouyang, G.; Liu, M. Self-assembly through coordination and π-stacking: Controlled switching of circularly polarized luminescence. Angew. Chem. 2019, 58, 5946–5950. [Google Scholar] [CrossRef]
  42. Jiang, Z.; Bhaskaran, A.; Aitken, H.M.; Shackleford, I.C.; Connal, L.A. Using synergistic multiple dynamic bonds to construct polymers with engineered properties. Macromol. Rapid Commun. 2019, 40, e1900038. [Google Scholar] [CrossRef]
  43. Zhang, R.; **ng, R.; Jiao, T.; Ma, K.; Chen, C.; Ma, G.; Yan, X. Carrier-Free, Chemophotodynamic Dual Nanodrugs via Self-Assembly for Synergistic Antitumor Therapy. ACS Appl. Mater. Interfaces 2016, 8, 13262–13269. [Google Scholar] [CrossRef]
  44. Wang, H.; **e, H.; Wang, J.; Wu, J.; Ma, X.; Li, L.; Wei, X.; Ling, Q.; Song, P.; Zhou, L.; et al. Self-Assembling Prodrugs by Precise Programming of Molecular Structures that Contribute Distinct Stability, Pharmacokinetics, and Antitumor Efficacy. Adv. Funct. Mater. 2015, 25, 4956–4965. [Google Scholar] [CrossRef]
  45. Liu, L.; Bao, Y.; Wang, J.; **ao, C.; Chen, L. Construction of carrier-free porphyrin-based drug self-framed delivery system to reverse multidrug resistance through photodynamic-chemotherapy. Dyes Pigments 2020, 177, 107922. [Google Scholar] [CrossRef]
  46. Steiner, T. The hydrogen bond in the solid state. Angew. Chem. Int. Ed. 2002, 41, 48–76. [Google Scholar] [CrossRef]
  47. Zhang, H.; Zhu, J.; Fang, T.; Li, M.; Chen, G.; Chen, Q. Supramolecular biomaterials for enhanced cancer immunotherapy. J. Mater. Chem. B 2022, 10, 7183–7193. [Google Scholar] [CrossRef] [PubMed]
  48. Zhao, M.; Zheng, Y.-H.; Zhao, Q.-Y.; Zheng, W.; Yang, J.-H.; Pei, H.-Y.; Liu, L.; Liu, K.-J.; Xue, L.-L.; Deng, D.-X. Synthesis and evaluation of new compounds bearing 3-(4-aminopiperidin-1-yl) methyl magnolol scaffold as anticancer agents for the treatment of non-small cell lung cancer via targeting autophagy. Eur. J. Med. Chem. 2021, 209, 112922. [Google Scholar] [CrossRef] [PubMed]
  49. Ji, H.; Wang, W.; Li, X.; Han, X.; Zhang, X.; Wang, J.; Liu, C.; Huang, L.; Gao, W. Natural Small Molecules Enabled Efficient Immunotherapy through Supramolecular Self-Assembly in P53-Mutated Colorectal Cancer. ACS Appl. Mater. Interfaces 2022, 14, 2464–2477. [Google Scholar] [CrossRef] [PubMed]
  50. Wang, J.; Qiao, W.; Li, X.; Zhao, H.; Zhang, H.; Dong, A.; Yang, X. A directed co-assembly of herbal small molecules into carrier-free nanodrugs for enhanced synergistic antitumor efficacy. J. Mater. Chem. B 2021, 9, 1040–1048. [Google Scholar] [CrossRef]
  51. Li, Y.; Zhang, W.; Shi, N.; Li, W.; Bi, J.; Feng, X.; Shi, N.; Zhu, W.; **e, Z. Self-assembly and self-delivery of the pure nanodrug dihydroartemisinin for tumor therapy and mechanism analysis. Biomater. Sci. 2023, 11, 2478–2485. [Google Scholar] [CrossRef] [PubMed]
  52. Grimme, S. Do special noncovalent π–π stacking interactions really exist? Angew. Chem. Int. Ed. Engl. 2008, 47, 3430–3434. [Google Scholar] [CrossRef] [PubMed]
  53. Wang, H.; Chen, J.; Xu, C.; Shi, L.; Tayier, M.; Zhou, J.; Zhang, J.; Wu, J.; Ye, Z.; Fang, T. Cancer nanomedicines stabilized by π-π stacking between heterodimeric prodrugs enable exceptionally high drug loading capacity and safer delivery of drug combinations. Theranostics 2017, 7, 3638–3652. [Google Scholar] [CrossRef] [PubMed]
  54. Zhuang, W.-R.; Wang, Y.; Cui, P.-F.; **ng, L.; Lee, J.; Kim, D.; Jiang, H.-L.; Oh, Y.-K. Applications of π-π stacking interactions in the design of drug-delivery systems. J. Control. Release 2019, 294, 311–326. [Google Scholar] [CrossRef]
  55. Fu, S.; Yang, X. Recent advances in natural small molecules as drug delivery systems. J. Mater. Chem. B 2023, 11, 4584–4599. [Google Scholar] [CrossRef]
  56. Ethirajan, M.; Chen, Y.; Joshi, P.; Pandey, R.K. The role of porphyrin chemistry in tumor imaging and photodynamic therapy. Chem. Soc. Rev. 2011, 40, 340–362. [Google Scholar] [CrossRef]
  57. Li, X.; Lee, S.; Yoon, J. Supramolecular photosensitizers rejuvenate photodynamic therapy. Chem. Soc. Rev. 2018, 47, 1174–1188. [Google Scholar] [CrossRef]
  58. Ren, C.; Wang, H.; Mao, D.; Zhang, X.; Fengzhao, Q.; Shi, Y.; Ding, D.; Kong, D.; Wang, L.; Yang, Z. When Molecular Probes Meet Self-Assembly: An Enhanced Quenching Effect. Angew. Chem. Int. Ed. Engl. 2015, 54, 4823–4827. [Google Scholar] [CrossRef]
  59. Wang, M.; Li, F.; Lu, T.; Wu, R.; Yang, S.; Chen, W. Photodynamic and ferroptotic Ce6@ ZIF-8@ ssPDA for head and neck cancer treatment. Mater. Des. 2022, 224, 111403. [Google Scholar] [CrossRef]
  60. Mai, Z.; Zhong, J.; Zhang, J.; Chen, G.; Tang, Y.; Ma, W.; Li, G.; Feng, Z.; Li, F.; Liang, X.-J. Carrier-free immunotherapeutic nano-booster with dual synergistic effects based on glutaminase inhibition combined with photodynamic therapy. ACS Nano 2023, 17, 1583–1596. [Google Scholar] [CrossRef]
  61. Wang, C.; Yu, H.; Yang, X.; Zhang, X.; Wang, Y.; Gu, T.; Zhang, S.; Luo, C. Elaborately engineering of a dual-drug co-assembled nanomedicine for boosting immunogenic cell death and enhancing triple negative breast cancer treatment. Asian J. Pharm. Sci. 2022, 17, 412–424. [Google Scholar] [CrossRef] [PubMed]
  62. Qin, X.; Zhang, M.; Zhao, Z.; Du, Q.; Li, Q.; Jiang, Y.; Xue, F.; Luan, Y. A Carrier-Free Photodynamic Nanodrug to Enable Regulation of Dendritic Cells for Boosting Cancer Immunotherapy. Acta Biomater. 2022, 47, 366–376. [Google Scholar] [CrossRef] [PubMed]
  63. Wang, T.; Zhang, Y.; Chen, K.; Huang, Y.; Liu, Y.; Xu, S.; Wang, W. CDK4/6 nano-PROTAC enhances mitochondria-dependent photodynamic therapy and anti-tumor immunity. Nano Today 2023, 50, 101890. [Google Scholar] [CrossRef]
  64. Richardson, J.J.; Björnmalm, M.; Caruso, F. Multilayer assembly. Technology-driven layer-by-layer assembly of nanofilms. Science 2015, 348, aaa2491. [Google Scholar] [CrossRef] [PubMed]
  65. Murray, J.S.; Politzer, P. The electrostatic potential: An overview. Wiley Interdiscip. Rev. Comput. Mol. Sci. 2011, 1, 153–163. [Google Scholar] [CrossRef]
  66. Zhao, L.; Rao, X.; Zheng, R.; Huang, C.; Kong, R.; Yu, X.; Cheng, H.; Li, S. Targeting glutamine metabolism with photodynamic immunotherapy for metastatic tumor eradication. J. Control. Release 2023, 357, 460–471. [Google Scholar] [CrossRef]
  67. Le, J.-Q.; Yang, F.; Yin, M.-D.; Zhao, R.-R.; Zhang, B.-C.; Li, C.; Lin, J.-F.; Fang, Y.-F.; Lin, Y.-T.; Shao, J.-W. Biomimetic polyphenol-coated nanoparticles by Co-assembly of mTOR inhibitor and photosensitizer for synergistic chemo-photothermal therapy. Colloids Surf. B 2022, 209, 112177. [Google Scholar] [CrossRef]
  68. Zhao, L.-P.; Chen, S.-Y.; Zheng, R.-R.; Rao, X.-N.; Kong, R.-J.; Huang, C.-Y.; Liu, Y.-B.; Tang, Y.; Cheng, H.; Li, S.-Y. Photodynamic Therapy Initiated Ferrotherapy of Self-Delivery Nanomedicine to Amplify Lipid Peroxidation via GPX4 Inactivation. ACS Appl. Mater. Interfaces 2022, 14, 53501–53510. [Google Scholar] [CrossRef]
  69. Li, S.; Yang, F.; Sun, X.; Wang, Y.; Zhang, X.; Zhang, S.; Zhang, H.; Kan, Q.; Sun, J.; He, Z. Precisely engineering a carrier-free hybrid nanoassembly for multimodal DNA damage-augmented photodynamic therapy. Chem. Eng. J. 2021, 426, 130838. [Google Scholar] [CrossRef]
  70. Chen, Y.; Li, Y.; Liu, J.; Zhu, Q.; Ma, J.; Zhu, X. Erythrocyte membrane bioengineered nanoprobes via indocyanine green-directed assembly for single NIR laser-induced efficient photodynamic/photothermal theranostics. J. Control. Release 2021, 335, 345–358. [Google Scholar] [CrossRef]
  71. Lan, J.-S.; Liu, L.; Zeng, R.-F.; Qin, Y.-H.; Hou, J.-W.; **e, S.-S.; Yue, S.; Yang, J.; Ho, R.J.; Ding, Y. Tumor-specific carrier-free nanodrugs with GSH depletion and enhanced ROS generation for endogenous synergistic anti-tumor by a chemotherapy-photodynamic therapy. Chem. Eng. J. 2021, 407, 127212. [Google Scholar] [CrossRef]
  72. Lynch, I.; Dawson, K.A. Protein-nanoparticle interactions. Nano Today 2008, 3, 40–47. [Google Scholar] [CrossRef]
  73. Walkey, C.D.; Chan, W.C. Understanding and controlling the interaction of nanomaterials with proteins in a physiological environment. Chem. Soc. Rev. 2012, 41, 2780–2799. [Google Scholar] [CrossRef] [PubMed]
  74. Dobrovolskaia, M.A.; Patri, A.K.; Zheng, J.; Clogston, J.D.; Ayub, N.; Aggarwal, P.; Neun, B.W.; Hall, J.B.; Mcneil, S.E. Interaction of colloidal gold nanoparticles with human blood: Effects on particle size and analysis of plasma protein binding profiles. Nanomedicine 2009, 5, 106–117. [Google Scholar] [CrossRef] [PubMed]
  75. Li, S.; Jiang, S.; Rahman, M.S.U.; Mei, J.; Wang, X.; Jiang, J.; Chen, Y.; Xu, S.; Liu, Y. Pre-Induced ICD Membrane-Coated Carrier-Free Nanoparticles for the Personalized Lung Cancer Immunotherapy. Small Methods 2023, 7, e2201569. [Google Scholar] [CrossRef]
  76. Souza, C.; Pellosi, D.S.; Tedesco, A.C. Prodrugs for targeted cancer therapy. Expert Rev. Anticancer Ther. 2019, 19, 483–502. [Google Scholar] [CrossRef]
  77. Walther, R.; Rautio, J.; Zelikin, A.N. Prodrugs in medicinal chemistry and enzyme prodrug therapies. Adv. Drug Deliv. Rev. 2017, 118, 65–77. [Google Scholar] [CrossRef]
  78. Harrisson, S.; Nicolas, J.; Maksimenko, A.; Bui, D.T.; Mougin, J.; Couvreur, P. Nanoparticles with in vivo anticancer activity from polymer prodrug amphiphiles prepared by living radical polymerization. Angew. Chem. Int. Ed. Engl. 2013, 52, 1678–1682. [Google Scholar] [CrossRef]
  79. Li, G.; Sun, B.; Li, Y.; Luo, C.; He, Z.; Sun, J. Small-molecule prodrug nanoassemblies: An emerging nanoplatform for anticancer drug delivery. Small 2021, 17, e2101460. [Google Scholar] [CrossRef]
  80. Luo, C.; Sun, J.; Sun, B.; He, Z. Prodrug-based nanoparticulate drug delivery strategies for cancer therapy. Trends Pharmacol. Sci. 2014, 35, 556–566. [Google Scholar] [CrossRef]
  81. Low, L.E.; Wu, J.; Lee, J.; Tey, B.T.; Goh, B.H.; Gao, J.; Li, F.; Ling, D. Tumor-responsive dynamic nanoassemblies for targeted imaging, therapy and microenvironment manipulation. J. Control. Release 2020, 324, 69–103. [Google Scholar] [CrossRef]
  82. Nguyen, A.; Böttger, R.; Li, S.-D. Recent trends in bioresponsive linker technologies of prodrug-based self-assembling nanomaterials. Biomaterials 2021, 275, 120955. [Google Scholar] [CrossRef] [PubMed]
  83. Li, S.; Shan, X.; Wang, Y.; Chen, Q.; Sun, J.; He, Z.; Sun, B.; Luo, C. Dimeric prodrug-based nanomedicines for cancer therapy. J. Control. Release 2020, 326, 510–522. [Google Scholar] [CrossRef] [PubMed]
  84. Xue, P.; Wang, J.; Han, X.; Wang, Y. Hydrophobic drug self-delivery systems as a versatile nanoplatform for cancer therapy: A review. Colloids Surf. B 2019, 180, 202–211. [Google Scholar] [CrossRef] [PubMed]
  85. Hou, M.; Ye, M.; Liu, L.; Xu, M.; Liu, H.; Zhang, H.; Li, Y.; Xu, Z.; Li, B. Azide-locked prodrug co-assembly into nanoparticles with indocyanine green for chemophotothermal therapy. Mol. Pharm. 2022, 19, 3279–3287. [Google Scholar] [CrossRef]
  86. Kyu Shim, M.; Yang, S.; Sun, I.C.; Kim, K. Tumor-activated carrier-free prodrug nanoparticles for targeted cancer Immunotherapy: Preclinical evidence for safe and effective drug delivery. Adv. Drug Deliv. Rev. 2022, 183, 114177. [Google Scholar] [CrossRef] [PubMed]
  87. Torchilin, V.P. Multifunctional, stimuli-sensitive nanoparticulate systems for drug delivery. Nat. Rev. Drug Discov. 2014, 13, 813–827. [Google Scholar] [CrossRef]
  88. Dong, X.; Brahma, R.K.; Fang, C.; Yao, S.Q. Stimulus-responsive self-assembled prodrugs in cancer therapy. Chem. Sci. 2022, 13, 4239–4269. [Google Scholar] [CrossRef]
  89. Mellman, I.; Fuchs, R.; Helenius, A. Acidification of the endocytic and exocytic pathways. Annu. Rev. Biochem. 1986, 55, 663–700. [Google Scholar] [CrossRef]
  90. Kanamala, M.; Wilson, W.R.; Yang, M.; Palmer, B.D.; Wu, Z. Mechanisms and biomaterials in pH-responsive tumour targeted drug delivery: A review. Biomaterials 2016, 85, 152–167. [Google Scholar] [CrossRef]
  91. Zhang, Z.; Chen, X.; Chen, L.; Yu, S.; Cao, Y.; He, C.; Chen, X. Intracellular pH-Sensitive PEG-block-Acetalated-Dextrans as Efficient Drug Delivery Platforms. ACS Appl. Mater. Interfaces 2013, 5, 10760–10766. [Google Scholar] [CrossRef] [PubMed]
  92. Belowich, M.E.; Stoddart, J.F. Dynamic imine chemistry. Chem. Soc. Rev. 2012, 41, 2003–2024. [Google Scholar] [CrossRef] [PubMed]
  93. Hu, X.; Jazani, A.M.; Oh, J.K. Recent advances in development of imine-based acid-degradable polymeric nanoassemblies for intracellular drug delivery. Polymer 2021, 230, 124024. [Google Scholar] [CrossRef]
  94. Wang, Z.; Yao, J.; Guan, Z.; Wu, H.; Cheng, H.; Yan, G.; Tang, R. pH-triggered small molecule nano-prodrugs emulsified from tryptamine-cinnamaldehyde twin drug for targeted synergistic glioma therapy. Colloids Surf. B 2021, 207, 112052. [Google Scholar] [CrossRef]
  95. Raguz, S.; Yague, E. Resistance to chemotherapy: New treatments and novel insights into an old problem. Br. J. Cancer 2008, 99, 387–391. [Google Scholar] [CrossRef] [PubMed]
  96. Fan, B.; Li, Q.; Jiang, Y.; Shen, W.; **ng, Y.; Liang, G.; Wu, Q.; Ban, S.; Zhang, R. Development of carrier-free nanodrugs based on low molecular weight heparin-doxorubicin conjugate assembly with smart pH-triggered drug release characteristics for combinatorial antitumor therapy. New J. Chem. 2022, 46, 820–831. [Google Scholar] [CrossRef]
  97. Wu, P.; Zhang, H.; Sun, M.; Mao, S.; He, Q.; Shi, Y.; Deng, Y.; Dong, Z.; Xu, Q.; Zhao, C. Manipulating Offense and Defense Signaling to Fight Cold Tumors with Carrier-Free Nanoassembly of Fluorinated Prodrug and siRNA. Adv. Mater. 2022, 34, e2203019. [Google Scholar] [CrossRef] [PubMed]
  98. Fang, J.; Seki, T.; Maeda, H. Therapeutic strategies by modulating oxygen stress in cancer and inflammation. Adv. Drug Deliv. Rev. 2009, 61, 290–302. [Google Scholar] [CrossRef] [PubMed]
  99. Wang, Q.; Guan, J.; Wan, J.; Li, Z. Disulfide based prodrugs for cancer therapy. RSC Adv. 2020, 10, 24397–24409. [Google Scholar] [CrossRef]
  100. Zhang, Y.; Wang, F.; Li, M.; Yu, Z.; Qi, R.; Ding, J.; Zhang, Z.; Chen, X. Self-Stabilized Hyaluronate Nanogel for Intracellular Codelivery of Doxorubicin and Cisplatin to Osteosarcoma. Adv. Sci. 2018, 5, 1700821. [Google Scholar] [CrossRef]
  101. Sun, B.; Luo, C.; Yu, H.; Zhang, X.; Chen, Q.; Yang, W.; Wang, M.; Kan, Q.; Zhang, H.; Wang, Y. Disulfide bond-driven oxidation-and reduction-responsive prodrug nanoassemblies for cancer therapy. Nano Lett. 2018, 18, 3643–3650. [Google Scholar] [CrossRef] [PubMed]
  102. Pei, Q.; Hu, X.; Liu, S.; Li, Y.; **e, Z.; **g, X. Paclitaxel dimers assembling nanomedicines for treatment of cervix carcinoma. J. Control. Release 2017, 254, 23–33. [Google Scholar] [CrossRef] [PubMed]
  103. Kang, W.; Ji, Y.; Cheng, Y. Van der Waals force-driven indomethacin-ss-paclitaxel nanodrugs for reversing multidrug resistance and enhancing NSCLC therapy. Int. J. Pharm. 2021, 603, 120691. [Google Scholar] [CrossRef]
  104. Luo, Y.; Zhang, X.; Wang, Y.; Han, F.; Xu, F.; Chen, Y. Mediating physicochemical properties and paclitaxel release of pH-responsive H-type multiblock copolymer self-assembly nanomicelles through epoxidation. J. Mater. Chem. B 2017, 5, 3111–3121. [Google Scholar] [CrossRef] [PubMed]
  105. Gamcsik, M.P.; Kasibhatla, M.S.; Teeter, S.D.; Colvin, O.M. Glutathione levels in human tumors. Biomarkers 2012, 17, 671–691. [Google Scholar] [CrossRef] [PubMed]
  106. Feng, W.; Lv, Y.; Chen, Z.; Wang, F.; Wang, Y.; Pei, Y.; **, W.; Shi, C.; Wang, Y.; Qu, Y. A carrier-free multifunctional nano photosensitizer based on self-assembly of lactose-conjugated BODIPY for enhanced anti-tumor efficacy of dual phototherapy. Chem. Eng. J. 2021, 417, 129178. [Google Scholar] [CrossRef]
  107. Yue, C.; Zhang, C.; Alfranca, G.; Yang, Y.; Jiang, X.; Yang, Y.; Pan, F.; de la Fuente, J.M.; Cui, D. Near-infrared light triggered ROS-activated theranostic platform based on Ce6-CPT-UCNPs for simultaneous fluorescence imaging and chemo-photodynamic combined therapy. Theranostics 2016, 6, 456–469. [Google Scholar] [CrossRef]
  108. Yang, B.; Wei, L.; Wang, Y.; Li, N.; Sun, J. Oxidation-strengthened disulfide-bridged prodrug nanoplatforms with cascade facilitated drug release for synergetic photochemotherapy. Asian J. Pharm. Sci. 2020, 15, 637–645. [Google Scholar] [CrossRef]
  109. Lavis, L.D. Ester bonds in prodrugs. ACS Chem. Biol. 2008, 3, 203–206. [Google Scholar] [CrossRef]
  110. Yang, L.; Xu, J.; **e, Z.; Song, F.; Wang, X.; Tang, R. Carrier-free prodrug nanoparticles based on dasatinib and cisplatin for efficient antitumor in vivo. Asian J. Pharm. Sci. 2021, 16, 762–771. [Google Scholar] [CrossRef]
  111. Li, X.; Yu, N.; Li, J.; Bai, J.; Ding, D.; Tang, Q.; Xu, H. Novel “Carrier-Free” nanofiber codelivery systems with the synergistic antitumor effect of paclitaxel and tetrandrine through the enhancement of mitochondrial apoptosis. ACS Appl. Mater. Interfaces 2020, 12, 10096–10106. [Google Scholar] [CrossRef] [PubMed]
  112. Fan, Z.; Wang, Y.; **ang, S.; Zuo, W.; Huang, D.; Jiang, B.; Sun, H.; Yin, W.; **e, L.; Hou, Z. Dual-self-recognizing, stimulus-responsive and carrier-free methotrexate–mannose conjugate nanoparticles with highly synergistic chemotherapeutic effects. J. Mater. Chem. B 2020, 8, 1922–1934. [Google Scholar] [CrossRef] [PubMed]
  113. Olson, O.C.; Joyce, J.A. Cysteine cathepsin proteases: Regulators of cancer progression and therapeutic response. Nat. Rev. Cancer 2015, 15, 712–729. [Google Scholar] [CrossRef] [PubMed]
  114. Kessenbrock, K.; Plaks, V.; Werb, Z. Matrix metalloproteinases: Regulators of the tumor microenvironment. Cell 2010, 141, 52–67. [Google Scholar] [CrossRef]
  115. Mu, J.; Lin, J.; Huang, P.; Chen, X. Development of endogenous enzyme-responsive nanomaterials for theranostics. Chem. Soc. Rev. 2018, 47, 5554–5573. [Google Scholar] [CrossRef]
  116. Wang, Y.; Cheetham, A.G.; Angacian, G.; Su, H.; **e, L.; Cui, H. Peptide-drug conjugates as effective prodrug strategies for targeted delivery. Adv. Drug Deliv. Rev. 2017, 110, 112–126. [Google Scholar] [CrossRef]
  117. Zhang, C.; Wu, W.; Li, R.Q.; Qiu, W.X.; Zhuang, Z.N.; Cheng, S.X.; Zhang, X.Z. Peptide-based multifunctional nanomaterials for tumor imaging and therapy. Adv. Funct. Mater. 2018, 28, 1804492. [Google Scholar] [CrossRef]
  118. Qi, G.B.; Gao, Y.J.; Wang, L.; Wang, H. Self-assembled peptide-based nanomaterials for biomedical imaging and therapy. Adv. Mater. 2018, 30, e1703444. [Google Scholar] [CrossRef]
  119. Wilhelm, S.; Tavares, A.J.; Dai, Q.; Ohta, S.; Audet, J.; Dvorak, H.F.; Chan, W.C. Analysis of nanoparticle delivery to tumours. Nat. Mater. 2016, 1, 16014. [Google Scholar] [CrossRef]
  120. Akinyemi, A.O.; Pereira, G.; Rocha, F.V. Role of Cathepsin B in Cancer Progression: A Potential Target for Coordination Compounds. Mini-Rev. Med. Chem. 2021, 21, 1612–1624. [Google Scholar] [CrossRef]
  121. Shim, M.K.; Moon, Y.; Yang, S.; Kim, J.; Cho, H.; Lim, S.; Yoon, H.Y.; Seong, J.-K.; Kim, K. Cancer-specific drug-drug nanoparticles of pro-apoptotic and cathepsin B-cleavable peptide-conjugated doxorubicin for drug-resistant cancer therapy. Biomaterials 2020, 261, 120347. [Google Scholar] [CrossRef] [PubMed]
  122. Shim, N.; Jeon, S.I.; Yang, S.; Park, J.Y.; Jo, M.; Kim, J.; Choi, J.; Yun, W.S.; Kim, J.; Lee, Y. Comparative study of cathepsin B-cleavable linkers for the optimal design of cathepsin B-specific doxorubicin prodrug nanoparticles for targeted cancer therapy. Biomaterials 2022, 289, 121806. [Google Scholar] [CrossRef]
  123. Hu, C.; He, X.; Chen, Y.; Yang, X.; Qin, L.; Lei, T.; Zhou, Y.; Gong, T.; Huang, Y.; Gao, H. Metformin mediated PD-L1 downregulation in combination with photodynamic-immunotherapy for treatment of breast cancer. Adv. Funct. Mater. 2021, 31, 2007149. [Google Scholar] [CrossRef]
  124. Zhou, J.; Rao, L.; Yu, G.; Cook, T.R.; Chen, X.; Huang, F. Supramolecular cancer nanotheranostics. Chem. Soc. Rev. 2021, 50, 2839–2891. [Google Scholar] [CrossRef] [PubMed]
  125. Tu, L.; Fan, Z.; Zhu, F.; Zhang, Q.; Zeng, S.; Chen, Z.; Ren, L.; Hou, Z.; Ye, S.; Li, Y. Self-recognizing and stimulus-responsive carrier-free metal-coordinated nanotheranostics for magnetic resonance/photoacoustic/fluorescence imaging-guided synergistic photo-chemotherapy. J. Mater. Chem. B. 2020, 8, 5667–5681. [Google Scholar] [CrossRef]
  126. Liu, B.; Hu, F.; Zhang, J.; Wang, C.; Li, L. A biomimetic coordination nanoplatform for controlled encapsulation and delivery of drug-gene combinations. Angew. Chem. Int. Ed. Engl. 2019, 58, 8804–8808. [Google Scholar] [CrossRef]
  127. He, C.; Liu, D.; Lin, W. Nanomedicine applications of hybrid nanomaterials built from metal–ligand coordination bonds: Nanoscale metal-organic frameworks and nanoscale coordination polymers. Chem. Rev. 2015, 115, 11079–11108. [Google Scholar] [CrossRef]
  128. Liu, S.; Xu, X.; Ye, J.; Wang, J.; Wang, Q.; Liu, Z.; Xu, J.; Fu, Y. Metal-coordinated nanodrugs based on natural products for cancer theranostics. Chem. Eng. J. 2023, 456, 140892. [Google Scholar] [CrossRef]
  129. Huang, S.; Le, H.; Hong, G.; Chen, G.; Zhang, F.; Lu, L.; Zhang, X.; Qiu, Y.; Wang, Z.; Zhang, Q. An all-in-one biomimetic iron-small interfering RNA nanoplatform induces ferroptosis for cancer therapy. Acta Biomater. 2022, 148, 244–257. [Google Scholar] [CrossRef]
  130. Jiang, Y.; Liu, Y.; Wang, M.; Li, Z.; Su, L.; Xu, X.; **ng, C.; Li, J.; Lin, L.; Lu, C. siRNA-Based Carrier-Free System for Synergistic Chemo/Chemodynamic/RNAi Therapy of Drug-Resistant Tumors. ACS Appl. Mater. Interfaces 2021, 14, 361–372. [Google Scholar] [CrossRef]
  131. Liu, J.; Zuo, W.; **, Q.; Liu, C.; Liu, N.; Tian, H.; Zhu, X. Mn (II)-directed dual-photosensitizers co-assemblies for multimodal imaging-guided self-enhanced phototherapy. Mater. Sci. Eng. C. 2021, 129, 112351. [Google Scholar] [CrossRef] [PubMed]
  132. Wen, Y.; Hu, J.; Liu, J.; Li, M. Degradable Carrier-Free Metal-Phenolic Network Theranostic Agent with Targeted Mitochondrial Damage for Efficient Cancer Theranostics. Chem. Mater. 2021, 33, 7089–7099. [Google Scholar] [CrossRef]
  133. Zhang, P.; Hou, Y.; Zeng, J.; Li, Y.; Wang, Z.; Zhu, R.; Ma, T.; Gao, M. Coordinatively unsaturated Fe3+ based activatable probes for enhanced MRI and therapy of tumors. Angew. Chem. Int. Ed. 2019, 58, 11088–11096. [Google Scholar] [CrossRef] [PubMed]
  134. Zhuang, Y.; Han, S.; Fang, Y.; Huang, H.; Wu, J. Multidimensional transitional metal-actuated nanoplatforms for cancer chemodynamic modulation. Coord. Chem. Rev. 2022, 455, 214360. [Google Scholar] [CrossRef]
  135. Ren, Z.; Sun, S.; Sun, R.; Cui, G.; Hong, L.; Rao, B.; Li, A.; Yu, Z.; Kan, Q.; Mao, Z. A metal-polyphenol-coordinated nanomedicine for synergistic cascade cancer chemotherapy and chemodynamic therapy. Adv Mater. 2020, 32, 1906024. [Google Scholar] [CrossRef]
  136. Li, J.; Li, X.; Gong, S.; Zhang, C.; Qian, C.; Qiao, H.; Sun, M. Dual-mode avocado-like all-iron nanoplatform for enhanced T1/T2 MRI-guided cancer theranostic therapy. Nano Lett. 2020, 20, 4842–4849. [Google Scholar] [CrossRef]
  137. Zhang, L.; McClements, D.J.; Wei, Z.; Wang, G.; Liu, X.; Liu, F. Delivery of synergistic polyphenol combinations using biopolymer-based systems: Advances in physicochemical properties, stability and bioavailability. Crit. Rev. Food Sci. Nutr. 2020, 60, 2083–2097. [Google Scholar] [CrossRef]
  138. Shang, L.; Yang, T.; Yang, C.; Li, Z.; Kong, L.; Zhang, Z. Metal ions-mediated self-assembly of nanomedicine for combinational therapy against triple-negative breast cancer. Chem. Eng. J. 2021, 425, 131420. [Google Scholar] [CrossRef]
  139. Fan, Z.; Shi, D.; Zuo, W.; Feng, J.; Ge, D.; Su, G.; Yang, L.; Hou, Z. Trojan-Horse Diameter-Reducible Nanotheranostics for Macroscopic/Microscopic Imaging-Monitored Chemo-Antiangiogenic Therapy. ACS Appl. Mater. Interfaces 2022, 14, 5033–5052. [Google Scholar] [CrossRef]
  140. Chen, J.; Wang, X.; Zhang, Y.; Zhang, S.; Liu, H.; Zhang, J.; Feng, H.; Li, B.; Wu, X.; Gao, Y. A redox-triggered C-centered free radicals nanogenerator for self-enhanced magnetic resonance imaging and chemodynamic therapy. Biomaterials 2021, 266, 120457. [Google Scholar] [CrossRef]
  141. Hao, Y.-N.; Zhang, W.-X.; Gao, Y.-R.; Wei, Y.-N.; Shu, Y.; Wang, J.-H. State-of-the-art advances of copper-based nanostructures in the enhancement of chemodynamic therapy. J. Mater. Chem. B 2021, 9, 250–266. [Google Scholar] [CrossRef]
  142. Pi, W.; Wu, L.; Lu, J.; Lin, X.; Huang, X.; Wang, Z.; Yuan, Z.; Qiu, H.; Zhang, J.; Lei, H. A metal ions-mediated natural small molecules carrier-free injectable hydrogel achieving laser-mediated photo-Fenton-like anticancer therapy by synergy apoptosis/cuproptosis/anti-inflammation. Bioact. Mater. 2023, 29, 98–115. [Google Scholar] [CrossRef]
  143. Lu, X.; Gao, S.; Lin, H.; Yu, L.; Han, Y.; Zhu, P.; Bao, W.; Yao, H.; Chen, Y.; Shi, J. Bioinspired copper single-atom catalysts for tumor parallel catalytic therapy. Adv. Mater. 2020, 32, 2002246. [Google Scholar] [CrossRef]
  144. Koo, S.; Park, O.K.; Kim, J.; Han, S.I.; Yoo, T.Y.; Lee, N.; Kim, Y.G.; Kim, H.; Lim, C.; Bae, J.-S. Enhanced chemodynamic therapy by Cu-Fe peroxide nanoparticles: Tumor microenvironment-mediated synergistic Fenton reaction. ACS Nano 2022, 16, 2535–2545. [Google Scholar] [CrossRef] [PubMed]
  145. Feng, G.N.; Huang, X.T.; Jiang, X.L.; Deng, T.W.; Li, Q.X.; Li, J.X.; Wu, Q.N.; Li, S.P.; Sun, X.Q.; Huang, Y.G.; et al. The Antibacterial Effects of Supermolecular Nano-Carriers by Combination of Silver and Photodynamic Therapy. Front. Chem. 2021, 9, 666408. [Google Scholar] [CrossRef] [PubMed]
  146. Chen, L.; Zuo, W.; **ao, Z.; **, Q.; Liu, J.; Wu, L.; Liu, N.; Zhu, X. A carrier-free metal-coordinated dual-photosensitizers nanotheranostic with glutathione-depletion for fluorescence/photoacoustic imaging-guided tumor phototherapy. J. Colloid Interface Sci. 2021, 600, 243–255. [Google Scholar] [CrossRef] [PubMed]
  147. Guan, G.; Zhang, C.; Liu, H.; Wang, Y.; Dong, Z.; Lu, C.; Nan, B.; Yue, R.; Yin, X.; Zhang, X.B. Ternary Alloy PtWMn as a Mn Nanoreservoir for High-Field MRI Monitoring and Highly Selective Ferroptosis Therapy. Angew. Chem. Int. Ed. Engl. 2022, 134, e202117229. [Google Scholar] [CrossRef]
  148. Xu, K.F.; Jia, H.R.; Zhu, Y.X.; Liu, X.; Gao, G.; Li, Y.H.; Wu, F.G. Cholesterol-Modified Dendrimers for Constructing a Tumor Microenvironment-Responsive Drug Delivery System. ACS Biomater. Sci. Eng. 2019, 5, 6072–6081. [Google Scholar] [CrossRef] [PubMed]
  149. Zhang, J.; Xu, M.; Mu, Y.; Li, J.; Foda, M.F.; Zhang, W.; Han, K.; Han, H. Reasonably retard O2 consumption through a photoactivity conversion nanocomposite for oxygenated photodynamic therapy. Biomaterials 2019, 218, 119312. [Google Scholar] [CrossRef]
  150. Zhang, H.; Liu, K.; Li, S.; **n, X.; Yuan, S.; Ma, G.; Yan, X. Self-Assembled Minimalist Multifunctional Theranostic Nanoplatform for Magnetic Resonance Imaging-Guided Tumor Photodynamic Therapy. ACS Nano 2018, 12, 8266–8276. [Google Scholar] [CrossRef]
  151. **ng, R.; Zou, Q.; Yuan, C.; Zhao, L.; Chang, R.; Yan, X. Self-Assembling Endogenous Biliverdin as a Versatile Near-Infrared Photothermal Nanoagent for Cancer Theranostics. Adv. Mater. 2019, 31, e1900822. [Google Scholar] [CrossRef] [PubMed]
  152. Geng, Z.; Chen, F.; Wang, X.; Wang, L.; Pang, Y.; Liu, J. Combining anti-PD-1 antibodies with Mn2+-drug coordinated multifunctional nanoparticles for enhanced cancer therapy. Biomaterials. 2021, 275, 120897. [Google Scholar] [CrossRef] [PubMed]
  153. Zheng, C.; Wang, Y.; Phua, S.Z.F.; Lim, W.Q.; Zhao, Y. ZnO-DOX@ZIF-8 Core-Shell Nanoparticles for pH-Responsive Drug Delivery. ACS. Biomater. Sci. Eng. 2017, 3, 2223–2229. [Google Scholar] [CrossRef]
  154. Chu, C.; Lin, H.; Liu, H.; Wang, X.; Wang, J.; Zhang, P.; Gao, H.; Huang, C.; Zeng, Y.; Tan, Y. Tumor microenvironment-triggered supramolecular system as an in situ nanotheranostic generator for cancer phototherapy. Adv. Mater. 2017, 29, 1605928. [Google Scholar] [CrossRef]
  155. Zeng, Y.; Li, H.; Li, Z.; Luo, Q.; Zhu, H.; Gu, Z.; Zhang, H.; Gong, Q.; Luo, K. Engineered gadolinium-based nanomaterials as cancer imaging agents. Appl. Mater. Today 2020, 20, 100686. [Google Scholar] [CrossRef]
  156. Fan, J.X.; Zheng, D.W.; Mei, W.W.; Chen, S.; Chen, S.Y.; Cheng, S.X.; Zhang, X.Z. A metal-polyphenol network coated nanotheranostic system for metastatic tumor treatments. Small 2017, 13, 1702714. [Google Scholar] [CrossRef] [PubMed]
  157. Stigliano, C.; Key, J.; Ramirez, M.; Aryal, S.; Decuzzi, P. Radiolabeled Polymeric Nanoconstructs Loaded with Docetaxel and Curcumin for Cancer Combinatorial Therapy and Nuclear Imaging. Adv. Funct. Mater. 2015, 25, 3371–3379. [Google Scholar] [CrossRef]
  158. Fan, Z.; Jiang, B.; Zhu, Q.; **ang, S.; Tu, L.; Yang, Y.; Zhao, Q.; Huang, D.; Han, J.; Su, G. Tumor-specific endogenous FeII-activated, MRI-guided self-targeting gadolinium-coordinated theranostic nanoplatforms for amplification of ROS and enhanced chemodynamic chemotherapy. ACS Appl. Mater. Interfaces 2020, 12, 14884–14904. [Google Scholar] [CrossRef]
  159. Detappe, A.; Kunjachan, S.; Rottmann, J.; Robar, J.; Tsiamas, P.; Korideck, H.; Tillement, O.; Berbeco, R. AGuIX nanoparticles as a promising platform for image-guided radiation therapy. Cancer Nanotechnol. 2015, 6, 4. [Google Scholar] [CrossRef]
  160. Huang, Z.; Yao, D.; Ye, Q.; Jiang, H.; Gu, R.; Ji, C.; Wu, J.; Hu, Y.; Yuan, A. Zoledronic Acid-Gadolinium Coordination Polymer Nanorods for Improved Tumor Radioimmunotherapy by Synergetically Inducing Immunogenic Cell Death and Reprogramming the Immunosuppressive Microenvironment. ACS Nano 2021, 15, 8450–8465. [Google Scholar] [CrossRef]
  161. Zhong, H.; Huang, P.Y.; Yan, P.; Chen, P.L.; Shi, Q.Y.; Zhao, Z.A.; Chen, J.X.; Shu, X.; Wang, P.; Yang, B.; et al. Versatile Nanodrugs Containing Glutathione and Heme Oxygenase 1 Inhibitors Enable Suppression of Antioxidant Defense System in a Two-Pronged Manner for Enhanced Photodynamic Therapy. Adv. Healthc. Mater. 2021, 10, e2100770. [Google Scholar] [CrossRef] [PubMed]
  162. Fan, Z.; Sun, L.; Huang, Y.; Wang, Y.; Zhang, M. Bioinspired fluorescent dipeptide nanoparticles for targeted cancer cell imaging and real-time monitoring of drug release. Nat. Nanotechnol. 2016, 11, 388–394. [Google Scholar] [CrossRef] [PubMed]
  163. Shi, L.; Wu, W.; Duan, Y.; Xu, L.; Li, S.; Gao, X.; Liu, B. Carrier-Free Hybrid DNA Nanoparticles for Light-Induced Self-Delivery of Functional Nucleic Acid Enzymes. ACS Nano 2021, 15, 1841–1849. [Google Scholar] [CrossRef] [PubMed]
  164. Fan, Y.; Cui, Y.; Hao, W.; Chen, M.; Liu, Q.; Wang, Y.; Yang, M.; Li, Z.; Gong, W.; Song, S.; et al. Carrier-free highly drug-loaded biomimetic nanosuspensions encapsulated by cancer cell membrane based on homology and active targeting for the treatment of glioma. Bioact. Mater. 2021, 6, 4402–4414. [Google Scholar] [CrossRef] [PubMed]
  165. Yu, G.; Yu, S.; Saha, M.L.; Zhou, J.; Chen, X. A discrete organoplatinum(II) metallacage as a multimodality theranostic platform for cancer photochemotherapy. Nat. Commun. 2018, 9, 4335. [Google Scholar] [CrossRef]
  166. Zhu, H.; Li, Q.; Shi, B.; Ge, F.; Liu, Y.; Mao, Z.; Zhu, H.; Wang, S.; Yu, G.; Huang, F.; et al. Dual-Emissive Platinum(II) Metallacage with a Sensitive Oxygen Response for Imaging of Hypoxia and Imaging-Guided Chemotherapy. Angew. Chem. Int. Ed. Engl. 2020, 59, 20208–20214. [Google Scholar] [CrossRef]
  167. Zhou, J.; Zhang, Y.; Yu, G.; Crawley, M.R.; Fulong, C.R.P.; Friedman, A.E.; Sengupta, S.; Sun, J.; Li, Q.; Huang, F.; et al. Highly Emissive Self-Assembled BODIPY-Platinum Supramolecular Triangles. J. Am. Chem. Soc. 2018, 140, 7730–7736. [Google Scholar] [CrossRef]
  168. Fu, X.; Yin, W.; Shi, D.; Yang, Y.; Zhang, D. Shuttle-Shape Carrier-Free Platinum-Coordinated Nanoreactors with O2 Self-Supply and ROS Augment for Enhanced Phototherapy of Hypoxic Tumor. ACS Appl. Mater. Interfaces 2021, 13, 32690–32702. [Google Scholar] [CrossRef]
  169. **ng, L.; Yang, C.X.; Zhao, D.; Shen, L.J.; Zhou, T.J.; Bi, Y.Y.; Huang, Z.J.; Wei, Q.; Li, L.; Li, F. A carrier-free anti-inflammatory platinum (II) self-delivered nanoprodrug for enhanced breast cancer therapy. J. Control. Release 2021, 331, 460–471. [Google Scholar] [CrossRef]
  170. Orrenius, S.; Zhivotovsky, B.; Nicotera, P. Regulation of cell death: The calcium-apoptosis link. Nat. Rev. Mol. Cell Biol. 2003, 4, 552–565. [Google Scholar] [CrossRef]
  171. Yoon, M.J.; Kim, E.H.; Kwon, T.K.; Park, S.A.; Choi, K.S. Simultaneous mitochondrial Ca2+ overload and proteasomal inhibition are responsible for the induction of paraptosis in malignant breast cancer cells. Cancer Lett. 2012, 324, 197–209. [Google Scholar] [CrossRef] [PubMed]
  172. Kim, S.; Palanikumar, L.; Choi, H.; Jeena, M.T.; Kim, C.; Ryu, J.-H. Intra-mitochondrial biomineralization for inducing apoptosis of cancer cells. Chem. Sci. 2018, 9, 2474–2479. [Google Scholar] [CrossRef] [PubMed]
  173. Chenguang, L.; Lingxiao, G.; Yong, W.; Jianting, Z.; Caiyun, F. Delivering metal ions by nanomaterials: Turning metal ions into drug-like cancer theranostic agents. Coord. Chem. Rev. 2023, 494, 215332. [Google Scholar]
  174. Qin, M.; Li, M.; Song, G.; Yang, C.; Wu, P.; Dai, W.; Zhang, H.; Wang, X.; Wang, Y.; Zhou, D. Boosting innate and adaptive antitumor immunity via a biocompatible and carrier-free nanovaccine engineered by the bisphosphonates-metal coordination. Nano Today 2021, 37, 101097. [Google Scholar] [CrossRef]
  175. Huang, X.; Qiu, M.; Wang, T.; Li, B.; Zhang, S.; Zhang, T.; Liu, P.; Wang, Q.; Qian, Z.R.; Zhu, C. Carrier-free multifunctional nanomedicine for intraperitoneal disseminated ovarian cancer therapy. J. Nanobiotechnol. 2022, 20, 93. [Google Scholar] [CrossRef] [PubMed]
  176. Möschwitzer, J. Drug nanocrystals in the commercial pharmaceutical development process. Int. J. Pharm. 2013, 453, 142–156. [Google Scholar] [CrossRef]
  177. Hollis, C.; Weiss, H.; Leggas, M.; Evers, B.; Gemeinhart, R.A.; Li, T. Biodistribution and bioimaging studies of hybrid paclitaxel nanocrystals: Lessons learned of the EPR effect and image-guided drug delivery. J. Control. Release 2013, 172, 12–21. [Google Scholar] [CrossRef]
  178. Shete, G.; Pawar, Y.; Thanki, K.; Jain, S.; Bansal, A. Oral bioavailability and pharmacodynamic activity of hesperetin nanocrystals generated using a novel bottom-up technology. Mol. Pharm. 2015, 12, 1158–1170. [Google Scholar] [CrossRef]
  179. Shegokar, R.; Müller, R. Nanocrystals: Industrially feasible multifunctional formulation technology for poorly soluble actives. Int. J. Pharm. 2010, 399, 129–139. [Google Scholar] [CrossRef]
  180. Wang, D.; Wang, Y.; Zhao, G.; Zhuang, J.; Wu, W. Improving systemic circulation of paclitaxel nanocrystals by surface hybridization of DSPE-PEG2000. Colloids Surf. B Biointerfaces 2019, 182, 110337. [Google Scholar] [CrossRef]
  181. Parmar, P.; Wadhawan, J.; Bansal, A. Pharmaceutical nanocrystals: A promising approach for improved topical drug delivery. Drug Discov. Today 2021, 26, 2329–2349. [Google Scholar] [CrossRef] [PubMed]
  182. Miao, X.; Yang, W.; Feng, T.; Lin, J.; Huang, P. Drug nanocrystals for cancer therapy. Wiley Interdiscip. Rev. Nanomed. Nanobiotechnol. 2018, 10, e1499. [Google Scholar] [CrossRef] [PubMed]
  183. **ang, H.; Xu, S.; Li, J.; Li, Y.; Xue, X.; Liu, Y.; Li, J.; Miao, X. Functional drug nanocrystals for cancer-target delivery. J. Drug Deliv. Sci. Technol. 2022, 76, 103807. [Google Scholar] [CrossRef]
  184. Yang, X.; Liu, Y.; Zhao, Y.; Han, M.; Guo, Y.; Kuang, H.; Wang, X. A stabilizer-free and organic solvent-free method to prepare 10-hydroxycamptothecin nanocrystals: In vitro and in vivo evaluation. Int. J. Nanomed. 2016, 11, 2979–2994. [Google Scholar] [CrossRef]
  185. Zhang, C.; Long, L.; **ong, Y.; Wang, C.; Peng, C.; Yuan, Y.; Liu, Z.; Lin, Y.; Jia, Y.; Zhou, X. Facile engineering of indomethacin-induced paclitaxel nanocrystal aggregates as carrier-free nanomedicine with improved synergetic antitumor activity. ACS Appl. Mater. Interfaces 2019, 11, 9872–9883. [Google Scholar] [CrossRef] [PubMed]
  186. Su, Y.; Gao, J.; Dong, X.; Wheeler, K.; Wang, Z. Neutrophil-Mediated Delivery of Nanocrystal Drugs via Photoinduced Inflammation Enhances Cancer Therapy. ACS Nano 2023, 17, 15542–15555. [Google Scholar] [CrossRef]
  187. Lázár, V.; Snitser, O.; Barkan, D.; Kishony, R. Antibiotic combinations reduce Staphylococcus aureus clearance. Nature 2022, 610, 540–546. [Google Scholar] [CrossRef]
  188. Brochado, A.R.; Telzerow, A.; Bobonis, J.; Banzhaf, M.; Mateus, A.; Selkrig, J.; Huth, E.; Bassler, S.; Zamarreño Beas, J.; Zietek, M.; et al. Species-specific activity of antibacterial drug combinations. Nature 2018, 559, 259–263. [Google Scholar] [CrossRef]
  189. Li, T.; Wang, P.; Guo, W.; Huang, X.; Tian, X.; Wu, G.; Xu, B.; Li, F.; Yan, C.; Liang, X.-J.; et al. Natural Berberine-Based Chinese Herb Medicine Assembled Nanostructures with Modified Antibacterial Application. ACS Nano 2019, 13, 6770–6781. [Google Scholar] [CrossRef]
  190. Feng, W.; Chittò, M.; Moriarty, T.F.; Li, G.; Wang, X. Targeted Drug Delivery Systems for Eliminating Intracellular Bacteria. Macromol. Biosci. 2022, 23, e2200311. [Google Scholar] [CrossRef]
  191. Wang, H.; Lin, F.; Wu, Y.; Guo, W.; Chen, X.; **ao, C.; Chen, M. Carrier-Free Nanodrug Based on Co-Assembly of Methylprednisolone Dimer and Rutin for Combined Treatment of Spinal Cord Injury. ACS Nano 2023, 17, 12176–12187. [Google Scholar] [CrossRef] [PubMed]
  192. Tang, L.; Di, Z.; Zhang, J.; Yin, F.; Li, L.; Zheng, L. Coordination-driven self-assembly of metallo-nanodrugs for local inflammation alleviation. Nano Res. 2023. [Google Scholar] [CrossRef]
  193. Kim, H.; Zhang, W.; Hwang, J.; An, E.-K.; Choi, Y.K.; Moon, E.; Loznik, M.; Huh, Y.H.; Herrmann, A.; Kwak, M.; et al. Carrier-free micellar CpG interacting with cell membrane for enhanced immunological treatment of HIV-1. Biomaterials 2021, 277, 121081. [Google Scholar] [CrossRef] [PubMed]
  194. Chen, M.; He, X.; Sun, Z.; Huo, X.; Hou, Y.; Xu, X.; Wu, H.; Shi, L.; Ma, G. Natural carrier-free self-assembled diterpene nanoparticles with its efficient anti-inflammation through the inhibition of NF-κB pathway for accelerated wound healing. Biomed. Pharmacother. 2023, 165, 115041. [Google Scholar] [CrossRef] [PubMed]
Scheme 1. Schematic representation of the carrier-free nanodrugs based on various assembly mechanisms and strategies for cancer treatment.
Scheme 1. Schematic representation of the carrier-free nanodrugs based on various assembly mechanisms and strategies for cancer treatment.
Molecules 28 07065 sch001
Figure 1. (A) Preparation of HK-NPS by intermolecular hydrogen bonding interactions. (B) HK-NPS for efficient immunotherapy of colorectal cancer with P53 mutations. Reprinted with permission from ref. [49]. Copyright 2022, American Chemical Society.
Figure 1. (A) Preparation of HK-NPS by intermolecular hydrogen bonding interactions. (B) HK-NPS for efficient immunotherapy of colorectal cancer with P53 mutations. Reprinted with permission from ref. [49]. Copyright 2022, American Chemical Society.
Molecules 28 07065 g001
Figure 2. (A) Preparation of C9SN by π-π stacking and hydrophobic interactions. (B) C9SN and laser irradiation synergistically mediate anti-tumor immunotherapy. Reprinted with permission from ref. [60]. Copyright 2023, American Chemical Society.
Figure 2. (A) Preparation of C9SN by π-π stacking and hydrophobic interactions. (B) C9SN and laser irradiation synergistically mediate anti-tumor immunotherapy. Reprinted with permission from ref. [60]. Copyright 2023, American Chemical Society.
Molecules 28 07065 g002
Figure 3. The preparation and proposed mechanism of BVC for photodynamic activated cascade immunotherapy through tumor metabolism reprogramming. (A) BVC preparation by self-assemble BMS, V9302 and Ce6. (B) BVC could passively accumulate at tumor site to inhibit ASCT2 and down-regulate glutamine and GSH, leading to the metabolism reprogramming of tumor cells to increase the expressions of Fas and PDL1. The elevated production of 1O2 could enhance PDT efficacy to eliminate primary tumor cells and induce ICD to release tumor associated antigens of CRT and HMGB1. Reprinted with permission from ref. [66]. Copyright 2023, Elsevier.
Figure 3. The preparation and proposed mechanism of BVC for photodynamic activated cascade immunotherapy through tumor metabolism reprogramming. (A) BVC preparation by self-assemble BMS, V9302 and Ce6. (B) BVC could passively accumulate at tumor site to inhibit ASCT2 and down-regulate glutamine and GSH, leading to the metabolism reprogramming of tumor cells to increase the expressions of Fas and PDL1. The elevated production of 1O2 could enhance PDT efficacy to eliminate primary tumor cells and induce ICD to release tumor associated antigens of CRT and HMGB1. Reprinted with permission from ref. [66]. Copyright 2023, Elsevier.
Molecules 28 07065 g003
Figure 4. Schematic representation of VP-16@PPa co-assembly and DNA damage-mediated synergetic chemo-photodynamic therapy. (A) The VP-16@PPa PEG2K NPs was precisely delivered into tumor cells, multimodal DNA damage-mediated synergetic chemo-photodynamic therapy was realized under laser irradiation; (B) Molecular docking simulation of VP-16 and PPa. Reprinted with permission from ref. [69]. Copyright 2021, Elsevier.
Figure 4. Schematic representation of VP-16@PPa co-assembly and DNA damage-mediated synergetic chemo-photodynamic therapy. (A) The VP-16@PPa PEG2K NPs was precisely delivered into tumor cells, multimodal DNA damage-mediated synergetic chemo-photodynamic therapy was realized under laser irradiation; (B) Molecular docking simulation of VP-16 and PPa. Reprinted with permission from ref. [69]. Copyright 2021, Elsevier.
Molecules 28 07065 g004
Figure 5. Schematic illustration of the self-assembly, in vivo transport and combinational antitumor therapy of the LD NDs. Reprinted with permission from ref. [94]. Copyright 2021, Elsevier.
Figure 5. Schematic illustration of the self-assembly, in vivo transport and combinational antitumor therapy of the LD NDs. Reprinted with permission from ref. [94]. Copyright 2021, Elsevier.
Molecules 28 07065 g005
Figure 6. Schematic illustration of indomethacin-ss-paclitaxel nanodrugs for reversing multidrug resistance and enhancing NSCLC therapy. Reprinted with permission from ref. [103]. Copyright 2021, Elsevier.
Figure 6. Schematic illustration of indomethacin-ss-paclitaxel nanodrugs for reversing multidrug resistance and enhancing NSCLC therapy. Reprinted with permission from ref. [103]. Copyright 2021, Elsevier.
Molecules 28 07065 g006
Figure 7. (A) The synthetic route of CP-DDA NPs; (B) Drug delivery mechanism of nanoparticles and drug release mechanism in tumor cells. Reprinted with permission from ref. [110]. Copyright 2021, Elsevier.
Figure 7. (A) The synthetic route of CP-DDA NPs; (B) Drug delivery mechanism of nanoparticles and drug release mechanism in tumor cells. Reprinted with permission from ref. [110]. Copyright 2021, Elsevier.
Molecules 28 07065 g007
Figure 8. (A) Schematic synthetic procedure of VES-Ir prodrug and construction of self-assembled VES-Ir NPs, (B) co-delivery and low pH and esterase release and synergistic antitumor effect after endo-lysosomal internalization into tumor cell. Reprinted with permission from ref. [122]. Copyright 2022, Elsevier.
Figure 8. (A) Schematic synthetic procedure of VES-Ir prodrug and construction of self-assembled VES-Ir NPs, (B) co-delivery and low pH and esterase release and synergistic antitumor effect after endo-lysosomal internalization into tumor cell. Reprinted with permission from ref. [122]. Copyright 2022, Elsevier.
Molecules 28 07065 g008
Figure 9. (A) The preparation process of p-QDF@M. (B) The biodistribution and mechanism of p-QDF@M for synergistic chemotherapy. Reprinted with permission from ref. [138]. Copyright 2021, Elsevier.
Figure 9. (A) The preparation process of p-QDF@M. (B) The biodistribution and mechanism of p-QDF@M for synergistic chemotherapy. Reprinted with permission from ref. [138]. Copyright 2021, Elsevier.
Molecules 28 07065 g009
Figure 10. (A) Preparation of Cu-siMDR-CDDP. (B) Mechanisms of Cu-siMDR-CDDP in overcoming drug resistance. Reprinted with permission from ref. [130]. Copyright 2021, American Chemical Society.
Figure 10. (A) Preparation of Cu-siMDR-CDDP. (B) Mechanisms of Cu-siMDR-CDDP in overcoming drug resistance. Reprinted with permission from ref. [130]. Copyright 2021, American Chemical Society.
Molecules 28 07065 g010
Figure 11. (A) Preparation of red blood cell membrane-coated “multi-in-one” core-shell metal ion-drug nanoparticles. (B) Schematic of the combination therapy in the tumor microenvironment. Reprinted with permission from ref. [152]. Copyright 2021, Elsevier.
Figure 11. (A) Preparation of red blood cell membrane-coated “multi-in-one” core-shell metal ion-drug nanoparticles. (B) Schematic of the combination therapy in the tumor microenvironment. Reprinted with permission from ref. [152]. Copyright 2021, Elsevier.
Molecules 28 07065 g011
Figure 12. (A) Preparation of tumor acid-responsive ICG@cur-Gd NPs (B) Schematic representation of ICG@cur-Gd NPs in imaging-guided cancer therapy. Reprinted with permission from ref. [132]. Copyright 2021, American Chemical Society.
Figure 12. (A) Preparation of tumor acid-responsive ICG@cur-Gd NPs (B) Schematic representation of ICG@cur-Gd NPs in imaging-guided cancer therapy. Reprinted with permission from ref. [132]. Copyright 2021, American Chemical Society.
Molecules 28 07065 g012
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zhang, X.; Hu, S.; Huang, L.; Chen, X.; Wang, X.; Fu, Y.-n.; Sun, H.; Li, G.; Wang, X. Advance Progress in Assembly Mechanisms of Carrier-Free Nanodrugs for Cancer Treatment. Molecules 2023, 28, 7065. https://doi.org/10.3390/molecules28207065

AMA Style

Zhang X, Hu S, Huang L, Chen X, Wang X, Fu Y-n, Sun H, Li G, Wang X. Advance Progress in Assembly Mechanisms of Carrier-Free Nanodrugs for Cancer Treatment. Molecules. 2023; 28(20):7065. https://doi.org/10.3390/molecules28207065

Chicago/Turabian Style

Zhang, **aoyu, Shuyang Hu, Lifei Huang, **yue Chen, **n Wang, Ya-nan Fu, Hui Sun, Guofeng Li, and **ng Wang. 2023. "Advance Progress in Assembly Mechanisms of Carrier-Free Nanodrugs for Cancer Treatment" Molecules 28, no. 20: 7065. https://doi.org/10.3390/molecules28207065

Article Metrics

Back to TopTop