Juan
Li
a,
Shan
Yan
a,
Jie
Xu
a,
Cao
Li
*a and
Qi
Yu
*ab
aKey Laboratory of Fermentation Engineering (Ministry of Education), National “111” Center for Cellular Regulation and Molecular Pharmaceutics, School of Life and Health Sciences, Hubei University of Technology, Wuhan 430068, China. E-mail: licao@hbut.edu.cn; yuqi@hbut.edu.cn
bState Key Laboratory of Organic Electronics and Information Displays & Institute of Advanced Materials (IAM), Nanjing University of Posts & Telecommunications, 9 Wenyuan Road, Nanjing 210023, China. E-mail: yuqi@hbut.edu.cn
First published on 13th August 2025
Continuous advances have been witnessed in the booming of size variable nanosystems for molecular imaging and therapy. These nanosystems usually exhibit in situ size transformation, which promotes optimized biodistribution, retention and penetration in lesions. Bioorthogonal reactions have been introduced as a useful tool to develop size variable nanosystems. In general, researchers modify controllable block (including pH adjustment, disulfide reduction, and/or enzymatic hydrolysis) masked bioorthogonal handles on the nanoparticles or small molecules to develop biocompatible size variable nanosystems. These nanosystems undergo precise click cycloaddition and self-assemble into nanoaggregates in situ, showing enhanced tissue accumulation and retention. These advantages have demonstrated great promise in improving imaging quality and therapeutic outcomes with high effectiveness and controllability. To date, this strategy has been widely introduced to construct bioimaging probes or nanomedicines. To gain a comprehensive understanding of the strategy, in this review, we focus on bioorthogonal reaction mediated size variable nanosystems reported in the last five years, present their application in bioimaging and therapy, and elucidate the mechanism of bioorthogonal reactions. Based on these efforts, challenges and future research directions in this area are also discussed at the end.
In general, the size of nanostructures affects their accumulation and penetration in targeted lesions. Designing of size variable nanosystems that enable transforming their size in situ has become an apparent and steerable strategy.15,16 As the renal filtration size is below 5.5 nm, on the one hand, nanosystems designed should be larger than 5.5 nm to allow time for diagnosis and treatment and avoid rapid renal clearance.17–19 On the other hand, nanosystems are preferable to be degraded to ultrasmall nanoparticles below 5.5 nm, which guarantees the easy clearance from the body and avoids the long-term toxicity post-diagnosis or therapy.20,21 The upper limit of parameters of nanosystems differ from hundreds of nanometers to several micrometers, which are largely affected by the vascular endothelial space in various types of tissues.22 Additionally, the size of the nanostructures also plays a crucial role in the phagocytosis from reticuloendothelial system (liver and spleen) and the exudation of blood vessels.23 Apart from the tissue accumulation and circulation, the promotion of penetration should be taken into consideration because small-sized nanoparticles (<20 nm) are more conducive to penetrate the lesions.14 In view of this idea, the fabrication of size transformation clustered nanosystems, which are amenable to smart control of their size in situ, is an appropriate way to prolong drug circulation, promote tissue accumulation and penetration.
Bioorthogonal reactions that occur within biological organisms without the intervention of their normal biochemical processes refer to a highly selective and efficient chemical tool. Pioneered by Bertozzi et al. in 2000,24 bioorthogonal reactions, involving the copper-catalyzed ones and copper free ones, have been well developed and widely used in specific protein labelling, molecular imaging and therapy due to the improved reaction rate and effectiveness.25 Currently, bioorthogonal reactions have been employed to design size variable nanosystems through the modification of the corresponding bioorthogonal reactive motifs on the molecules or nanoparticles.26,27 The intramolecular interaction facilitates the occurrence of in situ nano-to-cluster aggregation, which is beneficial to prolong the circulation time and retention time in targeted tissues. Moreover, these nanosystems can be further engineered using intelligent materials that are responsive in the presence of internal or external stimuli, such as overexpressed enzymes, light and slight acidity.28 The functionalization not only efficiently improves high specificity and precise controllability to accumulate at targeted sites, but also has potential to show secondary cluster-to-nano size transformation in nanocarriers for deep penetration in tissues. Precisely speaking, using bioorthogonal reaction-mediated size variable nanosystems is considered as an apparent and steerable strategy for benefiting optimal site-specific accumulation during diagnosis and therapy.
In this review, we will focus on bioorthogonal reaction-mediated size variable nanosystems reported in the last five years, and present an overview of the versatile strategies to optimize the biodistribution and site-specific accumulation of these nanosystems for improved molecular imaging and drug delivery (Scheme 1). In contrast to the passive targeting strategy, the in situ size variable strategy offers precise controllability of enhanced retention and accumulation of nanosystems. In the molecular imaging section, the design strategies of these contrast agents, improved signal to noise ratios (SNR) and applications in various bioimaging modes will be summarized and discussed. Furthermore, the improved retention and accumulation in targeted sites also facilitate the effective drug release and improved therapeutic efficacy. Similarly, the design strategies of these intelligent nanodrugs and improving drug delivery efficiency for the treatment of cancer, bacterial infection and other life-threatening diseases will be presented and discussed. In the end, we discuss the trends of the employment of size variable nanosystems as diagnostic and/or therapeutic platforms, followed by giving our opinions on their future direction addressing the challenges. We expect that this review will be helpful in developing more strategies to achieve precise controllability of disease diagnosis and drug delivery.
![]() | ||
Scheme 1 Schematic illustration of size transformation nanosystems for potentiating bioimaging and drug delivery. |
Apart from the entry into target tissues, particle size also has an important role in cellular uptake. Cells have their endocytosis and exocytosis processes, which usually have a size-dependent reverse mechanism to work on cellular uptake.38 Large-sized nanoparticles (>500 nm) tend to enter cells via phagocytosis,39 but are hard to be eliminated from the cells via exocytosis.40 Small ones are preferrable to internalize cells via endocytosis. For instance, nanoparticles with a parameter of 20–100 nm and 120–150 nm enter cells via caveolae and clathrin-dependent endocytosis, respectively.41,42 Based on these, a cellular uptake study should also take the size impact of nanoparticles into consideration.
![]() | ||
Fig. 1 Mechanism of bioorthogonal reactions: (A) CuAAC, (B) SPAAC, (C) iEDDA, (D) 1,2-amino thiol-CBT click reaction and (E) SPSAC. |
Imaging modalities | Probe | Responsive agent | The impact of size-variable probes on imaging | Reaction type | Size transformation, incubation time | Ref. |
---|---|---|---|---|---|---|
FL | Cbz-GPC(StBu)K(Cou)-CBT | FAP-α/GSH | “Turn-on” coumarin excimer emission, FL SNR↑ | CBT-Cys | From molecules to 72.3 ± 9.5 nm nanoaggregates, 4 h | 71 |
Cou-D/L-CBT | GSH | Enabling long-time and high-contrast fluorescence imaging | CBT-Cys | From molecules to nanotubes with an outer diameter of 156 nm, 12 h | 72 | |
Ala-Biotin-QMT | GSH/LAP | “Turn-on” the “dual AIE” fluorescence signal | CBT-Cys | From molecules to 140.9 ± 15.6 nm nanoaggregates, 6 h | 73 | |
QMT-CBT | Caspase-1/GSH | “Dual-AIE” fluorescence signal enhanced imaging of AD in vivo | CBT-Cys | From molecules to 185.5 ± 18.8 nm nanoaggregates, 4 h | 74 | |
AuNPs-Cy5.5-A&C | AEP | AEP triggered aggregation and emit strong fluorescence | CBT-Cys | From 50–60 nm to 422.2 ± 9.69 nm, 12 h | 75 | |
CyNAP-SS-FK | GSH/Cat B | FL off–on, FL SNR↑ | CBT-Cys | From molecules to ∼100 nm nanoaggregates, 12 h | 76 | |
PET | [18F]1 | GSH | Good imaging contrast and long retention time | CBT-Cys | From molecules to 138.2 ± 16.3 nm nanoaggregates | 81 |
[18F]SF-DEVD | GSH/Caspase-3 | Improve the in situ assembly efficiency and sensitive PET imaging of caspase-3 | CBT-Cys | From molecules to ∼150 nm nanoaggregates | 82 | |
[18F]SF-Glu | GGT/GSH | Improve the in situ assembly efficiency and sensitive PET imaging of GGT activity | CBT-Cys | — | 82 | |
[18F]-C-SNAT4 | GSH/Caspase-3 | Enhanced retention to improve PET imaging contrasts | CBT-Cys | From molecules to 200 nm nanoaggregates, overnight | 83 | |
[68Ga]NOTA-SFCVM [68Ga]NOTA-SFCVHEM | GSH/Cat B | Enhance tumor retention and PET signal intensity | CBT-Cys | — | 85 | |
CBT-NODA/CBT-NODA-Ga/CBT-NODA-68Ga | GSH/Furin | Prolonged tumor retention and amplified the microPET signal | CBT-Cys | From molecules to 356.8 ± 95.9 nm nanoaggregates, 7 h | 86 | |
PA | Cypate-CBT | GSH/CTSB | Enhanced retention to improve the PA signal | CBT-Cys | From molecules to 207.7 ± 15.9 nm nanoaggregates, 2 h | 92 |
NI-C-CBT | GSH/NTR | Triggered “on” and “enhanced” PA signals via intra-and intermolecular fluorescence quenching | CBT-Cys | From molecules to 117.9 ± 15.6 nm nanoaggregates, 5 h | 93 | |
MRI | DEVDCS-Gd-CBT | GSH/Caspase-3 | Effectively improved MRI sensitivity under low magnetic fields | CBT-Cys | From molecules to 85.1 ± 15.1 nm nanoaggregates, 4.5 h | 99 |
RI | nanoSABER | GSH/Legumain | Enhanced the intracellular accumulation, and improved and long-lasting Raman signals | CBT-Cys | From molecules to 147 ± 20 nm nanoaggregates, 3 h | 105 |
Yne-CBT | GSH/Cat B | Long retention in cell and relative Raman intensity↑ | CBT-Cys | — | 106 | |
QPI | P-SiO2 NPs | GSH/Legumain | In situ high-contrast refractive index imaging of enzyme activity | CBT-Cys | — | 110 |
NIR-II FL/PA | AuNNP@DEVD-IR1048 | GSH/Caspase-3 | Turn on both NIR-II FL and PA imaging signals | CBT-Cys | From 41.5 nm to 365.1 nm, 5 h | 118 |
PET/PA | [18F]-IR780-1 | GSH/Caspase-3 | Prolong probe retention and accumulation to enhance PA and PET signals | CBT-Cys | From molecules to 122 nmnanoaggregates, 1 h | 121 |
PA/MRI | Gd-IR 780 | GSH/Caspase-3 | In situ self-assembly enhanced PAI and MRI signals | CBT-Cys | From molecules to 110 ± 23 nm nanoaggregates, 2 h | 122 |
Gao et al. combined CBT-Cys cycloaddition with excimer formation to prepare a fibroblast-activated protein-α (FAP-α)-responsive “turn-on” fluorescent probe (Cbz-GPC(StBu)K(Cou)-CBT) (Fig. 2A).71 Coumarin with a planar polycyclic aromatic structure was employed as the emissive excimer fluorophore. FAP-α cleavable Cbz-Gly-Pro peptide sequence, and StBu-protected Cys and CBT units were also included in the probe. In the tumor microenvironment, FAP-α cleavage of the peptide segment and glutathione (GSH) reduction generated a Cys intermediate, which subsequently triggered intermolecular CBT-Cys click reactions to form cyclic dimers and in situ self-assembly into nanoparticles (Cou-CBT-NPs). As seen in Fig. 2B, the probe exhibited a fluorescence peak located at 475 nm, corresponding to the emission of the coumarin monomer, while the presence of 50 U mL−1 FAP-α red-shifted the emission to 550 nm, revealing the formation of the excimer. The excimer was ascribed to two coumarin molecules with a close proximity in the rigid cyclic dimer. The occurrence of in situ self-assembly led to the generation of nanoparticles with a diameter of 72.3 ± 9.5 nm via the H-aggregates (Fig. 2C). FAP-α-positive MIA PaCa-2 tumor cells and FAP-α-negative L929 normal cells were introduced to be incubated with the fluorescent probe for imaging. A time-dependent shift of the fluorescence signal from the blue to green channel was observed in MIA PaCa-2 cells due to the transition of monomer-to-excimer of the coumarin moiety (Fig. 2D). On the other hand, L929 cells only showed the enhanced fluorescence at the monomer channel. These results proved the probe sensitivity to FAP-α. Time-course fluorescence imaging of MIA PaCa-2 tumor-bearing BALB/c nude mice exhibited “turn-on” fluorescence and reached the maximum value at 6 h after the treatment with the probe, while no fluorescence was detected when the control probe removed the FAP-α sensitive peptide sequence. Additionally, Yang et al. reported a D, L alternated peptide containing probe Cys(StBu)-D-Glu-Lys-(coumarin)-D-Glu-CBT (Cou-D/L-CBT), which can react with GSH to trigger CBT-Cys cycloaddition to generate a fluorescent excimer of coumarin (Fig. 2E).72 Different from Cbz-GPC(StBu)K(Cou)-CBT, Cou-D/L-CBT included D-glutamic acid (D-Glu) that effectively underwent protonation in the lysosomal acidic environment (pH 4.5–5.5). The presence of alternated D,L-Glu in the probe facilitated the assembly into nanotubes, instead of nanoparticles. The performance prolonged the retention time of the probe in cells. The retention rates of Cou-D/L-CBT at 24 h and 36 h timepoint were 7.5 and 2.5 times higher than those of Cou-L-CBT, respectively, and the cell retention half-life was extended from 6 h (Cou-L-CBT) to 29 h (Cou-L-CBT). (Fig. 2F).
![]() | ||
Fig. 2 (A) Schematic diagram of the mechanism of a FAP-α-responsive “turn-on” fluorescent probe (Cbz-GPC(StBu)K(Cou)-CBT). (B) Fluorescence spectra of Cbz-GPC(StBu)K(Cou)-CBT treated with or without FAP-α. Inset: Corresponding pictures illuminated by an UV lamp. (C) TEM image of Cou-CBT-NPs. (D) Confocal fluorescence images of FAP-α-overexpressing MIA PaCa-2 cells or FAP-α-deficient L929 cells after incubation with Cbz-GPC(StBu)K(Cou)-CBT for 0.5 h and 2 h, adapted with permission from ref. 71. Copyright 2022, American Chemical Society. (E) Mechanism of a D, L alternated peptide containing probe Cou-D/L-CBT. (F) Intracellular fluorescence intensity over different time periods after probe incubation, adapted with permission from ref. 72. Copyright 2025, American Chemical Society. |
Commonly used fluorescent probes compromise the signal attenuation during in vivo imaging due to the aggregation-caused quenching effects at high concentrations. The emerging aggregation-induced emission luminogens (AIEgens) via limiting the molecular rotation resolve the issue. Hence, the development of an activatable fluorescent probe based on AIEgens has good potential for fluorescence imaging. Current AIEgen-based activatable probes usually form aggregates in the presence of specific targets, but the limited “turn-on” signal obtained results in compromised imaging sensitivity. Deng et al. developed an activatable AIEgen β-tBu-Ala-Cys(StBu)-Lys(Biotin)-Pra(QMT)-CBT (denoted as Ala-Biotin-QMT) to realize “two-step” aggregation in situ for enhanced fluorescence imaging (Fig. 3A).73 This probe consisted of four functional modules: β-tBu-Ala as a leucine aminopeptidase (LAP)-specific recognition site, a StBu-protected CBT-Cys click reaction group, a biotin tumor-targeting ligand, and the NIR-emissive AIE luminogen QMT. The probe targeted cell membranes through biotin–receptor interactions, followed by GSH-mediated thiol release and LAP enzymatic hydrolysis within tumor cells to generate the active intermediate Biotin-QMT. Biotin-QMT underwent a CBT-Cys click reaction to form cyclic dimers and in situ self-assembly, which completed “two-step” aggregation in cells. Experimental data validated that Ala-Biotin-QMT enhanced 4.8-fold and 7.9-fold fluorescence in HepG2 cells and HepG2 tumor-bearing mouse models in contrast to “biotin + LAP inhibitor” pre-treated control groups.
![]() | ||
Fig. 3 (A) Schematic diagram of activatable Ala-Biotin-QMT for in situ “two-step” aggregation, adapted with permission from ref. 73. Copyright 2024, American Chemical Society. (B) Schematic diagram of QMT-CBT for enhanced AD imaging. (C) Fluorescence spectra of QMT-CBT and QMT-CBT-Ctrl treated with or without caspase-1. (D) Schematic diagram of a QMT-dimer in an aqueous environment via molecular dynamics simulations. (E) Quantitative fluorescence signals of QMT-CBT and its control group in WT and AD mouse brain, adapted with permission from ref. 74. Copyright 2023, American Chemical Society. |
AIEgen-based fluorescence probes have been applied for Alzheimer's disease (AD) imaging. For instance, Ac-Trp-Glu-His-Asp-Cys(StBu)-Pra(QMT)-CBT (QMT-CBT) for enhanced AD imaging was prepared.74 As illustrated in Fig. 3B, QMT-CBT underwent GSH-triggered reduction of the disulfide bond and caspase-1(a key biomarker of AD neuroinflammation)-mediated peptide hydrolysis so that the probe induced the formation of cyclic dimers (primary aggregation) via CBT-Cys cross-linking and self-assembled nanoparticles (secondary aggregation) via hydrophobic interactions. The “two-step” aggregation efficiently turned on the AIE fluorescence in the presence of caspase-1. Most importantly, the “two-step” aggregation-based strategy led to 15.7-fold enhancement of fluorescence, which was more sensitive than that of the control group (Ac-Trp-Glu-His-Asp-Cys(tBu)-Pra(QMT)-CBT, QMT-CBT-Ctrl) with signal aggregation (Fig. 3C). Theoretical calculations demonstrated that this phenomenon was ascribed to tighter stacked QMT in “two-step” aggregation forming nanoparticles compared to QMT-CBT-Ctrl, resulting in a stronger role of limited intramolecular motions (Fig. 3D). After pre-treatment of cyclosporine to overcome the blood–brain barrier, QMT-CBT was employed for in vivo imaging of caspase-1-associated neuroinflammation containing AD mouse brain. A NIR fluorescence signal was observed in QMT-CBT treated AD mice which was 1.4-fold superior to that in the QMT-CBT-Ctrl control group (Fig. 3E). Another example is an asparagine endopeptidase (AEP)-activatable two-component nanoprobe system (AuNPs-Cy5.5-A&C) comprising functionalized gold nanoparticles AuNPs-Cy5.5-AK (containing the enzyme-cleavable peptide Ala-Ala-Asn-Cys-Lys) and AuNPs-Cy5.5-CBT. AEP-dependent click cycloaddition facilitated the utilization of the nanoprobe for AD imaging in transgenic APPswe/PS1dE9 mice.75
Apart from AIE and excimer formation, intramolecular charge transfer (ICT) is a strategy to be used for the activation of a fluorescent probe. Xu et al. constructed a GSH/cathepsin B (Cat B) dual-activatable nitrile-aminothiol (NAT) bioorthogonal probe, CyNAP-SS-FK, for specific fluorescence diagnosis of hepatocellular carcinoma (HCC) (Fig. 4A).76 Different from traditional bioorthogonal fluorogenic luminophores, this NAT one introduced a strong electron-withdrawing nitrile group in the asymmetric hemicyanine structure to obtain a donor–π–acceptor (D–π–A), thus resulting in the blocking of ICT and fluorescence quenching. The NAT luminophore was further modified with a GSH-responsive disulfide bond and a Cat B-reactive peptide moiety (Ac-FK) to obtain an activatable probe. To realize rational design of this NAT biorthogonal fluorogenic luminophore, a series of nitrile-substituted hemicyanines (CyN-X) were prepared and investigated the fluorescence. The C-6 position is preferrable to introduce electron-withdrawing groups to achieve optimized fluorescence. The spacer between reactive units and the indole ring was also investigated by either attaching an alkyl chain or a short PEG chain. The favorable electron transfer capability of the compound decorated with the flexible PEG chain (CyNAP) was demonstrated to be preferable to facilitate intramolecular macrocyclization. Consequently, CyNAP was identified for further investigation due to its superior intramolecular cyclization kinetics and highest fluorescence recovery efficiency. The first-order reaction rate constant for CyNAP's intramolecular cyclization (7.97 × 10−5 s−1) was significantly higher than its intermolecular reaction rate constant with free Cys. The rapid kinetics promoted the intramolecular cyclization pathway instead of intermolecular condensation, thereby enabling self-assembly into nanoaggregates with “turn-on” fluorescence. In vitro experiments demonstrated that CyNAP-SS-FK generated 3.9-fold higher signal intensity in HCC-LM3 cells compared to GSH inhibitor-pretreated groups. CyNAP-SS-FK was further used for real-time imaging orthotopic HCC in living mice. CyNAP-SS-FK facilitated to detect cancerous lesions with a diameter of ∼2 mm, and exhibited an improved signal to background ratio (SBR), which is ∼5-fold enhancement relative to that of the “always-on” control probe CyNAP-T (Fig. 4B). An extended detection window (∼36 h) was also detected (Fig. 4C). This CyNAP-SS-FK is different from the previously reported bioorthogonal CBT-Cys reaction mediated probes, that are intrinsically fluorescent prior to in situ self-assembly. Instead, CyNAP-SS-FK shows the “off-to-on” mode of signal response, which is beneficial to overcome the “false-positive” output and improve sensitivity and the SBR.
![]() | ||
Fig. 4 (A) Schematic illustration of CyNAP-SS-FK for specific fluorescence diagnosis of HCC. (B) SBRs of CyNAP-SS-FK or CyNAP-T. (C) Fluorescence intensity in livers of CyNAP-SS-FK or CyNAP-T injected mice at different timepoints, adapted with permission from ref. 76. Copyright 2025, Springer Nature. |
![]() | ||
Fig. 5 (A) Chemical structures of 1, 1a, and 1b. (B) MicroPET imaging of [18F]1 and [18F]1 + Block (pretreated with biotin) in HeLa tumor-bearing mice. (C) Tumor and muscle uptake of [18F]1 in HeLa tumor bearing mice at different times, adapted with permission from ref. 81. Copyright 2021, American Chemical Society. (D) Chemical structures of SF, adapted with permission from ref. 82. Copyright 2022, American Chemical Society. (E) Chemical structures of PET tracers [18F]SF-DEVD and [18F]SF-Glu. |
Another example is a novel molecule SF that was introduced with two benzyl groups attached on the abovementioned intramolecular CBT-Cys macrocylization scaffold.82 For SF, 4-(aminomethyl) benzoic acid was selected as a rigid linker and a flexible glycine residue was employed to adjust the macrocycle size (Fig. 5D). The synergetic role of both moieties optimized the intramolecular CBT-Cys cycloaddition kinetics, and improved the reaction rate, resulting in the reaction completion within 1 min and forming a stable macrocyclic structure (SF-C) under physiological conditions. The intramolecular reactivity not only eliminates the high-concentration dependency of traditional probes but also exhibits robust resistance to free Cys interference, consequently improving both in situ assembly efficacy and imaging specificity. Furthermore, SF was further modified with the radiolabelled precursor [18F]AmBF3 (18F labelled aminomethyl trifluoroborate) and an enzyme specific substrate, including a caspase-3 or γ-glutamyltranspeptidase (GGT) responsive DEVD peptide and a Glu motif to obtain two PET tracers [18F]SF-DEVD and [18F]SF-Glu (Fig. 5E). Both PET tracers have been applied for imaging of enzymatic activity in vivo. Taking caspase-3 detection as an example, HeLa tumor bearing mice were intratumorally injected with DOX to induce apoptosis for the detection of caspase-3 activity (Fig. 6A). Compared with the saline treated control group, DOX treated tumors showed significantly higher radioactivity (7.74 ± 1.56% ID per mL, 15 min post-injection). The further treatment of non-radioactive probe, [18F]SF-DEVD, further enhanced tumor uptake to 10.29 ± 1.11% ID per mL and prolonged signal retention at 30 min. The tumor-to-muscle ratio was 2.42 ± 0.54 and 6.91 ± 0.66 in [18F]SF-DEVD and the co-injection group. Compared to the previously reported [18F] tracer without a benzyl linker, [18F]SF-DEVD achieved excellent tumor uptake (7.74 ± 1.56% vs. 4.25 ± 0.97% ID per mL), and the acquiring dose of nonradioactive compounds was reduced 16 times (25 nmol vs. 400 nmol), which was attributed to the rapid intramolecular reaction mechanism of SF, enabling concentration independent self-assembly and enhancing the retention of nanoparticles in tumors. Chen et al. developed a caspase-3-sensitive PET tracer ([18F]-C-SNAT4) (Fig. 6B), in which the luciferin motif was replaced with 2-pyrimidinecarbonitrile and a benzyl linker to increase serum stability.83 In the presence of caspase-3, the occurrence of the thiol-nitrile intramolecular reaction promoted the in situ self-assembly into nanoparticles with an average size of around 200 nm. [18F]-C-SNAT4 was used to detect cell death in cisplatin-treated drug-sensitive NCI-H460 non-small cell lung cancer cells. In NCI-H460 (drug-sensitive) and NCI-H1299 (drug-resistant) tumor-bearing mouse models, PET signal intensity in NCI-H460 tumors exhibited a positive correlation with cisplatin dosage. Significantly, using PET apoptotic signals the chemotherapeutic outcomes can be predicted, obtaining consistent results with tumor volume changes. Again, the tracer successfully differentiated checkpoint inhibitor responders from non-responders in a colorectal cancer immunotherapy model, confirming its capability to predict therapeutic efficacy.
![]() | ||
Fig. 6 (A) MicroPET images of HeLa-tumor-bearing mice with or without DOX treatment, adapted with permission from ref. 82. Copyright 2022, American Chemical Society. (B) Schematic illustration of the mechanism of caspase-3-sensitive PET tracer [18F]-C-SNAT4, adapted with permission from ref. 83. Copyright 2021, Springer Nature. (C) Schematic diagram of GSH and Cat B-responsive PET probes [68Ga]NOTA-SFCVM and [68Ga]NOTA-SFCVHEM, adapted with permission from ref. 85. Copyright 2024, American Chemical Society. (D) Chemical structures of the unlabelled precursor CBT-NODA, the non-radioactive gallium complex CBT-NODA-Ga, and the 68Ga-labeled radioactive tracer CBT-NODA-68Ga, adapted with permission from ref. 86. Copyright 2021, American Chemical Society. |
Apart from 18F, 68Ga labelling PET probes have been used due to its matchable radioactive half-life (67.7 min) with the pharmacokinetics of biological molecules, such as peptides and oligonucleotides.84 Li et al. developed two Cat B-responsive PET probes, [68Ga]NOTA-SFCVM and [68Ga]NOTA-SFCVHEM (Fig. 6C).85 Both probes were integrated with lysosome-targeting morpholine and Cat B/GSH dual activatable modules for the CBT-Cys click condensation reaction, but only [68Ga]NOTA-SFCVHEM included a histidine–glutamate–histidine–glutamate–histidine–glutamate sequence (HEHEHE) to promote tumor uptake and hepatic metabolism. In vitro experiments confirmed both probes specifically recognized tumor cell lines (U87 and A549) with differential Cat B expression levels, and co-localization studies demonstrated selective localization in lysosomes. In vivo PET imaging revealed that [68Ga]NOTA-SFCVHEM and [68Ga]NOTA-SFCVM exhibited superior tumor uptake in Cat B-positive ones. HEHEHE containing [68Ga]NOTA-SFCVHEM exhibited remarkable lower liver uptake compared to [68Ga]NOTA-SFCVM. Chen et al. developed three compounds, containing a furin sensitive RVRR substrate, and StBu and CBT motifs, denoted as the unlabelled precursor AcRVRRC(StBu)K(NODAGA)-CBT (CBT-NODA), the non- radioactive gallium complex AcRVRRC(StBu)K(NODAGA-Ga)-CBT (CBT-NODA-Ga), and the 68Ga-labelled radioactive tracer AcRVRRC(StBu)K(NODAGA-68Ga)-CBT (CBT-NODA-68Ga) (Fig. 6D).86 These compounds have been demonstrated to activate the CBT-Cys click reaction in a furin-rich tumor microenvironment. Experiments demonstrated that sole injection of the radioactive tracer CBT-NODA-68Ga resulted in nonspecific condensation between unpurified residual precursors and endogenous Cys in tumor cells, thereby limiting nanoparticle generation efficiency. Instead, co-injection of the 68Ga-labelled PET tracer and its nonradioactive analogue blocked nonspecific interactions, ensuring efficient synthesis of hybrid 68Ga nanoparticles within tumor cells, which prolonged tumor retention and amplified the microPET signal.
![]() | ||
Fig. 7 (A) The mechanism of PA probe Cypate-CBT. (B) Time-dependent PA intensity of probes in each group after incubation with MDA-MB-231 cells. (C) Time-dependent PA intensity after injection of probes in MDA-MB-231 tumors, adapted with permission from ref. 92. Copyright 2021, Wiley-VCH GmbH. (D) Schematic diagram of the mechanism of NI-C-CBT, adapted with permission from ref. 93. Copyright 2023, Wiley-VCH GmbH. |
Another example is a nitroreductase (NTR)-responsive smart photoacoustic probe, NI-C-CBT, comprising 2-nitroimidazole (NI), StBu-protected cysteine-diaminopimelic acid (Cys(StBu)-Dap), near-infrared (NIR) chromophore IR780 and CBT (Fig. 7D).93 In the hypoxic tumor microenvironment, elevated GSH and NTR synergistically triggered the reductive cleavage of NI-C-CBT to release C-CBT, and underwent the CBT-Cys bioorthogonal click reaction. Dimerization and self-assembly of the probe subsequently induced intramolecular and intermolecular fluorescence quenching of IR780, switching on and amplifying the PA signal. In vitro experiments demonstrated a 1.9-fold overall PA signal enhancement in hypoxic HeLa cells compared to normoxic cells. Further investigation demonstrated highly sensitive and specific photoacoustic imaging of tumor hypoxia in a HeLa tumor-bearing mouse model.
![]() | ||
Fig. 8 (A) Schematic illustration of caspase-3-responsive probe DEVDCS-Gd-CBT for T1-weighed MRI imaging of apoptosis. (B) T1-weighted MR images and the lesion/body grayscale value ratios of MR images of zebrafish in different groups, adapted with permission from ref. 99. Copyright 2023, American Chemical Society. |
![]() | ||
Fig. 9 (A and B) Schematic diagram of nanoSABER for targeted tumor RI imaging. (C) Ratio of the alkyne to nitrile Raman peak intensities of DU145, LNCaP and RWPE1 cells treated with nanoSABER. (D) Score plot of MCR2 versus MCR3 for the DU145 and LNCaP cells, adapted from ref. 105, under the license CC-BY, published by Wiley-VCH GmbH. (E) Schematic illustration of Yne-CBT for the detection of intracellular enzyme activity. (F) Thermogram of CTSB enzyme activity in a single MDA-MB-231 cell, adapted with permission from ref. 106. Copyright 2024, American Chemical Society. |
Wang et al. designed a self-referencing Raman probe, Val-Cit-Cys(StBu)-Pra-Gly-CBT (Yne-CBT), for the detection of intracellular enzyme activity (Fig. 9E).106 The integration of Yne-CBT with intracellular overexpressing Cat B formed long-retained cyclic dimers, which displayed a distinct increase in the ratio of the alkyne to nitrile Raman signals. Additionally, uniform Au@SiO2 NPs were used to enhance the Raman signal with a detection limit of 61.4 U L−1. Using a custom microfluidic channel coupled with confocal Raman microscopy, real-time monitoring of MDA-MB-231 cells (high Cat B expression) co-incubated with Yne-CBT and Au@SiO2 NPs demonstrated dynamic shifts in relative signal intensities at 2120 and 2227 cm−1, confirming Cat B-activated CBT-Cys click reactions. Enzymatic activity heatmaps revealed significant Raman signal variability due to cellular heterogeneity (Fig. 9F).
![]() | ||
Fig. 10 (A) Schematic diagram of P-SiO2 NPs for QPI. Volume percentage of the refractive index between 1.4 and 1.5 in DU145 (B) and LNCaP (C) cells treated with different groups, adapted with permission from ref. 110. Copyright 2023, American Chemical Society. |
![]() | ||
Fig. 11 (A) Schematic diagram of AuNNP@DEVD-IR1048 for imaging of apoptosis. (B) Bio-TEM images of AuNNP@DEVD-IR1048-treated HepG2 cells under varying X-ray doses (2–8 Gy). Photoacoustic (C) and NIR-II fluorescence (D) intensity of AuNNP@DEVD-IR1048-treated HepG2 cells under varying X-ray doses. (E) Western blot analysis of cleaved caspase-3 expression as a marker of apoptotic activation. (F) Pearson's correlation coefficient of the relationship between the relative tumor volume changes and the caspase-3 expression, the ΔPA changes or the ΔFL changes after RT, adapted with permission from ref. 118. Copyright 2021, Wiley-VCH GmbH. |
The 2-cyano-6-hydroxyquinoline (CHQ)-Cys macrocyclization reaction has been commonly used to design activatable probes based on Cys-Luc-CHQ or Cys-Ben-PMN scaffolds for signal mode molecular imaging, such as PET, FL, or PA imaging.119,120 However, the introduction of these scaffolds to design probes for multimodal imaging is challenging due to the difficulty in installing two imaging tags together in one molecule. Also, the two imaging tags introduced would amplify the molecular size and induce steric hindrance that slows down the macrocyclization kinetics. To address this issue, Wang et al. designed a caspase-3 activatable probe ([18F]-IR780-1) based on the triazole IR780 scaffold for PA/PET bimodal imaging of apoptosis (Fig. 12A).121 [18F]-IR780-1 modified CBT-Cys reaction pairs and 18F labelled zwitterionic trifluoromethyl borate ([18F]-AMBF3) onto the IR780 scaffold to enable reliable PA/PET dual-modal imaging. To optimize macrocyclization kinetics, different acyclic precursors that were differed in linkers containing no glycine residue (8a) or one glycine residue (8b) on the D-Cys sites were screened and designed. Experimental results revealed that the macrocyclization kinetics of 8a (first-order rate constant k = (3.9 ± 0.1) × 10−3 s−1) were significantly faster than that of 8b (k = (2.6 ± 0.2) × 10−3 s−1) (Fig. 12B). The half-life of 8a was comparable to those of previously reported scaffolds like Cys-Luc-CHQ or Cys-Ben-PMN. In [18F]-IR780-1, the caspase-3 cleavage collaborated with GSH reduction facilitated in situ self-assembly into nanoparticles, which quenched the fluorescence of IR780 and enhanced the PA signal. The cold compound of [18F]-IR780-1, IR780-1 has demonstrated 4-fold enhancement of the PA signal at 855 nm (Fig. 12C). In U87MG tumor-bearing mice, pretreatment with DOX and [18F]-IR780-1 significantly enhanced the PET signal and PA intensity in tumors, compared to the saline group, and the group preinjected with a caspase-3 inhibitor, Z-VAD-fmk (Fig. 12D). Moreover, the PET tracer in the IR780 fluorophore based bimodal probe can be replaced by Gd-DOTA for PA and MRI imaging of tumor apoptosis (Fig. 13A).122 The caspase-3 triggered self-assembly concurrently turned on the PA signal (∼4.3-fold at 855 nm) (Fig. 13B) and increased r1 from 7.98 to 19.66 mM−1 s−1 in MRI at 0.5 T (Fig. 13C).
![]() | ||
Fig. 12 (A) Schematic diagram of [18F]-IR780-1 based on the triazole IR780 fluorophore for PA/PET bimodal imaging of apoptosis. (B) Calculation of first-order constants of intramolecular macrocyclization of 8a and 8b. (C) PA images (insets) and normalized PA intensities acquired at 790 and 855 nm before and after incubation of IR780-1 with caspase-3. (D) Representative axial and coronal PET images and transverse PA images of U87MG tumors in living mice after different treatments, adapted with permission from ref. 121. Copyright 2022, Wiley-VCH GmbH. |
![]() | ||
Fig. 13 (A) Schematic diagram of Gd-IR780 for PA and MRI imaging of tumor apoptosis. (B) PA images (insets) and normalized PA intensities acquired at 790 and 855 nm before and after incubation of Gd-IR780 with caspase-3. (C) Comparative r1 analysis of Gd-IR780 based on 1/T1versus Gd concentration plots before and after caspase-3 incubation, adapted with permission from ref. 122. Copyright 2022, Elsevier. |
Treatment mode | Material | Responsive agent | Reaction type | Disease | Size transformation, incubation time | Ref. |
---|---|---|---|---|---|---|
Chemotherapy | SPr-CPT@PEG | GSH | CBT-Cys | 4T1 tumor | From molecules to fusiform in shape with a length of 400 nm and a width of 20 nm | 131 |
iCPDNDBCO/iCPDNN3 | pH 6.5 | SPAAC | 4T1 tumor | From 107 nm to 1216 nm, 15 min; from 1216 nm to 10 nm, 24 h | 134 | |
NP@DOXDBCO + iCPPAN3 | pH 6.5 | SPAAC | 4T1 tumor | From 112.0 nm to 848.1 nm, 15 min; from 848.1 nm to 10 nm, 12 h | 135 | |
AuNPs-D&H-R&C | Furin | CBT-Cys | MCF-7/ADR tumor | From 38.2 ± 1.2 nm to 263.2 ± 10.6 nm, 48 h | 136 | |
Cip-CBT-Ada/CD-M | GSH/Caspase-1 | CBT-Cys | S. aureus infection | From molecules to 20.0 ± 2.5 nm nanoaggregates, 18 h | 137 | |
Ag-P&C NPs | pH 6.5 | CBT-Cys | MRSA-nfected wounds, anti-biofilm models, periodontitis | From ∼50 nm to ∼400 nm, 24 h | 140 | |
MEM/DON/INS ICV | Legumain | CBT-Cys | SAMP8 mice | From 198–203 nm to 800 nm–4.5 μm, 6 h | 141 | |
Photothermal therapy | AuNP@1 | GSH/Furin | CBT-Cys | MDA-MB-468 tumor | From 25.7 ± 10.2 nm to 103.5 ± 12.3 nm, 10 h | 147 |
SPIO@1NPs | GSH/Furin | CBT-Cys | MDA-MB-468 tumor | From ∼50 nm to 139.72 ± 24.69 nm, 6 h | 148 | |
AuNPs-ImLND and AuNPs-DBCO-RGD | DBCO | BCR | 4T1 tumor | From 16–24 nm to over 1000 nm, 12 h | 149 | |
Cu-POM NCs, antibacterial molecule 6 | Cu-based catalysts | CuAAC | S. aureus and E. coli infected biofilm | From 7–10 nm to 500 nm (pH 4.5), 22.5 h | 150 | |
CNDs(CNDs@N3 and CNDs@C![]() |
Cu-based catalysts | CuAAC | VRE bacterial infection | From 6 ± 4 nm to 40 ± 20 nm | 151 | |
Radiotherapy | MnAuNP-C&B | GSH | CBT-Cys | CT26 tumor | From 29 nm to over 200 nm, 24 h | 156 |
[131I]IM(HE)3 AAN | Legumain/GSH | CBT-Cys | HCT116 tumor | From molecules to 86 ± 9 nm | 157 | |
Photodynamic therapy | P-FFGd-TCO + 775NP-Tz + SA-Tz | ALP | iEDDA | HeLa tumor | From 40–240 nm to ∼1 μm, 15 min | 162 |
Chemodynamic therapy | cPFCDBCO and cPFCN3 | pH 6.5 | SPAAC | 4T1 tumor | From ∼100 nm to 995.07 nm, 15 min; from 995.07 nm to 10 nm, 12 h | 167 |
Synergistic therapy | D-NP and C-NP | pH 6.5 | CBT-Cys | 4T1 tumor | From ∼65 nm to 2611 ± 115 nm, 24 h | 169 |
SIA-αTSLs | CTSB | CBT-Cys | CT26 tumor | From 148.13 ± 4.24 nm to 498.70 ± 23.89 nm, 2 h; from ∼600 nm to small sized nanoparticles, 10 min | 172 | |
NGOPC@PTX | Legumain | CBT-Cys | 4T1 tumor | From ∼60 nm to ∼900 nm, 24 h | 173 | |
Ce6-Leu@Mn2+ | LAP/GSH | CBT-Cys | HepG2 tumor | From molecules to ∼80 nm aggregates, from ∼80 nm to ∼23 nm | 174 | |
NPCe6-DBCO and TK-PAMAMPR104A-N3 | pH 6.5 | SPAAC | 4T1 tumor | From ∼100 nm to ∼740 nm; from 740 nm to 10 nm, 0.5 h | 175 |
![]() | ||
Fig. 14 (A) Schematic diagram of the SPr-G2 dendrimer and the mechanism of self-assembly. (B) MHC-I expression on 4T1 cells. (C) Schematic illustration of the synthesis of SPr-CPT@PEG. (D) Primary and distant tumor weights after treatment with different groups, adapted with permission from ref. 131. Copyright 2024, Wiley-VCH GmbH. (E) Schematic diagram of tumor acidity and a bioorthogonal reaction dual-responsive nanosystem to co-deliver NO donor and chemotherapeutic drug DOX for enhanced chemotherapy. (F) FL images of 4T1 murine models after i.v. administration of: (I) iCPDNCy5.5 + iCPDNDBCO/Cy5.5 and (II) iCPDNN3/Cy5.5 + iCPDNDBCO/Cy5.5 at different time points. (G) Quantification of the DOX concentration in major organs and tumor tissues at 24 h post-injection with DOX, iCPDN+ iCPDNDBCO, and iCPDNN3 + iCPDNDBCO, adapted with permission from ref. 134. Copyright 2022, American Chemical Society. |
Apart from insufficient retention, another main challenge in cancer chemotherapy is chemo-resistance.132 A tumor hypoxic environment upregulates hypoxia-inducible factor-1α (HIF-1α), which has been reported to enhance the expression level of P-glycoprotein, cause drug efflux and contribute to chemo-resistance.133 Hence, researchers developed various strategies to overcome hypoxia and improve the treatment outcome of chemotherapy. For instance, Wang et al. employed a nitric oxide (NO) donor to increase oxygen (O2) perfusion within a tumor to overcomer hypoxia and introduced a tumor acidity and bioorthogonal reaction dual-responsive nanosystem to co-deliver the NO donor and a chemotherapeutic drug, doxorubicin (DOX), for enhanced chemotherapy (Fig. 14E).134 In this nanosystem, DOX was conjugated to polyamidoamine (PAMAM) via an acid-sensitive hydrazone bond to obtain PAMAM-DOX. PAMAM-DOX was further introduced with thiol groups and linked with tert-butyl nitrite to generate ultrasmall PAMAM-DOX/NO (PND) containing a pH sensitive DOX prodrug and a NO donor. PDN was cross-linked with pH-cleavable maleic acid amide to form clustered nanoparticles (iCPDN) (107 nm). iCPDN were modified by an azide (N3) motif containing poly(ethylene glycol) (PEG) or a dibenzocyclooctyne (DBCO) motif containing poly(2-azepane ethyl methacrylate) (PAEMA). In a normal physiological pH environment, DBCO was shed in the nanoparticles, while in a tumorous environment, mild acidity slowly deprotonated PAEMA, then exposed the DBCO moiety, and bioorthogonally reacted with N3 containing PEGylated iCPDN, thus allowing rapid formation of nanoaggregates (∼1216 nm) within 15 min. Subsequently, as the maleic acid amide moieties were cleaved, the size of nanoaggregates was shifted to 10 nm with the incubation time prolonged to 24 h. Cy5.5 labelled iCPDNN3 and iCPDNDBCO were introduced to track the biodistribution in the tumor bearing mice model (Fig. 14F). Markedly enhanced accumulation and a prolonged retention time within tumor tissue were observed in the iCPDNN3/Cy5.5 + iCPDNDBCO/Cy5.5 group. This demonstrated that the “two-step” size transformation facilitated superior tumor retention. The delivery of DOX to hypoxic tumor tissues is also revealed in Fig. 14G. The released NO could reverse hypoxia-induced chemo-resistance by downregulating HIF-1α levels, enhance the chemotherapeutic efficacy of DOX, and reprogram the tumor immune microenvironment to boost antitumor immune responses. Additionally, authors also utilized the same strategy to prepare DBCO masked DOX prodrug nanoparticles and N3 PEGylated hypoxia-activated prodrugs PR104A to construct a “two-step” size-transformation nanosystem (Fig. 15A), which efficiently overcame hypoxia and enhanced chemotherapy efficacy.135
![]() | ||
Fig. 15 (A) Schematic diagram of a dual-responsive nanosystem designed to overcome hypoxia-induced intratumoral heterogeneity and improve chemotherapy efficacy, adapted with permission from ref. 135. Copyright 2022, Elsevier. (B) The mechanism of furin protease sensitive AuNPs-D&H-R&C, adapted with permission from ref. 136. Copyright 2021, Elsevier. (C) Schematic illustration of Cip-CBT-Ada/CD-M for combating S. aureus infections in macrophages. (D) Fluorescence microscopy imaging of normal RAW 264.7 cells incubated with NBD-CBT-Ada/CD-M, S. aureus-infected RAW 264.7 cells treated with different groups, adapted with permission from ref. 137. Copyright 2023, Wiley-VCH GmbH. |
Different from the modulation of tumor microenvironments, the inhibition of cancerous pro-survival pathways is also a desirable strategy to decrease chemo-resistance. Xie et al. designed a tumorous overexpressed furin protease sensitive Au NP-based nanoplatform to co-deliver DOX and hydroxychloroquine (HCQ), which inhibited one of the pro-survival pathways, cell autophagy (Fig. 15B).136 In the nanoplatform, Au NPs were co-loaded with DOX and HCQ, and were modified with the RK peptide (RVRRCK) and 2-cyanobenzothiazole-polyethylene (CBT-PE), respectively. Furin cleaved the RK peptide sequence and subsequently promoted the CBT-Cys reaction, thus leading to the exposure of 1,2-aminothiol groups on Cys to form AuNP aggregates. The efficient uptake of AuNP aggregates delivered sufficient HCQ and DOX to MCF-7/ADR cells. HCQ interfered cell autophagy and promoted autophagic degradation, indicated by the blocking of more LC3-II recycling to LC3-I and high expression of p62 protein. Furthermore, HCQ plays a synergistic role with DOX in intensifying DNA damage and apoptosis pressure by activating the p53 signalling pathway. The nanoplatform was further used in chemo-resistant MCF-7/ADR breast tumors in vivo, eliciting superiority over free drug delivery systems in decreasing systemic toxicity.
Bioorthogonal chemistry-mediated chemotherapy can be also applied for different kinds of disease treatments. For instance, Zhan et al. designed a supramolecular antibiotic delivery system, Cip-CBT-Ada/CD-M, for combating S. aureus infections in macrophages.137 The antibiotic-peptide conjugate, Cip-CBT-Ada, consisted of ciprofloxacin (Cip), a caspase-1/GSH reactive bioorthogonal CBT-Cys module, and an adamantane (Ada) moiety (Fig. 15C). The antibiotic-peptide conjugate was linked with β-cyclodextrin-heptameric mannoside (CD-M) through guest–host recognition, and showed a strong affinity to macrophages due to abundant mannose receptors on the surface. Within infected macrophages, elevated GSH and caspase-1 exposed reactive Cys groups triggered CBT-Cys click reactions. Cyclic Cip-dimers were further formed and subsequently self-assembled into Cip nanoparticles (Nano-Cip, ∼20 nm) via hydrophobic interactions. The following esterase-mediated hydrolysis promoted the Cip release to eradicate intracellular S. aureus. 7-Nitro-1,2,3-benzoxadiazole (NBD)-labelled NBD-CBT-Ada was prepared to investigate the uptake of S. aureus-infected macrophages. Only infected cells exhibited NBD fluorescence at 2 h incubation (Fig. 15D). Mannose pretreatment (receptor blockade) or the CD-M motif-absent group decreased fluorescence by 75.8% and 65.8%, respectively, verifying mannose receptor-mediated endocytosis. The control group treated GSH inhibitors or caspase-1 inhibitors led to 62.1% or 85.8% decrease of fluorescence, confirming GSH/caspase-1-dependent activation of self-assembly. Cip-CBT-Ada/CD-M treated S. aureus-infected RAW264.7 cells and mouse infection models demonstrated superior bactericidal efficacy and inflammation alleviation capability. Silver nanoparticles (Ag NPs) are commonly used as antibiotics because they can release Ag+ to inhibit ATP generation and affect the membrane structure.138,139 Delivering sufficient Ag NPs to the biofilm is beneficial to antibacterial efficacy. Cheng et al. developed pH-responsive Ag NPs (Ag-P&C NPs) based on bioorthogonal chemistry, and employed an intelligent size-regulation strategy for drug-resistant bacterial infections (Fig. 16A).140 PEGylated Ag NPs were modified with a peptide sequence (NH2-Lys-Arg4-Gly-His4-Cys-CM) and a CBT moiety, respectively. In the neutral environment of healthy tissues, the peptide chain folded into a U-shaped conformation via hydrogen bonding, shedding the active thiol in the Cys group. Upon entering the acidic microenvironment of bacterial infection sites, the protonation of histidine imidazole groups induced surface charge reversal and triggered the following CBT-Cys cycloaddition. The system has been proved to acquire efficient bactericidal effects through multiple mechanisms, including disrupting membrane integrity, inhibiting ATP synthesis, and generating ROS. The system was successfully applied in three animal infection models, including MRSA-infected wounds, and for eradication of biofilms established on the surface of bone implants, ameliorating osseointegration, and treating periodontitis.
![]() | ||
Fig. 16 (A) Schematic illustration of pH-responsive Ag-P&C NPs, adapted with permission from ref. 140. Copyright 2022, Wiley-VCH GmbH. (B) Schematic diagram of bioorthogonal reaction-mediated crosslinked vesicles to intracerebrally co-deliver insulin, donepezil hydrochloride and memantine hydrochloride for AD treatment. (C) Distribution of fluorescent nanovesicles in biological tissues treated with either RhBVesicleAK or RhBICV after 12 h, adapted from ref. 141, under the license CC-BY-NC-ND 4.0, published by Elsevier. |
Another example is bioorthogonal reaction-mediated crosslinked vesicles to intracerebrally co-deliver insulin, donepezil hydrochloride (DON HCl) and memantine hydrochloride (MEM HCl) for Alzheimer's disease treatment.141 These vesicles are respectively coated with AD brain parenchyma specific expressed legumain cleavable AK peptide sequences (Ac-Ala-Ala-Asn-Cys-Asp) and the CBT unit, thus triggering the CBT-Cys cross-linking reaction and forming in situ aggregation, which effectively prevented the drug efflux pumps of the brain (Fig. 16B). Multi-drugs were validated to exhibit a neuroprotective effect and improve memory ability of SAMP8 mice. Most importantly, the cross-linking strategies improved the intracerebral retention in vivo and overcame the brain barrier efflux. The distribution of RhB labelling cross-linked vesicles revealed that the brain and liver total retention was more than 91% (Fig. 16C), while other major organ retention was less than 9% at 12 h, providing good potential for chemotherapy of the AD brain.
![]() | ||
Fig. 17 (A) Schematic of the mechanism of AuNP@1 for enhanced PTT in tumors. (B) TEM images of AuNP@1 treated with GSH and furin. (C) Visible–NIR absorption spectra of AuNPs, AuNP@1 or AuNP@1-Scr treated with GSH and furin. (D) The temperature curves of different groups irradiated by 808 nm NIR laser, adapted with permission from ref. 147. Copyright 2020, Wiley-VCH GmbH. (E) Schematic illustration of a BCR reaction based two-component nanoplatform for synergistic PTT and chemotherapy. (F) The photothermal conversion efficiency of different AuNP aggregation. (G) PA imaging of different groups. (H) HPLC results of the LND released from aggregated AuNPs, adapted with permission from ref. 149. Copyright 2024, Wiley-VCH GmbH. |
Bioorthogonal “Click and Release” (BCR) reactions, in which an unstable intermediate is formed via cycloaddition and then rapidly decomposes to release a molecular motif, offer the opportunities to involve nano-to-cluster conversion and drug release simultaneously. Yan et al. developed a BCR reaction based two-component nanoplatform for synergistic PTT and chemotherapy.149 In the nanoplatform, a specific BCR reaction was introduced based on [3+2] cycloaddition between a mesoionic prodrug iminosydnone-lonidamine (ImLND) and DBCO to trigger in situ nanoassembly of AuNPs and release chemotherapeutic drug lonidamine (LND) that can downregulate heat shock protein expression and avoid the cellular protection from hyperthermia damage (Fig. 17E). The two-component nanoplatform using HS-PEG5000-NH2 as a linker is composed of ImLND grafted AuNPs and DBCO-RGD grafted AuNPs. In the therapeutic regimen, DBCO-RGD grafted AuNPs specifically located at tumor sites through the specific affinity between RGD and integrin αvβ3. Subsequently, ImLND grafted AuNPs were administered, so that the BCR reaction triggered [3+2] cycloaddition to promote the AuNP aggregation and simultaneous release of LND. The generation of nanoaggregates through the BCR reaction facilitated excellent photothermal effects with a photothermal conversion rate of 57.3% and activated a strong photoacoustic signal (Fig. 17F and G). On the other hand, the release profile of LND was revealed by HPLC and LC-MS with the observation of an absorption peak at an elution time of 7.903 min (Fig. 17H) and the corresponding peak m/z = 406.0819, respectively. The release of LND in cells downregulated the HSP expression, which was confirmed via western-blotting assay. The synergistic roles of released LND and aggregation enhanced PTT facilitated a robust tumor suppression effect in vitro and in vivo.
Cu(I) catalyzed bioorthogonal CuAAC reaction has been employed to design prodrugs, but the introduction of a Cu(I) catalyst is harmful to living cells or organisms and the instability of Cu(I) compromised its catalytic activity. Qu et al. developed copper-doped molybdenum-based polyoxometalate nanoclusters (Cu-POM NCs) (Fig. 18A), which selectively self-assembled into large aggregates in an acidic bacterial biofilm through the hydrogen bonding and efficiently catalyzed the CuAAC reaction to synthesize antibacterial molecule 6 for antibiofilm therapy.150 Additionally, Cu-POM NCs consumed bacterial H2S through the MoVI-to-MoV conversion and enhanced the NIR-II photothermal effect for PTT (Fig. 18B). TEM images exhibited monodispersed spherical Cu-POM NCs with a diameter of 7–10 nm at pH 7.2, whereas formed nanoaggregates with a diameter of ∼220 nm at pH 5.5. The aggregates did not affect the catalytic activity, which was demonstrated by the consistent results in the model CuAAC reaction between 3-azido-7-hydroxycoumarin and phenylacetylene at different pH buffers. The time-dependent absorption spectra of Cu-POM treated with H2S showed gradually increased absorption ability in the NIR-II region. Acid/H2S dual responsiveness promoted the excellent photothermal effects with a photothermal conversion efficiency of 49.2%. This platform integrates bioorthogonal catalysis with NIR-II photothermal effects, offering a synergistic strategy to combat biofilm infections and reduce bacterial tolerance.
![]() | ||
Fig. 18 (A) Schematic illustration of Cu-POM NCs selectively self-assembled into large aggregates and efficiently catalyzed the CuAAC reaction to synthesize antibacterial molecule 6 for antibiofilm therapy. (B) NaHS-dependent temperature increases irradiation by 1064 nm NIR laser and the mechanism, adapted with permission from ref. 150. Copyright 2023, Wiley-VCH GmbH. (C) Schematic illustration of covalent assembly of CNDs for improved PTT. (D) The calculated HOMO and LUMO levels of CNDs@N3, CNDs@C![]() |
Another example is covalent assembly of carbon nanodots (CNDs) via a CuAAC reaction in a water-in-oil emulsion system for improved PTT.151 The system comprised azide-modified CNDs@N3 and alkynylated CNDs@CC (Fig. 18C), which formed triazole covalent crosslinks through CuAAC to yield nanocomposites with a particle size of 40 ± 20 nm. Compared to free CNDs with a diameter of (6 ± 4 nm), covalent assembled CNDs reduced the electronic transition bandgap from 3.45 eV to 3.05 eV (Fig. 18D) and shifted the energy dissipation pathway from radiative decay to non-radiative decay, which resulted in fluorescence quenching and the enhancement of NIR absorption from 600 to 1100 nm. Upon laser irradiation at 808 nm, covalent assembled CNDs possessed a photothermal conversion efficiency of 32.34%. The assembled CNDs were further applied for in vitro and in vivo photothermal antibacterial efficacy. According to standard plate counting assays, the assembled CNDs showed 99% bactericidal efficiency against vancomycin-resistant enterococcus (VRE). In a subcutaneous VRE-infected mouse model, NIR irradiation plus assembled CND treatment led to almost eradication of the infected abscessed area on the third day, giving strong evidence of excellent photothermal antibacterial efficacy of the assembled CNDs in vivo.
![]() | ||
Fig. 19 (A) Schematic illustration of intelligent nanoplatform MnAuNP-C&B, adapted with permission from ref. 156. Copyright 2024, American Chemical Society. (B) Chemical structure and the mechanism of [131I]IM(HE)3AAN. |
![]() | ||
Fig. 20 (A) Schematic diagram of a tumor pre-targeting theranostic strategy to alleviate tumor hypoxia and achieve multimodal imaging-guided PDT. (B) Epifluorescence imaging of HeLa tumor cells in different treatment groups. I: P-FFGd-TCO, II: P-FFGd-TCO and 775NP-Tz (containing R6G) + SA-Tz; III: P-FFGd-TCO and 775NPs-OMe (containing R6G) + SA-Tz; IV: 775NP-Tz (containing R6G) + SA-Tz. (C) Bio-SEM images of blank HeLa cells, and HeLa cells incubated with P-FFGd-TCO or P-FFGd-TCO + 775NP-Tz + SA-Tz, adapted with permission from ref. 162. Copyright 2023, Wiley-VCH GmbH. |
![]() | ||
Fig. 21 (A) Schematic diagram of the mechanism of cPFCDBCO and cPFCN3 nanoparticles for enhanced tumor therapy. (B) Confocal fluorescence imaging of ˙OH generation in 4T1 cells treated with different groups under normal and hypoxic conditions, adapted with permission from ref. 167. Copyright 2022, Elsevier. |
![]() | ||
Fig. 22 (A) Schematic diagram of bioorthogonal in situ assembled nanodepots for enhanced tumor accumulation and cocktail therapy. (B) Intravital CLSM real-time image of Cy5 labeled D-NP/C-NP and S-NP/C-NP in GFP-4T1 tumor-bearing mice, adapted from ref. 169, under the license CC-BY, published by Springer Nature. (C) Schematic diagram of a dual mode imaging guided nanotheranostic agent based on bioorthogonal click chemistry for precise cancer treatment. (D) Abemaciclib release curves for different groups. (E) Cell cycle arrest of CT26 cells treated with different groups, adapted from ref. 172, under the license CC-BY, published by Wiley-VCH GmbH. |
PTT has been testified to trigger cells to release damage-associated molecular patterns (DAMPs), such as heat shock proteins (HSP), high mobility group protein B1 (HMGB1), and calreticulin (CRT), which can promote maturation of dendritic cells (DC) and motivate host's immune system to combat tumors.170,171 Therefore, PTT can be combined with immunotherapy for cancer treatment. Sang et al. designed themoresponsive liposomes to codeliver magnetic Fe3O4 nanoparticles, chemotherapeutic drug Abemaciclib and photothermal reagents IR780 (Fig. 22C).172 The liposomes were modified with the ACKFRGD peptide sequence or CBT motif so that these liposomes could reach tumor sites and form aggregates in the presence of highly expressed Cat B via CBT-Cys cycloaddition. Upon laser irradiation, SPIONs/IR780 generated localized hyperthermia to trigger liposomal phase transition. Abemaciclib and IR780 were released in deep tumor tissues resulting in a synergistic photothermal/chemo/immuno therapeutic cascade. Experimental validation demonstrated that laser irradiation of in situ aggregated liposomes significantly enhanced Abemaciclib release and induced the highest rate of G1 phase arrest (49.42 ± 2.65%) (Fig. 22D and E), promoted release of CRT, HMGB1, ATP and increased T-cell infiltration, leading to potent tumor suppression and long-term immune memory. Moreover, magnetic resonance/NIR fluorescence imaging is beneficial to precisely localize the liposomes in vivo and guide a multimodal therapy. Another example is a legumain-responsive in situ self-assembled nanosystem containing a photothermal reagent, nano-graphene oxide (NGO) and paclitaxel (PTX) (Fig. 23A).173 Legumain triggered the specific CBT-Cys cross-linking to generate large-size aggregates, which ensured prolonged retention and sustained PTX release. As illustrated in Fig. 23B, NGOPC@PTX exhibits a sustained drug release profile in the presence of 4T1 lysates. The 808 nm laser irradiation slightly enhanced PTX liberation by 9.5%, indicating a photothermal release behavior. Ex vivo fluorescence imaging of tumors and major organs of mice bearing IR783 labeled NGOPC revealed maximal fluorescence intensity localized within tumor regions, whereas a negligible signal was detected in liver, indicating that the formed aggregates augmented the tumor accumulation and benefited for drug retention (Fig. 23C). Accompanied by the blocking immune escaping role of PTX, the nanosystem promoted DC maturation via binding to Toll-like receptors (TLRs) on DC surfaces and selectively killed regulatory T cells (Tregs) through Bcl-2/Bax-mediated apoptosis, thereby promoting a shift from an immunosuppressive to immunostimulatory tumor microenvironment.
![]() | ||
Fig. 23 (A) Schematic illustration of a legumain-responsive in situ self-assembled nanosystem. (B) In vitro PTX release curves of different groups. (C) Ex vivo fluorescence imaging of major organs and tumors in different groups, adapted with permission from ref. 173. Copyright 2020, Elsevier. (D) Schematic illustration of GSH and LAP-responsive size-transformable smart nanoprobe Ce6-Leu, adapted with permission from ref. 174. Copyright 2021, American Chemical Society. |
PDT has also been commonly used to work with other modalities to enhance therapeutic efficacy. Wang et al. designed a size-transformable smart nanoprobe, Ce6-Leu, capable of specifically responding to leucine aminopeptidase (LAP) and GSH (Fig. 23D).174 By chelating with Mn2+ to obtain Ce6-Leu@Mn2+, the probe realized efficient multimodal tumor imaging guided photodynamic/radiotherapy. Ce6-Leu@Mn2+ consists of three components: the photosensitizer Ce6, a leucine and disulfide bond unit, and a CBT group. The amphiphilic characteristics induced the probe to self-assemble as nanoparticles (∼80 nm) via the hydrophobic interactions and π–π stacking of Ce6, while the presence of LAP and GSH spontaneously cleaved the probe to yield its cyclic dimer via CBT-Cys cycloaddition with a diameter of ∼23 nm. More importantly, the Mn2+-chelation in the probe exhibited peroxidase-like catalytic ability to transfer endogenous H2O2 to O2, which was beneficial to PDT and radiotherapy in a hypoxic environment. Jiang et al. reported DBCO masked photosensitizer Ce6 nanoparticles and N3 functionalized hypoxia-activated prodrugs PR104A to construct “two-step” size-transformation nanodepots (Fig. 24A).175 Bioorthogonal click reaction-induced macro-aggregates were formed within tumor adjacent regions, enabling effective co-delivery of prodrugs and photosensitizers. Upon laser irradiation, Ce6-generated ROS facilitated the transformation of N3-functionalized PR104A into ultrasmall nanoparticles with enhanced permeability. Fluorometric analysis confirmed 60% release of RhB-labeled PAMAM from crosslinked depots upon 15 min post-irradiation (Fig. 24B), demonstrating efficient photo-triggered decrosslinking. Negligible release in non-irradiated control groups validated the design feasibility. Furthermore, in vitro experiments showed hypoxia-activated PR104A release in 4T1 cells by HPLC quantification (Fig. 24C). Using 4T1 multicellular spheroids (MCSs) as an in vitro model, photoirradiation/ROS-mediated permeation of PAMAM-PR104A was verified by the observation of red fluorescence at 50 μm depth in the nanodepot incubated group treated in an acidic environment and with light irradiation (Fig. 24D), ultimately achieving synergistic PDT and chemotherapy within hypoxic tumors.
![]() | ||
Fig. 24 (A) Schematic illustration of the ultra-fast pH-responsive “two-step” size-transformation nanodepots achieving synergistic PDT and chemotherapy in hypoxic tumors. (B) Release of PAMAMRhB upon 660 nm laser exposure of crosslinked depots. (C) HPLC analysis of hypoxia-triggered PR104A prodrug release from 4T1 cells at varying incubation times. (D) Confocal fluorescence image of the MCSs incubated with TK-PAMAMRhB-N3 and Nce6-DBCO after different treatments at different depths, adapted with permission from ref. 175. Copyright 2022, Elsevier. |
The type of size transformation | Applications | |
---|---|---|
Small to large | From molecules to the dimer, the dimer self-assembles into nanoaggregates | FL, PET, PA, MRI, RI, QPI, multi modality, chemotherapy, and PDT |
From molecules to the dimer, the dimer self-assembles into nanotubes | FL | |
From molecules to the dimer | PET, RI, and RT | |
From nanoparticles to nanoaggregates | FL, chemotherapy, PTT, RT, and synergistic therapy | |
From nanoparticles to nanoaggregates and release drugs | PTT | |
Multi-size transformation | From nanoparticles to nanoaggregates, and then release smaller particles | Chemotherapy, CDT, and synergistic therapy |
From molecules into nanoaggregates and then reassemble into dimers | Synergistic therapy |
Firstly, it is necessary to modulate bioorthogonal responsive units that allows elevating tissue selectivity. Most reported bioorthogonal responsive nanosystems rely on the responsiveness between probes and overexpressed biomarkers, such as caspase-1, Cat B and GSH, to be transferred from the inert to activated state. However, these biomarkers do not have very closely relationship with specific tissues. For example, overexpressed caspase-1 can be found in S. aureus infected RAW264.7 cells and is associated with AD neuroinflammation. GSH exists in both normal and tumorous tissues. Hence, it is advisable to build multiple biomarker responsive platforms to amplify the selectivity of bioorthogonal reactions. Additionally, “photo-triggered” bioorthogonal reactions are another good candidate. For example, the light driven reaction between tetrazine and norbornene provides a powerful tool owing to temporospatial controllability. The unique “photo-triggered” reactive units can combine with other bioorthogonal reactions and allow building an “AND” logic gate that can simultaneously or sequentially response to multiple stimuli for precise targeted imaging and drug release. Despite the above-mentioned strategies being helpful to improve the bioorthogonal reactions’ selectivity, the off-target effects persist in the execution of bioorthogonal reactions in complex in vivo physiological environments. First, the bioorthogonal functional groups may undergo unintended size reactions with endogenous substances within biological systems, such as Cys, thereby some bioorthogonal macrocyclization reactions with relatively slow kinetics decrease the targeting capability. Another unresolved issue is that the complex in vivo physiological environments induce off-target effects in bioorthogonal reactions. The in vivo environments with a high dilution factor exhibit decreased target concentrations and affect the reaction kinetics.
Secondly, enriching bioorthogonal reactive pairs is essential. Some bioorthogonal reaction rates are relatively low (bioorthogonal-induced nanoparticle aggregation finishes in 24 or 48 h) leading to the clearance of original nanocarriers before tumoral accumulation happens. Therefore, more stable, biocompatible and rapid bioorthogonal reactions needed to be explored. In general, reactions with rapid kinetics (iEDDA > CuAAC > 1,2-aminothiol-CBT click reactions ≥ SPAAC > SPSAC) are preferrable for biological applications,176 but optimizing molecules with specific moieties facilitates to dramatically decrease the activation free energy and allows the reaction kinetics to be changed. For example, structural optimization of molecules significantly enhanced macrocyclization kinetics, achieving first-order reaction rates of 7.97 × 10−5 s−1 for CyNAP and (3.9 ± 0.1) × 10−3 s−1 for [18F-IR780-1], respectively.76,121 This accelerated reaction kinetics effectively prevents premature clearance of the original nanocarrier prior to targeted site accumulation. On the other hand, this process may be simplified in the future via the collaboration between artificial intelligence and organic chemistry. Then the optimization of the scheduling gap between the bioorthogonal reaction pairs, the administered dose, the site accumulation and in vivo clearance rates is still the main challenge needed to be resolved. Additionally, some derivatives of bioorthogonal reactive pairs can be introduced in size variable nanosystems to improve functionality. For example, dissociative bioorthogonal reactions should be considered to be integrated into nanocarriers. These dissociative bioorthogonal reactive motifs can be linked with prodrugs or protein activators for targeting theranostics. The green byproduct from bioorthogonal reactions can also participate in synergistic therapy. For example, the IEDDA click reaction between a thiocarbamate-functionalized TCO and a Tz can generate H2S in cells, which opens up a new direction for the combination of H2S gas therapy with other modalities.177
Finally, the bioorthogonal chemistry mediated size variable nanosystems commonly induce the nano-to-cluster morphology transformation. The cascaded utilization of “small-to-large” and “large-to-small” size transformation may be advantageous due to the good distribution, penetration and metabolism. The long-term biosafety of these developed nanoplatforms should also be taken into consideration to achieve clinical translation. Instead of the well-developed CuAAC, copper-free alternatives, such as IEDDA, with faster kinetics and lower toxicity are preferrable to design size transformation systems for in vivo applications. Notably, a bioorthogonal reaction-mediated prodrug activation strategy, SQ3370, has entered Phase II clinical trials for tumor therapy.178 This milestone has significantly heightened researcher interest in bioorthogonal reaction-mediated drug delivery systems.
All in all, this area is still in its infancy. We hope efficient and low-toxic bioorthogonal tools will aid in developing intelligent nanocarriers and speeding up the pace of nanocarrier innovations to clinical trials.
This journal is © The Royal Society of Chemistry 2025 |