Abstract
Bioorthogonal chemistry-mediated self-assembly holds great promise for dynamic molecular imaging in living organisms. However, existing approaches are limited to nanoaggregates with ‘always-on’ signals, suffering from high signal-to-background ratio (SBR) and compromised detection sensitivity. Herein we report a nitrile-aminothiol (NAT) bioorthogonal fluorogenic probe (CyNAP-SS-FK) for ultrasensitive diagnosis of orthotopic hepatocellular carcinoma. This probe comprises a nitrile-substituted hemicyanine scaffold with a cysteine tail dually locked with biomarker-responsive moieties. Upon dual cleavage by tumor-specific cathepsin B and biothiols, the 1,2-aminothiol residue is exposed and spontaneously reacts with nitrile group for in situ intramolecular macrocyclization, enabling near-infrared fluorescence (NIRF) turn-on as well as self-assembly. In living male mice, such ‘cleavage-click-assembly’ regimen allows for real-time and ultrasensitive detection of small cancerous lesions (~2 mm in diameter) with improved SBR (~5) and extended detection window (~36 h), outperforming conventional clinical assays. This study not only presents NAT click reaction-based fluorogenic probes but also highlights a generic dual-locked design of these probes.
Similar content being viewed by others
Introduction
Optical imaging enables noninvasive, real-time visualization and quantification of biomarkers at molecular levels in living subjects1,2,3,4,5. By integrating molecular probes with imaging instruments, fluorescence imaging has become an indispensable tool in preclinical research and clinical practice6,7,8,9. To enhance diagnostic sensitivity, activatable probes of which the fluorescence is switched on by specific stimuli have been developed to translate diseases-associated biomarkers into optical output10,11,12,13. To date, numerous activatable probes have been constructed for the detection of enzymes14,15,16,17, chemical mediators18,19,20,21, and physiological indexes in living animals22,23,24,25,26. Nevertheless, traditional activatable molecular probes encounter challenges associated with rapid diffusion and clearance from diseases sites, which lead to shortened imaging window, reduced signal-to-background ratios (SBRs) and eventually compromised diagnostic accuracy.
In vivo self-assembly involves the spontaneous organization of externally administered small molecules into nanostructures in response to pathological abnormalities, allowing them to accumulate in target tissues27,28,29. This approach simultaneously capitalizes on the pharmacokinetic advantages of small molecular probes for deep-tissue penetration and broad biodistribution, as well as the benefits of nanoparticles for extended local retention at disease sites30,31. In normal tissues, the inactive probes remain unassembled and are efficiently cleared out from living bodies30,32. Thereby, self-assembled probes lead to improved therapeutic outcomes and imaging performance with reduced systematic toxicity. However, due to the sophisticated and dynamic physiological environment in vivo, achieving highly controllable and biocompatible self-assembly of molecular probes in target tissues remains a significant challenge.
Bioorthogonal reactions including Staudinger ligation33,34, trans-cyclooctene/tetrazine cycloaddition35,36 and 2-cyano-benzothiazole (CBT)/cysteine (Cys) condensation37,38,39,40 (Fig. 1a) have been utilized to construct in vivo self-assembled systems for drug delivery, polymer synthesis, and bioimaging. Among them, CBT-Cys click reaction and its biomarker-instructed self-assembly design have given rise to a novel family of nanoprobes for imaging of pH levels, reactive oxygen species and nitrogen species, as well as enzymatic activities41,42,43,44,45,46. However, existing CBT-Cys probes are intrinsically fluorescent prior to forming nanoaggregates for signal enhancement (Fig. 1c), posing a high risk of “false-positive” output and suboptimal SBRs. In contrast, bioorthogonal fluorogenic probes which are natively non-fluorescent but turn on fluorescence exclusively upon specific click reactions provides a powerful solution to address these concerns (Fig. 1b). However, a suitable fluorogenic scaffold and the biomarker-instructed probe design are currently lacking in existing studies.
a, b Schematic of CBT-Cys click reaction and NAT click reaction based fluorogenic hemicyanine in this study. c, d Schematic illustration of the intramolecular cyclization and self-assembly mediated by CBT-Cys click reaction c and NAT click reaction d, the comparison between two strategies in terms of imaging contrast and SBR. e Molecular structures and sensing mechanism of the dual-locked activatable probe (CyNAP-SS-FK) with GSH/Cat B-triggered intramolecular cyclization and self-assembly for non-invasive NIRF in vivo imaging of orthotopic HCC.
Nitrile-aminothiol (NAT) click reaction is a well-known condensation reaction between nitriles and 1,2-aminothiols to form thiazolines, which occurs readily under physiological conditions with exceptional specificity, fast rates and high yields47,48,49. For example, the heteroaromatic nitriles, such as 2-cyanoisonicotinamide (CINA), 6-cyanonicotinic acid (CNA) as well as 2-cyano-6-hydroxyquinoline (CHQ) are capable of condensation with N-terminal cysteines with high reactivity and stability at biological environments, thus widely used for protein and peptide synthesis50,51. However, its utilization as fluorogenic self-assembled probes hasn’t been explored.
Herein we report a dual-locked NAT bioorthogonal fluorogenic probe (CyNAP-SS-FK) with biomarker-triggered near-infrared fluorescence (NIRF) turn-on and self-assembly for non-invasive diagnosis of hepatocellular carcinoma (HCC) (Fig. 1d, e). CyNAP-SS-FK comprises a NAT bioorthogonal fluorogenic scaffold (CyNAP) linked with an aminothiol tail that is jointly caged by glutathione (GSH)-responsive disulfide bond and a cathepsin B (Cat B)-cleavable peptide sequence (Ac-FK) (Fig. 1e). GSH and Cat B are biomarkers co-overexpressed by HCC cells and are highly associated with tumor progression, angiogenesis, and metastasis52. Prior to probe construction, the CyNAX scaffolds (nitrile-substituted hemicyanines) along with the length and rigidity of the spacer chain are carefully screened to optimize click reaction and intramolecular macrocyclization kinetics. CyNAP-SS-FK is intrinsically non-fluorescent due to the inhibited intramolecular charge transfer (ICT) by nitrile group. Upon removal of both caging groups by Cat B and GSH, the aminothiol head in CyNAP-SS-FK is exposed to undergo intramolecular cyclization with nitrile through NAT click reaction, which unlocks ICT for the recovery of NIRF emission, and simultaneously enables self-assembly to nanoaggregates. This particular ‘cleavage-click-assembly’ sensing mechanism by CyNAP-SS-FK allows for real-time ultrasensitive detection of small HCC (~2 mm in diameter) with improved SBR (~5) and prolonged detection window (~36 h), which outperforms the control ‘always on’ probe and clinical assays.
Results
Rational design of NAT bioorthogonal fluorogenic luminophores
Asymmetric hemicyanines with a donor-π-acceptor (D − π − A) motif are frequently used to construct activatable probes in response to various biomarkers17,53,54,55,56,57,58. Its sensing mechanism relies on adjusting the ICT between electron donating phenol group and the electron-withdrawing indole group on the hemicyanine scaffold. When the phenol group is caged with a biomarker-responsive moiety, the electron donating ability is significantly reduced, causing ICT blocking and fluorescence quenching. Upon the removal of caged moieties by specific targets, the probes are activated to ‘turn on’ fluorescence59,60,61. Given the importance of electron ‘push-pull’ effects in hemicyanine scaffold, we proposed that replacing the electron-donating hydroxyl group with a strong electron-withdrawing nitrile group could block ICT and ‘turn-off’ NIRF signal (Fig. 2a), which could then be restored when nitrile undergoes NAT click reaction to afford thiazoline, as the electron density is dramatically changed.
a Chemical structures and sensing mechanisms of nitrile-substituted hemicyanines with free cysteine via NAT click reaction. b Synthetic routes for CyN-1−5. Reagents and conditions: (i) 2-bromocyclohex-1-ene-1-carbaldehyde, Cs2CO3, DMF, rt, 12 h. (ii) 1,2,3,3-tetramethyl-3H-indolium iodide, AcONa, Ac2O, 60 °C, 2−4 h. (iii) vilsmeier haack reagent, EtOH, reflux, 4 h. (iv) 5-formyl-2,4-dihydroxybenzonitrile, K2CO3, DMF, 50 °C, 12 h. (v) Cs2CO3, CH3I, ACN, rt, overnight. c The second order rate constant (M−1 s−1) and fluorescence enhancement of CyN-1−5 in the presence of free cysteine. d Fluorescence spectra of CyN-1−5 (20 μM) in the absence or presence of Cysteine (100 μM) in PBS (containing 20% MeOH, pH = 7.4) at room temperature. The fluorescence excitation was set at 580 nm. The experiments in d were repeated independently three times with similar results. Source data are provided as a Source Data file.
To prove the proposed concept, we strategically designed and synthesized a series of nitrile-substituted hemicyanines (termed CyN-X in Fig. 2b and Supplementary Figs. 1, 2). Since the reactivity of the aromatic nitrile is affected by π electron density and steric properties, we synthesized several compounds with different nitrile groups at C-6 (CyN-1–2) and C-7 (CyN-3–5) positions to identify the optimal site for click reaction. As shown in Fig. 2b, the synthesis of CyN-1–3 involved two simple steps: firstly, each benzonitrile derivative was condensed with 2-bromocyclohex-1-ene-1-carbaldehyde in presence of Cs2CO3 as base to give compound 2; secondly, the resulted aldehyde was directly condensed with indole quaternary ammonium salt to yield the corresponding fluorophores (CyN-1–3). Differently, CyN-4–5 was obtained by treating a chloro-substituted cyanine with resorcin derivates in the presence of a base. Briefly, Vilsmeier Haack reagent was reacted with an indole quaternary ammonium salt in refluxing ethanol to obtain the chloro-substituted cyanine 4. Resorcin derivative as a nucleophile was substituted on the meso position, resulting in the formation of product CyN-4 via Retro-Knoevenagel reaction. Finally, methylation of CyN-4 by iodomethane yielded CyN-5.
To evaluate the reactivity of nitrile-substituted hemicyanines with 1,2-aminothiols, their optical profiles and the second-order rate constants were determined after the addition of free cysteine in phosphate-buffered saline (PBS) solution. For CyN-1 and CyN-3 with different nitrile group positions, the fluorescence signals were enhanced by 4.1- and 2.7-fold, respectively (Fig. 2c, d and Supplementary Fig. 3). This observation was consistent with previous studies suggesting that the C-6 position is preferable for the electron-donating group of hemicyanine to achieve optimal fluorescence activity62. Moreover, CyN-1 has a higher second-order rate constant of 0.309 M−1 s −1 compared to that of 0.221 M−1 s −1 for CyN-3 (Supplementary Fig. 4). The proposed mechanism of intermolecular cyclization involves an initial nucleophilic attack by the sulfhydryl group on the nitrile group to produce a thioimidate intermediate (Fig. 2a), followed by condensation reaction with amine to yield the 2-aminothiazolidine intermediate. Finally, ammonia is irreversibly eliminated and results in the formation of the thiazoline product. The detection of ESI-MS spectrum (497.34, 554.35 and 497.36 for CyNC-1−3, respectively) in Supplementary Fig. 5 confirmed the intermolecular cyclization.
While a reduction in electronegativity on the aromatic nitrile would theoretically accelerate the nitrile-aminothiol reaction, a comparison between CyN-1 and CyN-2 reveals that adding one halogen atom on the ortho position significantly decreases the reactivity of aromatic nitrile (0.151 M−1 s−1, Supplementary Fig. 4) and reduces fluorescence enhancement (~1.7-fold). This is likely due to the steric hindrance caused by chlorine atom, which retards the intramolecular cyclization. The reactivity of a hydroxyl-substituted (CyN-4) or an alkoxy-substituted aromatic nitrile (CyN-5) with cysteine was also tested. Since hydroxyl- and alkoxy-substituents offer electron-donating effects, their reactivities are extremely low, as evidenced by negligible fluorescence enhancement (Fig. 2c). Based on these structure-activity studies, CyN-1 stands out with the fastest reacting kinetics and the highest fluorescence enhancement. We herein selected CyN-1 as the optimal bioorthogonal fluorogenic luminophores for the subsequent studies.
Screening of intramolecular NAT bioorthogonal fluorogenic luminophores
To construct and screen intramolecular NAT reaction mediated fluorogenic luminophores, we synthesized luminophores CyNAX including CyNA4, CyNA7 and CyNAP using CyN-1 with a N-terminal cysteine residue attached to either an alkyl chain or a short PEG chain linked to the indole ring (Fig. 3a). To investigate how the length and rigidity of the chain affect intramolecular macrocyclization kinetics, we varied the lengths of the alkyl or PEG chains (Fig. 3a). The synthetic routes of CyNAX were shown in Fig. 3b and Supplementary Fig. 6. Briefly, the commercially available raw material potassium phthalimide as amine source was used to convert a serious of dibromo derivatives into the corresponding phthalimide alkylation species 6, which then reacted with the indole to obtain compound 7. The resulted indole derivate 7 was condensed with aldehyde 2a to give hemicyanine derivative 8, followed by deprotection in the presence of N2H4•H2O to yield the primary amine 9. This primary amine was then conjugated with a protected cysteine and then deprotected trityl group to afford the corresponding probe CyNAX. The chemical structures of the intermediates and final products were characterized by NMR spectroscopy and ESI-MS (see the Supplementary Information).
a Chemical structures and intramolecular condensation via NAT click reaction. b Synthetic routes for CyNAX. Reagents and conditions: (i) dibromo derivative, KI, Acetone, reflux, overnight. (ii) 2,3,3-trimethyl-3H-indole, ACN, reflux, 24 h. (iii) compound 2a, AcONa, Ac2O, 60 °C, 0.5 h. (iv) N2H4•H2O, DCM/MeOH, rt, 2 h. (v) Boc-Cys(Trt)-OH, HBTU, DIPEA, DCM, rt, 2 h. (vi) Et3SiH, TFA (20% in DCM), rt, overnight. c Molecular orbitals of CyNAX and their cyclized products CyNAX-C. The HOMO and LUMO energy levels are indicated. d Hole (red color) and electron (blue color) of optimized excited-state geometries of CyNAP and CyNAP-C calculated by TD-DFT method at B3LYP/6-311 G (d, p) level with Grimme’s DFT-D3 correction. e Representative confocal microscopy images of LM3 cells incubated with CyNAX (30 μM) for 0, 3, 6, 12 and 24 h, respectively. Blue and red fluorescence indicate a cell nucleus stained with DAPI and the signals from CyNAX, respectively. (Scale bar, 10 μm). f Fluorescence enhancement and condensation reaction rate constants of CyNAX (start materials, SM) in the presence of 1,2-aminothiols. Log D values of CyNAX and their cyclized products CyNAX-C (terminal materials, TM). g Quantification of fluorescence intensity in the panel e. (n = 3, mean ± s.d.). Two-tailed Student’s t test; Control versus probe-treated groups. NS, no statistically significant differences. The experiments in g were repeated independently three times with similar results. h Calculated energy levels of single and triplet states for distinct luminescence behaviors of the CyNAP and CyNAP-C molecules. Green arrows represent absorption, red arrows indicate emission, and grey wavy arrows depict nonradiative quenching processes. i Fluorescence spectra of CyNAX in PBS (containing 20% MeOH, pH = 7.4) at room temperature. The fluorescence excitation was set at 580 nm. The experiments in i was repeated independently three times with similar results. Source data are provided as a Source Data file.
To screen the optimal CyNAX for intramolecular condensation, their optical profiles and the first-order rate constant were determined in PBS and MeOH solution (Fig. 3f, i and Supplementary Fig. 7). At pH 7.4, both CyNA4 and CyNA7, which have alkyl chain substitutions, produced cyclized products, as confirmed by ~4-fold fluorescence enhancement at ~675 nm and the detection of a MALDI-TOF MS peak (536.24 and 578.29) (Supplementary Fig. 8), which corresponded to the mass of cyclized products (CyNA4-C and CyNA7-C). The first-order reaction rate constants were determined to be 2.91 × 10−5 s−1 for CyNA4 and 0.89 × 10−5 s−1 for CyNA7 by high-performance liquid chromatography (HPLC) (Supplementary Figs. 9, 10). In contrast, CyNAP exhibited a higher fluorescence enhancement (~5.1-fold) and faster kinetics (7.97 × 10−5 s−1), implying that the flexible PEG linker can promote intramolecular macrocyclization. To provide a theoretical basis for understanding the effects of intramolecular cyclization on their fluorescence enhancement, we calculated the lowest unoccupied molecular orbital (LUMO) and highest occupied molecular orbital (HOMO) energies as well as their spatial distributions in CyNAX and CyNAX-C using density functional theory (DFT) (Fig. 3c). For CyNAX, the LUMOs were distributed on hemicyanine scaffold while the HOMOs were found on the cysteine-containing aminothiol group. In contrast, for the cyclized products CyNAX-C, both LUMOs and HOMOs were uniformly distributed on hemicyanine-thiazoline backbone, validating the effectiveness of the intramolecular cyclization design in achieving fluorescence ‘Turn-On’. To elucidate the underlying causes of the marked differences in fluorescence between CyNAP and CyNAP-C, we embarked on a computational analysis of their emission mechanisms via time-dependent density functional theory (TD-DFT) at the B3LYP/6-311 G (d, p) level with Grimme’s DFT-D3 correction. For CyNAP, as shown in Fig. 3d the separation between the hole and electron distributions is large, resulting in low electron transport efficiency. Additionally, the predicted optical gap for CyNAP is 2.20 eV (f = 0.09), which can be ascribed to the S0 − S2 transition, rather than the S0 − S1 transition with a negligible oscillator strength (f = 0.00) (Fig. 3h). In contrast, in CyNAP-C, the hole and electron distributions are closer together, facilitating better electron transport. Furthermore, CyNAP-C exhibits a clear S0 − S1 absorption at 2.20 eV (f = 0.03), making it more likely to emit fluorescence from the S1 state.
To evaluate the optical performance of probes CyNAX in cells, confocal microscopy was performed on CyNAX-treated human HCC-LM3 cells. When LM3 cells were incubated with CyNAX, gradually enhanced fluorescence was observed (Fig. 3e). CyNA4 and CyNA7 showed slight fluorescence enhancement over time. In contrast, CyNAP turned on its fluorescence more rapidly. Fluorescence signal of CyNAP in LM3 cells was 3.4-fold and 3.6-fold higher than that of CyNA4 and CyNA7, respectively (Fig. 3g), at 3 h post-incubation. This could be attributable to CyNAP’s faster cyclization rate.
‘Click-Assembly’ and specificity of intramolecular condensation
Given that the endogenous free cysteine in cells and plasma ranges from 20 to 100 μM, the specificity of intramolecular condensation was studied in the presence of high concentration of free cysteine. This was done because locally free cysteine may also undergo intermolecular condensation with aromatic nitrile (Fig. 4a and Supplementary Fig. 11). To investigate this, CyNAP was co-incubated with varying mole ratios of free cysteine, and the condensation products were traced by HPLC. As shown in Fig. 4b, incubating CyNAP and cystine at a 1:1 molar ratio in PBS resulted in the rapid conversion of ~100% CyNAP to CyNAP-C, as confirmed by the detection of an HPLC peak with a retention time of about 20.6 min, corresponding to the intramolecular cyclized products (CyNAP-C). This high specificity was also observed at a 1:100 and even 1:1000 molar ratios of CyNAP to cystine, demonstrating that intramolecular condensation occurs preferentially over intermolecular condensation. This was ascribed to the first-order reaction rate constants (7.97 × 10−5 s−1) of CyNAP enables promoting intramolecular cyclization preferentially, consistent with previous studies42. To evaluate whether the intramolecular condensation can induce the formation of nanoaggregates, we used transmission electron microscopy (TEM) and dynamic light scattering (DLS) assays to analyze NAT click reaction mediated self-assembly of CyNAP. As shown in Fig. 4c, CyNAP was converted to the cyclized product CyNAP-C in PBS solution (pH 7.4), which spontaneously assembled into nanoparticles with a spherical morphology. Furthermore, the average hydrodynamic diameter for CyNAP-C was measured to be ~100 nm (Fig. 4d). The self-assembly was attributed to the increased rigidity and high hydrophobicity of CyNAP-C, reflected by a significantly higher Log D value (3.33) relative to CyNAP (Log D = 2.88; Fig. 3f).
a Competition studies between intramolecular and intermolecular condensation. b A competitive HPLC assay of CyNAP incubated with different molar ratio of free cysteine. c Representative TEM images of CyNAP-SS before and after incubation with GSH in buffer solution (scale bar, 200 nm). d Average diameters of CyNAP-SS before and after incubation with GSH in buffer solution by DLS. e Chemical structures and condensation mechanisms of CyNAP-SS and CyNAP-S in the response of GSH. f Representative confocal microscopy images of LM3 cells incubated with CyNAP-SS (20 μM) and CyNAP-S (20 μM) at 37 °C for 8 h in the presence or absence of the GSH synthesis inhibitor BSO. NIRF images acquired at excitation at 594 nm. g Quantification of fluorescence intensity in the panel f. Levels of fluorescence intensity at different groups (n = 5, mean ± s.d.). (Scale bar, 5 μm). Two-tailed Student’s t test; Control versus probe-treated groups. NS, no statistically significant differences. The experiments in b–d and f were repeated independently three times with similar results. Source data are provided as a Source Data file.
To further investigate the specificity of CyNAP self-cyclization in cells, we synthesized a GSH-responsive probe CyNAP-SS containing disulfide bond and a control probe CyNAP-S containing a single sulfide bond insensitive to GSH (Fig. 4e and Supplementary Figs. 12, 13). We hypothesized that the disulfide bond of CyNAP-SS would be reduced by cellular GSH to expose 1,2-aminothiols residues, which would then react with nitrile groups to form fluorescent cyclized products and self-assemble into nanoparticles. In contrast, CyNAP-S would not undergo self-cyclization as it is unable to be reduced by GSH (Fig. 4e). To test this, CyNAP-SS or CyNAP-S was respectively incubated with LM3 cells for 8 h before confocal fluorescence imaging. As shown in Fig. 4f, NIRF signal of CyNAP-SS was clearly observed in the cytoplasm of LM3 cells, indicating the formation of fluorescent cyclized products. However, in LM3 cells pre-treated with a GSH synthesis inhibitor buthionine sulfoximine (BSO)63, NIRF signals from CyNAP-SS decreased to basal levels. Notably, almost no fluorescence was observed in cells incubated with the control probe CyNAP-S, which was 2.5-fold lower than that of CyNAP-SS (Fig. 4g and Supplementary Fig. 14). These results confirmed that CyNAP had a high specificity for intramolecular condensation, even in the presence of endogenous free cysteine in tumor cells.
Design and characterization of dual-locked fluorogenic probes
Encouraged by the robustness of intramolecular cyclization and self-assembly of CyNAP, we developed a dual-locked activatable probe (CyNAP-SS-FK) for NIRF imaging and companion diagnosis of orthotopic HCC in living mice. CyNAP-SS-FK comprises three key units: a NAT bioorthogonal fluorogenic chromophore (CyNAP), a GSH-responsive disulfide bond and a cathepsin B (Cat B)-reactive peptide moiety (Ac-FK) (Fig. 5b). GSH and Cat B are biomarkers co-overexpressed by HCC cells, promoting tumor growth, angiogenesis, and metastasis56,64. CyNAP-SS-FK is intrinsically non-fluorescent and only activate its NIRF signals in the presence of both GSH and Cat B. This ‘dual-locked’ design ensures that the NIRF signal activation is specific to tumor cells but not normal tissues. To rule out ‘cleavage-click-assembly’ induced long-term retention effect in tumor imaging, we further synthesized CyNAP-T as an ‘always on’ probe for comparison (Supplementary Fig. 15). CyNAP-T has a similar chemical structure to the cyclized product CyNAP-C, but differs in that the macrocyclic ring is open. The synthetic route of CyNAP-SS-FK is shown in Fig. 5a and Supplementary Fig. 16. First, nitrile-substituted hemicyanine 9c was synthesized and then coupled with the protected cysteine derivate to give CyNAP-SS, which was further conjugated with the protected Cat B substrate (Acetyl-Phe–Lys-OtBu) to obtain CyNAP-SS-FK. All the intermediates and probes were characterized by 1H NMR spectra, LC-MS spectra or High-resolution mass spectrum.
a The synthesis route of CyNAP-SS-FK. b Activation mechanism of the dual-locked molecular probe CyNAP-SS-FK for LM3 cell detection. c Fluorescence spectra of CyNAP-SS-FK (20 µM) after incubation with/without GSH, Cat B, or GSH and Cat B at 37 °C in PBS buffer (pH 7.4). Fluorescence excitation at 580 nm. d Fluorescence intensity of CyNAP-SS-FK (40 µM) after incubation with the indicated enzymes, metal ions (60 µM), ROS (60 µM) in PBS (10 mM, pH 7.4) at 37 °C. Mg2+, magnesium chloride; Ca2+, calcium chloride; ClO-, sodium hypochlorite; H2O2, hydrogen peroxide, SO42-, sodium sulfate; GSH, glutathione; TCEP, Tris(2-carboxyethyl) phosphine hydrochloride; Cas-3, caspase 3; Galactosidase; Cat B, Cathepsin B. (n = 3, mean ± s.d.). Two-tailed Student’s t test; GSH and Cat B treated group versus other groups. NS, no statistically significant differences. e Mean NIRF intensities of HCC-LM3 after incubation with CyNAP-SS-FK (with or without BSO) in the panel h (n = 4, mean ± s.d.). n refers to cell imaging experiment repeat times. (Scale bar, 5 μm). Two-tailed Student’s t test; Control versus different incubation time. NS, no statistically significant differences. f Representative TEM image and DLS result of CyNAP-SS-FK before and after incubation with GSH and Cat B in buffer solution (scale bar, 150 nm). g Schematic illustration of the sensing mechanism of CyNAP-SS-FK in living cells. h Representative confocal microscopy images of LM3 cells incubated with CyNAP-SS-FK (25 μM) for 0, 4, 8 and 12 h (with or without BSO), respectively. Blue and red fluorescence indicate a cell nucleus stained with DAPI and the signals from CyNAP-SS-FK, respectively. NIRF images acquired at excitation at 594 nm. The experiments in c, d, f and h were repeated independently three times with similar results. Source data are provided as a Source Data file.
To investigate the responses of CyNAP-SS-FK towards GSH and Cat B, the optical properties were characterized in the absence or presence of these biomarkers. CyNAP-SS-FK initially showed one absorption peak at 550 nm (Supplementary Fig. 17 and was nonfluorescent with a low fluorescent quantum yield of about 0.8%. After incubation with GSH or Cat B alone, CyNAP-SS-FK remained nonfluorescent as a single cleavage by GSH or Cat B was insufficient to induce NAT reaction-mediated intramolecular cyclization (Fig. 5c). However, when incubated with both GSH and Cat B, the NIR fluorescence of CyNAP-SS-FK was turned on, showing a ~5-fold increase at 680 nm with an increased fluorescence quantum yield of ~4% (Fig. 5c). DLS and TEM analyses of CyNAP-SS-FK following incubation with GSH and Cat B further confirmed the intramolecular cyclization and in-situ formation of nanoaggregates with an average size of around 100 nm (Fig. 5f). Moreover, no significant NIRF change was observed in the presence of other analytes including reactive oxygen species, reductant, and enzymes, confirming its high stability and specificity (Fig. 5d). Given its minimal cytotoxicity (Supplementary Fig. 18), the capability of CyNAP-SS-FK to detect HCC LM3 was investigated. When LM3 cells were incubated with CyNAP-SS-FK, red fluorescence gradually increased with the incubation time (Fig. 5h). In contrast, minimal NIRF signals were observed in LM3 cells pre-treated with the GSH synthesis inhibitor BSO. Quantification results showed that NIRF signal in LM3 cells was 3.9-fold higher than in cells pretreated with BSO. These results confirmed that the NIRF signals of CyNAP-SS-FK can be specifically activated by endogenous GSH and Cat B in LM3 cells. The stability of CyNAP-SS-FK in circulation after injection was also confirmed (Supplementary Fig. 19).
Real-time and in vivo NIRF imaging of HCC
Early diagnosis of cancer is crucial for the treatment and accurate prognosis assessment. However, conventional diagnostic methods often struggle with low specificities and sensitivities, particularly for tumors smaller than 2 mm. The ability of CyNAP-SS-FK for real-time NIRF imaging of HCC was evaluated in a murine orthotopic HCC model, established according to previous studies60. At different time points post-tumor implantation (0, 5, 10, and 15 days), CyNAP-SS-FK or the control probe CyNAP-T was intravenously injected into living mice, followed by longitudinal NIRF imaging (Fig. 6a, b). NIRF imaging failed to detect the liver in control mice (0 day) throughout the entire imaging course (Fig. 6c). However, at 5 days post tumor implantation when tumor diameter was ~2 mm (Supplementary Fig. 20), the liver was clearly delineated, suggesting the activation of CyNAP-SS-FK by upregulated Cat B and GSH during tumor growth. The NIRF signals in the tumor regions increased gradually, peaking at 12 h post-injection of CyNAP-SS-FK and showing a 1.9-fold signal increase compared to those from healthy control mice. Similar signal patterns were observed at 10- and 15-day post tumor implantation, with maximum tumor signals 2.7 and 3.6 times higher than those from control mice (Fig. 6d), indicating further elevated Cat B and GSH levels.
a Chemical structure of CyNAP-SS-FK and the control probe CyNAP-T. b Schematic illustration of the development of orthotopic liver tumor model and NIRF imaging at different post tumor implantation time points. c Representative NIRF images of living mice after injection of CyNAP-SS-FK or CyNAP-T at different post tumor implantation timepoints. The white circles indicate the liver (Li) or bladder (Bl). NIRF images acquired at 710 nm upon excitation at 660 nm. d The dynamic NIRF intensities of liver in living mice as a function of time after injection of CyNAP-SS-FK or CyNAP-T for different groups (n = 3, mean ± s.d.). Two-tailed Student’s t test; Day-0 group versus different tumor incubation time groups. NS, no statistically significant differences. e, f Representative confocal fluorescence microscopy images of regional liver slices and photomicrographs of haematoxylin and eosin staining in paraffin-embedded liver sections from different mice groups. The blue, green and red signals originate from 4,6-diamidino-2-phenylindole (DAPI), Cat B antibody staining and CyNAP-SS-FK, respectively (scale bar, 200 µm). NIRF images acquired at excitation at 594 nm. Each experiment was repeated at least for 3 times. g The IVIS imaging images of the organs of mice after injection of CyNAP-SS-FK about 12 h. Data are the mean ± SD. n = 3 independent experiments. h Quantified mean fluorescence intensity (MFI) in Fig. 6c (n = 3, mean ± s.d.). Two-tailed Student’s t test; Day-0 group versus different tumor incubation time groups. *p < 0.05, **p < 0.01, ***p < 0.001, NS, no statistically significant differences. i-j Measurements of liver function in the mouse model of orthotopic liver cancer. Levels of ALT and AST at different groups (n = 5, mean ± s.d.). Two-tailed Student’s t test; Day-0 group versus different tumor incubation time groups. NS, no statistically significant differences. k Quantified SBRs of CyNAP-SS-FK or CyNAP-T at 12 h post-injection time for different groups. Levels of SBRs at different groups (n = 3, mean ± s.d.). l Schematic illustration of detection timeline of HCC comparing CyNAP-SS-FK to the clinical and preclinical assays. m Comparison of the retention effect and SBRs of CyNAP-SS-FK or CyNAP-T in entire tumor imaging window period. Source data are provided as a Source Data file.
Tumor imaging ability of control probe CyNAP-T was also studied. After 12 h post-injection, NIR signals of CyNAP-T in the liver were 2.6-fold lower than those of CyNAP-SS-FK. Similarly, ex vivo imaging showed that with prolonged tumor burden, the NIFR signal increased accordingly (Fig. 6g). The immunofluorescence studies also showed the NIRF signal of CyNAP-SS-FK in the liver was ∼2-fold higher than that of CyNAP-T (Fig. 6e, h and Supplementary Fig. 21). The SBR of CyNAP-SS-FK was ~5-fold higher than that of the ‘always-on’ probe CyNAP-T, owing to its active sensing mechanisms (Fig. 6k). Furthermore, at 15-day post tumor implantation, the NIRF signal of mice injected with CyNAP-SS-FK remained high for over 36 h, showing prolonged tumor retention and an extended imaging window (Fig. 6d, m). In contrast, the NIRF signal from CyNAP-T rapidly declined and became very weak at 6 h post-injection (Fig. 6d, m), attributed to its rapid diffusion and clearance in vivo. These findings suggested that the intramolecular cyclization and self-assembly of CyNAP-SS-FK through NAT bioorthogonal reaction enable an enhanced permeability and retention (EPR) effect, thereby allowing prolonged and high-contrast tumor imaging in vivo.
The measurement of serum ALT and AST is routinely used for of liver function assessment in clinics. We thus compared the detection ability of CyNAP-SS-FK with serum ALT/AST assays in our orthotopic HCC model. A statistically significant increase in ALT and AST was observed at 15-day post tumor implantation when tumor diameter reaches up to 6.5 mm (Fig. 6i, j), which was 3.3- and 3.9-fold higher than control groups, respectively. We also conducted measurements of kidney function in the mouse model to verify CyNAP-SS-FK could not impair the kidney function (Supplementary Fig. 22). Histological studies showed a small tumor area and a noticeable tumor area were observed at 5- and 10-day post tumor implantation, respectively (Fig. 6f), and no obvious changes of other main organs were found. (Supplementary Fig. 23). The first statistically significant change in CyNAP-SS-FK-NIRF signals was detected at least 10 days earlier than with ALT/AST assays and 5 days earlier than histological assays (Fig. 6l). These findings demonstrate that CyNAP-SS-FK not only enables early and sensitive detection of HCC, but also provides improved imaging contrast and a prolonged detection window.
Discussion
A substantial challenge facing diagnosis of diseases via optical imaging is a lack of fluorogenic probes that possess fast intramolecular condensation in vivo to prolong the imaging window periods and enhance SBRs. In situ self-assembly strategy via biorthogonal chemistry allow probes to have longer residence time in the disease site but faster excretion of unwanted small molecules from other organs. Unlike other bioorthogonal reactions such as copper-catalyzed azide-alkyne cycloaddition and palladium-mediated Alloc/Poc cleavage that required metal catalysts, NAT click reaction is readily proceed under physiological conditions without the need of any added catalyst and has been shown to be exquisitely specificity, fast reaction rates and high-yielding. Although CBT-Cys click reaction (analogical to NAT) mediated self-assembly has widely reported, those CBT-Cys probes and their resulting nanoaggregates have ‘always-on’ imaging signals and lack fluorogenicity, giving a relatively low SBR.
To tackle those challenges, we developed NAT bioorthogonal fluorogenic probes for real-time and noninvasive in vivo NIRF imaging. To identify the suitable fluorogenic luminophores, we synthesized a series of nitrile-substituted hemicyanines and screened for the optimal structure, CyN-1, which demonstrated the fastest click reaction rate and best fluorogenicity. We then systematically analyzed the structure–activity relationship of aromatic nitrile in CyN-1 and the aminothiol tail linked through varying spacers, identifying CyNAP with the highest intramolecular cyclization kinetics and the topmost fluorescence recovery. As the reaction rate constant of CyNAP’s intramolecular cyclization is orders of magnitude higher than that of its intermolecular reaction with free cysteine, intramolecular condensation preferentially occurs and results in spontaneous self-assembly into nanoaggregates, rather than undergo intermolecular condensation. We further developed a dual-locked activatable probe (CyNAP-SS-FK) by caging the aminothiol tail of CyNAP with both GSH-responsive moiety and Cat B-cleavage peptide. In HCC tumor cells with aberrantly high levels of GSH and Cat B, CyNAP-SS-FK was specifically activated to initiate intramolecular cyclization, igniting NIR fluorescence and simultaneously self-assembling into nanoparticles (~100 nm) for enhanced SBR and prolonged probe retention. Such ‘cleavage-click-assembly’ sensing mechanisms allowed CyNAP-SS-FK to real-timely and noninvasively imaging orthotopic HCC in living mice with improved SBR (~5, Supplementary Table 1) and prolonged imaging window period, which was unattainable for the control ‘always-on’ and non-cyclization probe CyNAP-T or the clinical approach the serum ALT/AST assay.
In summary, this study presents a bioorthogonal NIR fluorogenic probe utilizing NAT click reaction for in vivo imaging of HCC. Compared to broadly studied CBT-Cys probes with ‘always-on’ signals, NAT fluorogenic probes with ‘cleavage-click-assembly’ sensing mechanism offers enhanced imaging contrast and prolonged detection window. By simply varying biomarker-responsive moieties, this molecular design can be employed to create a versatile toolkit for ultrasensitive early diagnosis of various diseases beyond HCC.
Methods
General information
1H NMR and 13C NMR spectra were recorded on a Bruker AVANCE III 400 MHz (or 500 MHz) spectrometer (400 MHz or 500 MHz for 1H NMR, 100 MHz or 125 MHz for 13C NMR). Data were reported as follows: chemical shift, integration, multiplicity (s = singlet, d = doublet, t = triplet, q = quartet, m = multiplet, br = broad), coupling constants (Hz), and assignment. Low-resolution mass spectrometry (LRMS) data were recorded on Waters SQD2 with Waters H-Class UPLC. High-resolution mass spectrometry (HRMS) data were recorded on Waters with UPLC quattro premier XE. MALDI-MS spectrometric analyses were recorded at the Bruker ultraflex TOF/TOF instrument. Preparative HPLC was performed on a Shimadzu LC-16P system equipped with an Ascentis® C18 HPLC Column (25 cm × 21.2 mm, 5 μm) using gradient elution. The analytical HPLC was performed on a dual pump Shimadzu LC-16 system equipped with an Ascentis® C18 HPLC column (4.6 × 250 mm, 5 μm) using methanol/water (containing 0.1% Formic acid) as the mobile phase at a flow rate of 1 mL/min. DLS (Dynamic light scattering) experiment was operated on the Dynapro Plate Reader II.
Reactions were monitored by thin-layer chromatography (TLC, silica gel HSGF 254, 0.25 ± 0.02 mm). A silicone oil bath was used for experiments requiring heating. Reaction products were purified by column chromatography on silica gel (Qingdao Haiyang Chemical, zcx-II, 200−300 or 300−400 mesh). Chemical reagents and solvents were purchased from Aladdin Chemicals, Innochem, Energy Chemical and Bidepharm and used as received.
General procedure for the synthesis of 2a–c
A solution of 1 (0.1 mmol, 1.0 equiv.), 2-bromocyclohex-1-enecarbaldehyde (1.2 equiv.) and Cs2CO3 (3.0 equiv.) in DMF (2 mL) was stirred at room temperature overnight65. The reaction mixture was quenched with H2O and extracted with EA for three times. The combined organic layer was dried over Na2SO4 and concentrated under reduced pressure to afford the 2, which was used without further purification for the next step.
The procedure for the synthesis of 1,2,3,3-tetramethyl-3H-indol-1-ium iodide
A solution of 2,3,3-Trimethylindolenine (10.0 mmol, 1.59 g) and CH3I (50.0 mmol, 7.1 g) was stirred at 50 °C for 4 h53. After being cooled to room temperature, the reaction mixture was washed with Et2O for three times and the resulting residue was dried under reduced pressure to afford 1,2,3,3-tetramethyl-3H-indol-1-ium iodide (2,71 g, 9.0 mmol, 90% yield) as pink solid. 1H NMR (400 MHz, DMSO-d6) δ 7.92–7.90 (m, 1H), 7.84–7.82 (m, 1H), 7.70–7.54 (m, 2H), 3.98 (s, 3H), 2.77 (s, 3H), 1.53 (s, 6H).
General procedure for the synthesis of CyN-(1-3)
A solution of 2 (0.2 mmol, 1.0 equiv.), 1,2,3,3-tetramethyl-3H-indol-1-ium iodide (90.5 mg, 1.5 equiv.), AcONa (65.6 mg, 4.0 equiv.) in Ac2O (4 mL) was stirred at 60 °C for 2 h. After being cooled to room temperature, the reaction mixture was quenched with H2O, and extracted with DCM for three times. The combined organic layer was dried over Na2SO4, the filtrated was concentrated under reduced pressure. The purification using column chromatography on silica gel (eluted with DCM/MeOH = 20/1) was performed to afford the CyN-X.
CyN-1
Purplish red solid, 51.9 mg, 66% yield. 1H NMR (400 MHz, Chloroform-d): δ 8.60 (d, J = 15.2 Hz, 1H), 7.55–7.50 (m, 4H), 7.45–7.43 (m, 2H), 7.39–7.37 (m, 1H), 7.04 (d, J = 15.2 Hz, 1H), 6.93 (s, 1H), 4.27 (s, 3H), 2.94 (t, J = 6.0 Hz, 2H), 2.73 (t, J = 6.4 Hz, 2H), 2.00–1.94 (m, 2H), 1.84 (s, 6H). 13C NMR (125 MHz, Chloroform-d): δ 180.0, 156.5, 152.5, 146.2, 142.5, 142.0, 134.7, 129.7, 128.9, 128.3, 127.9, 127.7, 126.1, 125.8, 122.5, 119.0, 117.8, 117.1, 114.0, 113.3, 51.5, 35.9, 29.9, 29.8, 28.0, 25.0, 20.1. LRMS (m/z): [M]+ calcd. for C27H25N2O+ 393.20, found 393.37.
CyN-2
Purplish red solid, 48.7 mg, 57% yield. 1H NMR (500 MHz, Chloroform-d) δ 8.55 (d, J = 15.5 Hz, 1H), 7.56–7.51 (m, 4H), 7.42 (s, 1H), 7.37 (s, 1H), 7.09 (d, J = 15.5 Hz, 1H), 6.81 (s, 1H), 4.30 (s, 3H), 2.95 (t, J = 6.0 Hz, 2H), 2.72 (t, J = 6.0 Hz, 2H), 1.97–1.93 (m, 3H), 1.84 (s, 6H). LRMS (m/z): [M]+ calcd. for C27H24ClN2O+ 427.16, found 427.44.
CyN-3
Purplish red solid, 55.0 mg, 70% yield. 1H NMR (500 MHz, Chloroform-d): δ 8.60 (d, J = 15.5 Hz, 1H), 7.65 (dd, J = 8.5, 2.0 Hz, 1H), 7.60 (d, J = 2.0 Hz, 1H), 7.54–7.46 (m, 4H), 7.36 (d, J = 8.5 Hz, 1H), 6.97 (s, 1H), 6.88 (d, J = 15.5 Hz, 1H), 4.22 (s, 3H), 2.86 (t, J = 6.0 Hz, 2H), 2.71 (t, J = 6.0 Hz, 2H), 1.95–1.90 (m, 2H), 1.81 (s, 6H). 13C NMR (125 MHz, Chloroform-d): δ 179.9, 157.7, 154.9, 146.3, 142.3, 141.9, 134.2, 132.9, 131.1, 129.6, 1288, 127.8, 122.7, 122.6, 117.9, 117.1, 116.6, 114.0, 109.2, 108.6, 51.5, 30.2, 29.8, 29.6, 27.9, 24.8, 20.0. LRMS (m/z): [M]+ calcd. for C27H25N2O+ 393.20, found 393.37.
4
The 4 was preformed according to the reported literature66. A solution of 1,2,3,3-tetramethyl-3H-indol-1-ium iodide (602.2 mg, 2.0 mmol), N-(((E)-2-chloro-3-((phenylamino)methylene) cyclohex-1-enyl)methylene)benzenemin-ium (359.3 mg, 0.5 equiv.), NaOAc (166.0 mg, 1.0 equiv.) and EtOH (30 mL) was refluxed for 4 h. After being cooled to room temperature, the reaction mixture was concentrated under reduced pressure to remove EtOH, the resultant reside was quenched with H2O, and extracted with DCM for three times. The combined organic layer was dried over Na2SO4, the filtrates was concentrated under reduced pressure. The purification using column chromatography on silica gel (eluted with DCM/MeOH = 10/1) was performed to afford the 4 (299.6 mg, 62% yield) as green solid. 1H NMR (400 MHz, Chloroform-d): δ 8.36 (d, J = 14.0 Hz, 2H), 7.43–7.38 (m, 4H), 7.29–7.20 (m, 4H), 6.22 (d, J = 14.0 Hz, 2H), 3.76 (s, 6H), 2.75 (t, J = 6.4 Hz, 4H), 2.01–1.96 (m, 2H), 1.73 (s, 12H). 13C NMR (125 MHz, Chloroform-d): δ 172.9, 150.7, 144.4, 142.8, 141.0, 128.9, 127.8, 125.4, 122.2, 110.9, 101.7, 49.3, 32.7, 28.1, 26.8, 20.7.
CyN-4
The CyN-4 was preformed according to the reported literature67. A solution of 4 (30.0 mg, 0.05 mmol), 5-formyl-2,4-dihydroxybenzonitrile (9.8 mg, 1.2 equiv.), K2CO3 (13.8 mg, 2.0 equiv.) and DMF (4 mL) was stirred at 50 °C for 12 h. After being cooled to room temperature, the mixture reaction was quenched with H2O and extracted with DCM for three times. The combined organic layer was dried over Na2SO4, the filtrates was concentrated under reduced pressure. The purification using column chromatography on silica gel (eluted with DCM/MeOH = 10/1) was performed to afford the CyN-4 (9.6 mg, 47% yield) as blue solid. 1H NMR (400 MHz, Chloroform-d): δ 8.40 (d, J = 14.4 Hz, 1H), 7.46 (s, 1H), 7.42–7.35 (m, 2H), 7.15–7.11 (m, 2H), 6.78 (s, 1H), 6.03 (d, J = 14.4 Hz, 1H), 3.64 (s, 3H), 2.64 (t, J = 5.2 Hz, 2H), 2.54 (t, J = 5.2 Hz, 2H), 1.83–1.80 (m, 2H), 1.69 (s, 6H). LRMS (m/z): [M]+ calcd. for C27H25N2O2+ 409.19, found 409.60.
CyN-5
To a solution of CyN-4 (30.0 mg, 0.06 mmol), K2CO3 (16.5 mg, 2.0 equiv.) in MeCN (2 mL) was added dropwise CH3I (25.5 mg, 3.0 equiv.) at room temperature, and then the mixture reaction was stirred overnight. The volatiles were removed under reduce pressure, the resultant residue was quenched with H2O, and extracted with DCM for three times. The combined organic layer was dried over Na2SO4, filtered and concentrated under reduced pressure. The purification using column chromatography on silica gel (eluted with DCM/MeOH = 10/1) was performed to afford the CyN-5 (20.8 mg, 82% yield) as purplish red solid. 1H NMR (400 MHz, Chloroform-d) δ 8.74 (d, J = 15.6 Hz, 1H), 7.72–7.49 (m, 1H), 7.60 (d, J = 8.8 Hz, 1H), 7.55–7.51 (m, 3H), 7.18 (s, 1H), 7.11 (J = 15.6 Hz, 1H), 6.74 (d, J = 8.8 Hz, 1H), 4.28 (s, 3H), 4.00 (s, 3H), 2.93 (t, J = 6.0 Hz, 2H), 2.71 (t, J = 6.0 Hz, 2H), 1.97–1.94 (s, 2H), 1.83 (s, 6H). LRMS (m/z): [M]+ calcd. for C28H27N2O2+ 423.21, found 423.55.
General procedure for the synthesis of 6a–c
A solution of potassium phthalimide (5, 1.85 g,10 mmol), dibromide derivative (10.0 equiv.), KI (83.0 mg, 5 mol%) and Acetone (30 mL) was refluxed overnight68. After being cooled to room temperature, the reaction mixture was filtrated, washed with acetone for three time and the filtrate was concentrated under reduced pressure to afford the 6.
6a (Linker = (CH2)4)
White solid, 2.37 g, 84% yield. 1H NMR (400 MHz, CD3OD): δ 7.91−7.88 (m, 2H), 7.86−7.83 (m, 2H), 3.75 (t, J = 6.8 Hz, 2H), 3.54 (t, J = 6.8 Hz, 2H), 1.97−1.84 (m, 4H). 13C NMR (125 MHz, Chloroform-d): δ 168.5, 134.1, 132.1, 123.4, 37.0, 32.9, 29.9, 27.3;
6b (Linker = (CH2)7)
White solid, 2.37 g, 73% yield. 1H NMR (400 MHz, Chloroform-d): δ 7.82−7.78 (m, 2H), 7.70−7.66 (m, 2H), 3.64 (t, J = 7.2 Hz, 2H), 3.35 (t, J = 6.8 Hz, 2H), 1.84−1.77 (m, 2H), 1.69−1.62 (m, 2H), 1.42−1.31 (m, 6H);
6c (Linker = CH2CH2(OCH2CH2)2)
White solid, 1.98 mg, 58% yield. 1H NMR (400 MHz, Chloroform-d) δ 7.86−7.84 (m, 2H), 7.73−7.71 (m, 2H), 3.91 (t, J = 5.8 Hz, 2H), 3.75 (t, J = 6.0 Hz, 2H), 3.74 (t, J = 6.4 Hz, 2H), 3.66−3.61 (m, 4H), 3.38 (t, J = 6.2 Hz, 2H). 13C NMR (125 MHz, Chloroform-d) δ 168.4, 134.5, 134.1, 132.3, 123.8, 123.4, 71.3, 70.6, 70.3, 68.12, 37.4, 30.4.
General procedure for the synthesis of 7a-c
A solution of 6 (1.0 equiv., 2.0 mmol) and 2,3,3-trimethyl-3H-indole (636.9 mg, 2 equiv.) in MeCN (10 mL) was refluxed for three days. After being cooled to room temperature, the reaction mixture was concentrated under reduced pressure. The resulting reside was diluted with MeOH, and then the mixture was added to Et2O (20 mL) slowly, the precipitates was observed, filtrated and dried under vacuums to afford the 7, which was used without further purification for the next step.
7a (Linker = (CH2)4)
White solid, 643.4 mg, 89% yield. 1H NMR (400 MHz, Chloroform-d) δ 7.85–7.80 (m, 2H), 7.73–7.69 (m, 2H), 7.64–7.54 (m, 4H), 4.69 (t, J = 7.6 Hz, 2H), 3.66 (t, J = 6.8 Hz, 2H), 3.12 (s, 3H), 1.97–1.89 (m, 2H), 1.70–1.63 (m, 2H), 1.66 (s, 6H), 1.60 (s, 3H), 1.49–1.45 (m, 4H), 1.36–1.34 (m, 2H). 13C NMR (125 MHz, Chloroform-d): δ 196.5, 168.5, 141.7, 141.0, 134.3, 131.9, 130.3, 129.7, 123.5, 123.4, 115.6, 54.9, 49.7, 37.0, 26.0, 25.2, 23.4, 17.3. LRMS (m/z): [M]+ calcd. for C23H25N2O2+ 361.19, found 361.30.
7b (Linker = (CH2)7)
White solid, 645.7 mg, 80% yield. 1H NMR (400 MHz, Chloroform-d): δ 7.85–7.80 (m, 2H), 7.73–7.69 (m, 2H), 7.64–7.54 (m, 4H), 4.69 (t, J = 7.6 Hz, 2H), 3.66 (t, J = 6.8 Hz, 2H), 3.12 (s, 3H), 1.97–1.89 (m, 2H), 1.70–1.63 (m, 2H), 1.66 (s, 6H), 1.60 (s, 3H), 1.49–1.45 (m, 4H), 1.36–1.34 (m, 2H). 13C NMR (125 MHz, Chloroform-d): δ 195.8, 168.6, 141.8, 141.1, 134.1, 132.2, 130.3, 129.7, 123.5, 123.3, 115.5, 54.8, 50.3, 37.8, 28.7, 28.4, 28.0, 26.7, 26.5, 23.3, 17.2. LRMS (m/z): [M]+ calcd for C26H31N2O2+ 403.24, found 403.37.
7c (Linker = CH2CH2(OCH2CH2)2)
White solid, 354.0 mg, 42% yield. 1H NMR (400 MHz, Chloroform-d): δ 7.87–7.85 (m, 2H), 7.77–7.75 (m, 2H), 7.59–7.52 (m, 3H), 4.95 (t, J = 4.8 Hz, 2H), 4.00–3.98 (t, J = 4.8 Hz, 2H), 3.78–3.75 (t, J = 5.6 Hz, 2H), 3.58 (t, J = 5.6 Hz, 2H), 3.52–3.48 (m, 4H), 3.01 (s, 3H), 1.62 (s, 6H). 13C NMR (125 MHz, Chloroform-d) δ 198.0, 168.3, 141.5, 134.4, 132.1, 130.1, 129.7, 123.5, 123.2, 115.9, 89.9, 70.6, 70.1, 68.1, 67.2, 54.8, 50.6, 37.3, 23.2, 16.8, 15.4. LRMS (m/z): [M]+ calcd. for C25H29N2O4+ 421.21, found 421.29.
General procedure for the synthesis of 8a–c
A solution of 7 (1 mmol, 1.0 equiv.), 2a (261.0 mg, 1.1 equiv.) and NaOAc (328.1 mg, 4.0 equiv.) in Ac2O (10 mL) was stirred at 60 °C for 2 h. After being cooled to room temperature, the reaction mixture was quenched with H2O (50 mL) and extracted with DCM for three times. The combined organic layer was dried over Na2SO4, filtrated and the filtrates was concentrated under reduced pressure. The purification using column chromatography on silica gel (eluted with DCM/MeOH = 10/1) was performed to afforded the 8.
8a (Linker = (CH2)4)
Violet solid, 412.0 mg, 71% yield. 1H NMR (400 MHz, Chloroform-d): δ 8.60 (d, J = 15.2 Hz, 1H), 7.82–7.80 (m, 2H), 7.72–7.69 (m, 2H), 7.54–7.40 (m, 7H), 7.06 (d, J = 15.2 Hz, 1H), 6.99 (s, 1H), 4.91 (t, J = 5.6 Hz, 2H), 3.78 (t, J = 5.6 Hz, 2H), 2.98 (t, J = 6.0 Hz, 2H), 2.75 (t, J = 6.0 Hz, 2H), 1.98 (m, 6H), 1.83 (s, 6H). LRMS (m/z): [M]+ calcd. for C38H34N3O3+ 580.26, found 580.46.
8b (Linker = (CH2)7)
Violet solid, 385.8 mg, 62% yield. 1H NMR (400 MHz, Chloroform-d) δ 8.59 (d, J = 15.2 Hz, 1H), 7.82–7.80 (m, 2H), 7.71–7.69 (m, 2H), 7.55–7.38 (m, 7H), 7.05 (d, J = 15.2 Hz, 1H), 6.95 (s, 1H), 4.82 (t, J = 7.2 Hz, 2H), 3.64 (t, J = 7.2 Hz, 2H), 2.97 (d, J = 6.8 Hz, 2H), 2.74 (t, J = 6.4 Hz, 2H), 1.98–1.92 (m, 4H), 1.84 (s, 6H), 1.68–1.42 (m, 12H). LRMS (m/z): [M]+ calcd. for C41H40N3O3+ 622.31, found 622.59.
8c (Linker = CH2CH2(OCH2CH2)2)
Violet solid, 473.8 mg, 74% yield. 1H NMR (400 MHz, Chloroform-d) δ 8.57 (d, J = 15.6 Hz, 1H), 7.80–7.76 (m, 2H), 7.72–7.69 (m, 2H), 7.64 (d, J = 8.0 Hz, 1H), 7.53–7.35 (m, 6H), 7.10 (d, J = 15.6 Hz, f1H), 6.91 (s, 1H), 5.04 (t, J = 4.8 Hz, 2H), 4.03 (t, J = 4.8 Hz, 2H), 3.70 (t, J = 6.0 Hz, 2H), 3.56–3.47 (m, 6H), 2.91–2.88 (m, 2H), 2.2 (t, J = 6.0 Hz, 2H), 1.95–1.92 (m, 2H), 1.83 (s, 6H). LRMS (m/z): [M]+ calcd. for C40H38N3O5+ 640.28, found 640.51.
General procedure for the synthesis of 9a-c
A solution of 8 (1 mmol, 1.0 equiv.), N2H4•H2O (80% conc. in H2O, 20 equiv.) in DCM/MeOH (v/v = 1:4, 20 mL) was stirred at room temperature for 2 h. The reaction mixture was quenched with H2O, and extracted with DCM for three times. The combined organic layer was dried over Na2SO4, filtrated and the filtrates was concentrated under reduced pressure to afford the 9, which was used without further purification for the next step.
9a (Linker = (CH2)4)
1H NMR (500 MHz, Methanol-d4) δ 8.70 (d, J = 15.5 Hz, 1H), 7.70–7.69 (m, 1H), 7.65–7.61 (m, 2H), 7.52–7.44 (m, 4H), 7.14 (s, 1H), 6.66 (d, J = 15.5 Hz, 1H), 4.45 (t, J = 7.5 Hz, 2H), 2.91 (t, J = 7.5 Hz, 2H), 2.70 (t, J = 6.0 Hz, 2H), 2.66 (t, J = 6.0 Hz, 2H), 1.95–1.89 (m, 2H), 1.87–1.84 (m, 2H), 1.77 (s, 6H), 1.77–1.73 (m, 2H). LRMS (m/z): [M]+ calcd. for C30H32N3O+ 450.25, found 450.41.
9b (Linker = (CH2)7)
1H NMR (500 MHz, Methanol-d4) δ 8.75 (d, J = 15.5 Hz, 1H), 7.76− (s, 1H), 7.72−7.67 (m, 2H), 7.59−7.52 (s, 4H), 7.20 (s, 1H), 6.73 (d, J = 15.5 Hz, 1H), 4.48 (t, J = 7.5 Hz, 2H), 2.89 (t, J = 7.5 Hz, 2H), 2.78 (t, J = 6.0 Hz, 2H), 2.72 (t, J = 6.0 Hz, 2H), 1.96−1.90 (s, 4H), 1.83 (s, 6H), 1.67−1.61 (m, 2H), 1.50−1.37 (s, 6H). LRMS (m/z): [M]+ calcd. for C33H38N3O+ 492.30, found 492.47.
9c (Linker = CH2CH2(OCH2CH2)2)
LRMS (m/z): [M]+ calcd for C32H36N3O3+ 510.28, found 510.53.
General procedure for the synthesis of CyNA4, CyNA7, CyNAP
A solution of 9, Cys-NH(Boc)-S(Tri) (556.3 mg, 1.2 equiv.), HBTU (455.1 mg, 1.4 equiv.), DIPEA (181.0 mg, 1.4 equiv.) in DCM (20 mL) was stirred at room temperature for 2 h. The reaction mixture was quenched with H2O, and extracted with DCM for three times. The combined organic layer was dried over Na2SO4, filtrated and concentrated under reduced pressure to afforded the intermedia 10, which was used without further purification for the next reaction.
A solution of 10, Et3SiH (232.6 mg, 2.0 equiv.) in CF3COOH (20% in DCM, 10 mL) was stirred at room temperature for 6 h. The reaction mixture was quenched by H2O, and extracted with DCM for three times. The combined organic layer was dried over Na2SO4, filtrated and concentrated under reduced pressure. The purification using column chromatography on silica gel (eluted with DCM/MeOH = 5/1) was performed to afforded the CyNAX.
CyNA4
Purplish red solid, 22.1 mg, 4% yield in three steps. 1H NMR (400 MHz, Methanol-d4) δ 8.78 (d, J = 15.2 Hz, 1H), 7.77 (s, 1H), 7.71 (td, J = 7.2, 1.6 Hz, 2H), 7.61–7.52 (m, 3H), 7.29–7.16 (m, 2H), 6.75 (d, J = 15.2 Hz, 1H), 4.52 (t, J = 7.6 Hz, 2H), 4.21 (dt, J = 20.8, 5.2 Hz, 1H), 3.97 (t, J = 6.8 Hz, 1H), 3.44 (dt, J = 13.6, 6.8 Hz, 1H), 3.22 (dt, J = 13.6, 6.8 Hz, 1H), 2.98–2.90 (m, 3H), 2.74 (t, J = 6.0 Hz, 2H), 2.03–1.91 (m, 4H), 1.85 (s, 6H), 1.73–1.69 (m, 2H). HRMS (m/z): [M]+ calcd. for C33H37N4O2S+ 553.2632, found 553.2596.
CyNA7
Purplish red solid, 35.7 mg, 6% yield in three steps. 1H NMR (400 MHz, Methanol-d4) δ 8.77 (d, J = 15.2 Hz, 1H), 7.76 (s, 1H), 7.72 (d, J = 7.2 Hz, 1H), 7.68 (d, J = 7.6 Hz, 1H), 7.61–7.54 (m, 3H), 7.20–7.16 (m, 2H), 6.73 (d, J = 15.2 Hz, 1H), 4.48 (t, J = 7.6 Hz, 2H), 4.20 (dt, J = 16.4, 5.2 Hz, 1H), 3.97 (t, J = 6.0 Hz, 1H), 3.21–2.93 (m, 7H), 2.72 (t, J = 6.4 Hz, 2H), 1.96–1.92 (m, 2H), 1.84 (s, 6H), 1.57–139 (m, 8H). HRMS (m/z): [M]+ calcd. for C36H43N4O2S+ 595.3101, found 595.3096.
CyNAP
Purplish red solid, 18.4 mg, 3% yield in three steps. 1H NMR (500 MHz, Methanol-d4): δ 8.77 (d, J = 15.5 Hz, 1H), 7.78 (s, 1H), 7.74–7.69 (m, 2H), 7.60–7.54 (m, 4H), 7.20 (s, 1H), 6.90 (d, J = 15.5 Hz, 1H), 4.71 (t, J = 5.0 Hz, 2H), 3.99 (t, J = 5.0 Hz, 2H), 3.95 (t, J = 6.0 Hz, 1H), 3.60–3,58 (m, 2H), 3.51–3.49 (m, 2H), 3.42–3.36 (m, 3H), 3.24–3.16 (m, 2H), 2.94 (dd, J = 14.5, 5.0 Hz, 1H), 2.89 (dd, J = 14.5, 6.0 Hz, 1H), 2.80 (t, J = 6.0 Hz, 2H), 2.73 (t, J = 6.0 Hz, 2H), 1.97–1.93 (m, 2H), 1.86 (s, 6H). HRMS (m/z): [M]+ calcd. for C35H41N4O4S+ 613.2843, found 613.2808.
CyNAP-SS
A solution of 9c (51.0 mg, 0.1 mmol, 1.0 equiv.), Cys-NH(Fmoc)-S(S-tBu) (51.8 mg, 1.2 equiv.), HBTU (53.1 mg, 1.4 equiv.), DIPEA (18.1 mg, 1.4 equiv.) in DCM (5 mL) was stirred at room temperature for 2 h. The reaction mixture was quenched with H2O, and extracted with DCM for three times. The combined organic layer was dried over Na2SO4, filtrated and concentrated under reduced pressure to afforded the intermedia 11, which was used without further purification for the next reaction.
A solution of 11 in piperidine (20% in DCM) was stirred at room temperature for 2 h. The reaction mixture was quenched by H2O, and extracted with DCM for three times. The combined organic layer was dried over Na2SO4, filtrated and concentrated under reduced pressure. The purification using column chromatography on silica gel (eluted with DCM/MeOH = 5/1) was performed to afforded the CyNAP-SS (9.1 mg, 13% in two-step) as purplish red solid. 1H NMR (400 MHz, Methanol-d4) δ 8.82 (d, J = 15.6 Hz, 1H), 7.82 (s, 1H), 7.78–7.74 (m, 2H), 7.65–7.58 (m, 4H), 7.24 (s, 1H), 6.95 (d, J = 15.6 Hz, 1H), 4.76 (t, J = 5.2 Hz, 2H), 4.04 (t, J = 5.2 Hz, 2H), 3.64–3.63 (m, 2H), 3.55–3.52 (m, 2H), 3.45 (t, J = 5.6 Hz, 2H), 2.84 (t, J = 6.2 Hz, 2H), 2.77 (t, J = 6.2 Hz, 2H), 2.01–1.98 (m, 2H), 1.91 (s, 6H), 1.36 (s, 9H). HRMS (m/z): [M]+ calcd. for C39H49N4O4S2+ 701.3190, found 701.3188.
CyNAP-S
A solution of 9c (51.0 mg, 0.1 mmol, 1.0 equiv.), Cys-NH(Fmoc)-S(tBu) (55.63 mg, 1.2 equiv.), HBTU (53.1 mg, 1.4 equiv.), DIPEA (18.1 mg, 1.4 equiv.) in DCM (5 mL) was stirred at room temperature for 2 h. The reaction mixture was quenched with H2O, and extracted with DCM for three times. The combined organic layer was dried over Na2SO4, filtrated and concentrated under reduced pressure to afforded the intermedia 12, which was used without further purification for the next reaction.
A solution of 12 in piperidine (20% in DCM) was stirred at room temperature for 2 h. The reaction mixture was quenched by H2O, and extracted with DCM for three times. The combined organic layer was dried over Na2SO4, filtrated and concentrated under reduced pressure. The purification using column chromatography on silica gel (eluted with DCM/MeOH = 5/1) was performed to afforded the compound CyNAP-S (10.0 mg, 15% in two-step) as purplish red solid. 1H NMR (400 MHz, Methanol-d4) δ 8.82 (d, J = 15.2 Hz, 1H), 7.82 (s, 1H), 7.80–7.73 (m, 2H), 7.66–7.58 (m, 4H), 7.25 (s, 1H), 6.95 (d, J = 15.2 Hz, 1H), 4.77 (t, J = 4.8 Hz, 2H), 4.05 (t, J = 4.8 Hz, 2H), 3.94 (s, 1H), 3.65 (dd, J = 6.0, 3.2 Hz, 2H), 3.56 (dd, J = 6.0, 3.2 Hz, 2H), 3.45 (d, J = 5.3 Hz, 2H), 3.30 (s, 1H), 3.03 (s, 1H), 2.95 (s, 1H), 2.85 (t, J = 6.2 Hz, 2H), 2.78 (t, J = 6.2 Hz, 2H), 2.04–1.99 (m, 2H), 1.92 (s, 6H), 1.37 (s, 9H). HRMS (m/z): [M]+ calcd. for C39H49N4O4S+ 669.3469, found 669.3427.
CyNAP-T
A solution of compound 9c (20.0 mg, 0.04 mmol), D-Cys (48.4 mg, 10 equiv.) and MeOH/PBS (2.0 mL, V/V = 3:7, pH = 7.4) was stirred at room temperature overnight. The mixture reaction was concentrated under reduced prepress, and the resulted residue was purified with flash column chromatography on silica gel (eluted with DCM/MeOH = 10/1) to afforded the compound CyNAP-T as blue solid in 6% yield. LRMS (m/z): [M]+ calcd. for C35H40N3O5S+ 614.27, found 614.38.
CyNAP-SS-FK
A solution of compound CyNAP-SS (35 mg, 0.05 mmol), Ac-FK (43.5 mg, 2.0 equiv.), HBTU (26.5 mg, 1.4 equiv.), DIPEA (9.0 mg, 1.4 equiv.) in THF (5 mL) was stirred at room temperature for 12 h. The was removed under reduced pressure, and the resultant residue was extracted with DCM for three times. The combined organic layer was dried over Na2SO4, filtrated and concentrated under reduced pressure. The purification using fresh column chromatography on silica gel (eluted with DCM/MeOH = 5/1) was performed to afforded the compound CyNAP-SS-FK (1.1 mg) as purplish red solid in 4% yield. 1H NMR (500 MHz, Chloroform-d) δ 8.61 (d, J = 15.5 Hz, 1H), 7.70–7.64 (m, 2H), 7.58–7.50 (m, 3H), 7.46 (t, J = 8.0 Hz, 2H), 7.38–7.31 (m, 3H), 7.20 (q, J = 8.0, 7.0 Hz, 2H), 6.99 (d, J = 15.5 Hz, 1H), 6.91 (s, 1H), 5.06 (s, 4H), 4.71 (dd, J = 44.0, 16.5 Hz, 4H), 4.43 (s, 1H), 4.15 (s, 1H), 4.05 (d, J = 10.5 Hz, 2H), 3.57–3.49 (m, 3H), 3.34 (dd, J = 13.5, 3.5 Hz, 1H), 3.22–3.04 (m, 6H), 2.73 (s, 3H), 2.03 (s, 2H), 1.84 (s, 6H), 1.68 (s, 15H), 1.44 (d, J = 5.5 Hz, 2H), 1.31 (s, 9H), 1.27 (d, J = 12.0 Hz, 6H). HRMS (m/z): [M]+ calcd. for C61H80N7O9S2+ 1118.5453, found 1118.5460.
Kinetics study of intermolecular condensation reaction of CyN-(1−3) with Cysteine
The condensation reaction kinetics were performed according to the report literature46. A solution of CyN-X (X = 1, 2, 3) (50 μM), Cysteine (1.0 equiv., 50 μM), TCEP (100 equiv.) and PBS (containing 20% MeOH, adjust the pH of the reaction solution to 7.4 with NaHCO3 aq.) was attired at room temperature (30 °C). The progress at different time points was monitored by HPLC at 600 nm. The liner analysis of 1/[CyN-X] vs. time gives the 2nd order rate constant. CyN-1: 0.30956 M−1 s−1; CyN-2: 0.15179 M−1 s−1; CyN-3: 0.22190 M−1 s−1;
Kinetics study of intramolecular condensation reaction of CyNAX (X = 4, 7, P)
The condensation reaction kinetics were performed according to the report literature46. A solution of CyNAX (X = 4, 7, P) (50 μM), TCEP (100 equiv.) and PBS (containing 20% MeOH, adjust the pH of the reaction solution to 7.4 with NaHCO3 aq.) was attired at room temperature (30 °C). The progress at different time points was monitored by HPLC at 600 nm with the addition of methylene blue (MB, λex = 700 nm) as internal standard. The liner analysis of In[CyNAX] vs. time gives the 1st order rate constant. CyNA4: −2.91 × 10−5 s−1; CyNA7: −0.89 × 10−5 s−1; CyNAP: −7.97 × 10−5 s−1; CyNA4-C: MALDI-TOF MS: m/z: [M]+ Calcd for C33H34N3O2S+ 536.237; Found 536.241; CyNA7-C: MALDI-TOF MS: m/z: [M]+ Calcd for C36H40N3O2S+ 578.284; Found 578.291; CyNAp-C: 1H NMR (400 MHz, Methanol-d4) δ 8.82 (d, J = 15.2 Hz, 1H), 7.82 (s, 1H), 7.80 (m, 2H), 7.65–7.59 (m, 4H), 7.25 (s, 1H), 6.94 (d, J = 15.2 Hz, 1H), 4.77 (t, J = 5.2 Hz, 2H), 4.05–4.03 (m, 3H), 3.98 (t, J = 6.0 Hz, 1H), 3.65–3.63 (m, 3H), 3.56–3.53 (m, 3H), 3.45 (m, 2H), 2.98–2.95 (m, 1H), 2.85 (t, J = 6.0 Hz, 2H), 2.78 (t, J = 6.0 Hz, 2H), 2.00 (t, J = 6.0 Hz, 2H), 1.91 (s, 6H). MALDI-TOF MS (m/z): [M]+ calcd. for C35H38N3O4S+ 596.258, found 596.262;
Competitive reaction between intramolecular cyclization and intermolecular condensation of CyNAP in the presence of varying concentration of free Cys
The condensation reaction kinetics were performed according to the report literature46. A solution of CyNAP (20 μM), Cysteine (20 μM, 2 mM, and 20 mM), TCEP (40 μM, 4 mM, and 40 mM) and PBS (containing 20% MeOH, adjust the pH of the reaction solution to 7.4 with NaHCO3 aq.) was attired at room temperature (30 °C) for 24 h. The progress was monitored by HPLC at 600 nm.
Computational details
The geometric optimization and frequency analysis of all molecules were conducted using the density functional theory method at the B3LYP/6-311 G(d,p) level with Grimme’s DFT-d3 correction in Gaussian 16 program package. The visualization of frontier molecular orbitals was achieved using the IQmol molecular viewer package.
Measure of fluorescence quantum yields
Methylene Blue was used as a standard with a known fluorescence quantum yield (Φ) value of 0.49 in PBS69. Fluorescence quantum yields were calculated using the following equation:
where Φs and Φf are the fluorescence quantum yields of the standard and the samples, respectively; As and Af are the emission areas of the standard and the samples, respectively; Abss and Absf are the absorbance of the standard and the samples at the Wavelength of excitation; ηs and ηf are the refractive indices of the standard and the samples, respectively.
TEM sample preparation
1 μM CyNAP-SS-FK was incubated with TCEP and Cat B for 12 h in phosphate buffered solution (1×PBS, pH = 7.4), the carbon-coated side of the grid was gently immersed in the dispersion, the photo was taken by Hitachi TEM system. Magnification = 40,000, Accelerating Voltage = 100,000.
Enzyme-responsive experiment
The CyNAP-SS-FK at a concentration of 20 μM was separately incubated with GSH, Cat B enzyme, and a mixture of both at 37 °C for 6 h. After incubation, fluorescence spectra were collected to calculate the fold increase in fluorescence signal before and after the response.
Specificity experiment
The CyNAP-SS-FK (40 μM) was incubated with the specified enzymes, metal ions (60 μM), and ROS (60 μM) in PBS (10 mM, pH 7.4) at 37 °C. Mg2+ refers to magnesium chloride; Ca2+ to calcium chloride; ClO- to sodium hypochlorite; H2O2 to hydrogen peroxide; SO42- to sodium sulfate; GSH to glutathione; TCEP to Tris(2-carboxyethyl) phosphine hydrochloride; Cas-3 to caspase 3; β-Galactosidase (n = 3, mean ± standard deviation).
Cytotoxicity assay
The cytotoxicity of CyNAP was detected by Cell Counting Kit-8. Human hepatocellular carcinoma cells (HCC-LM3 cells) were obtained from American Type Culture Collection (ATCC). HCC-LM3 cells were grown in DMEM containing 10% fetal bovine serum (FBS), and 1% penicillin & streptomycin, cultured at 37 °C with 5% CO2. HCC-LM3 cells were implanted in 96-well plates (4000 cells per well in 100 μL complete culture medium) and cultured overnight, incubated with different concentrations of CyNAP (0, 20, 40, 60, 80, 100 μM) for 12 h. Then, 10 μL CCK-8 solution was added into each wall, followed by further incubation for 1 h. The relative cell viability was performed based on the absorbance of the formazan at 450 nm as OD value. Cells viability = [A (sample) − A (blank)]/[A (control) − A (blank)] × 100%, where A is corresponding to OD value.
Cell imaging study
Human hepatocellular carcinoma cells (HCC-LM3 cells) were obtained from the American Type Culture Collection (ATCC, CBP60654). HCC-LM3 cells were grown in DMEM containing 10% fetal bovine serum (FBS) and 1% penicillin and streptomycin and cultured at 37 °C with 5% CO2. Various concentrations of CyNA4, CyNA7, CyNAP, CyNAP-S, CyNAP-SS, CyNAP-SS-FK were incubated with HCC-LM3 cells (4000 cells per wall) in a 96-wall confocal plate. The control group was treated with or without inhibitor BSO (CAS S9728), and all groups were cultured overnight. After staining with DAPI, imaging was conducted by a confocal laser scanning microscope FV3000 (Olympus). The excitation and emission wavelengths were 590 ± 10 and 619 nm. Cellular fluorescence intensities were quantified using ImageJ software.
Biodistribution and biocompatibility of CyNAP in living mice
The mice were all male Balb/c-nu, 4–6 weeks old. All mice experimental procedures were approved by the Institutional Animal Care and Use Committee (IACUC), Sun Yat-sen University (SYSU) with the approval number SYSU-IACUC-2022000220. According to ethical standards, the tumor volume of the experimental mice should not exceed 2000 mm3. All our tumor-bearing experiments adhere to ethical requirements. To determine the clearance pathway and biocompatibility in living mice, 4–6 weeks old mice (Balb/c-nu, 15 mg/kg body weight) was injected with CyNAP intravenously. All mice were fed with non-fluorescent chow (Xietong Pharmaceutical Bioengineering CO., Ltd, XT19008) and kept in SPF environment. The experimental group and the control group of mice were both randomly assigned to their respective groups. The dynamic NIRF imaging and intensities of abdominal cavity and resected major organs (heart, liver, spleen, lung, kidneys, bladder, intestine, muscle, and skin) was performed and analyzed by IVIS Lumina XR Series III (Perkin Elmer Inc, USA) from mice post-injection at various time. The major organs were collected and placed into 4% paraformaldehyde (PFA) for histological examination.
Stability of the probe CyNAP-SS-FK
In response to the stability of the probe, we conducted further research. Initially, we injected CyNAP-SS-FK into the healthy mice via the tail vein. After 12 h post-injection, blood was collected from the heart, centrifuged after standing, and the supernatant was taken. A fluorescent probe stock solution was prepared and quantified by ultraviolet absorbance intensity to match the concentration of the fluorescent probe in the serum. Subsequently, fluorescence spectra were collected to evaluate the stability of the probe’s fluorescence performance by comparing the changes in fluorescence intensity before and after injection. The experimental results, as shown in the figure, indicate that there was hardly any significant change in the fluorescence spectra of the fluorescent probe before and after injection, suggesting that the fluorescent probe possesses a certain level of optical stability.
Establishment of orthotopic HCC animal model in living mice
Balb/c-nu mice (4–6 weeks, male) were randomly selected and utilized for human HCC cells implantation. HCC-LM3 cells were cultured and mixed 1:1 with Matrigel and PBS, temporarily stored on ice. Mice were completely anesthetized and the abdominal cavity was opened to expose their liver. A mixture of cultured HCC-LM3 cells and stromal gel was injected into the liver, followed by closing the abdominal cavity and returning the mice to the SPF environment for further rearing. All animals were fed with non-fluorescent chow.
Real-time NIRF imaging of HCC
Following implantation of tumors in mice with various time points (0, 5, 10, and 15 days), CyNAP were administered intravenously in the tail vein (15 mg/kg body weight), and NIRF imaging of the major organs was acquired at different time points post injection. Fluorescence images were conducted by an IVIS Lumina XR series III system with an acquisition time of 4.0 s (excitation at 620 ± 10 nm and emission at 710 ± 10 nm). Blood samples from each group were collected and placed at −20 °C for biochemical analysis. Tumors and major organs from each group were collected and placed into 4% paraformaldehyde (4% PFA) at 4 °C for histological examination and immunofluorescence assays.
Histopathologic studies and immunofluorescence assay
All collected tissue were fixed with 4% PFA. After routine dehydration, paraffin-embedded tissues were sectioned (10 μm of thickness) for H&E staining (hematoxylin and eosin staining) according to the instructions. The stained sections were observed under the EVOS M7000 microscope. After fixing the mouse tumor tissue in PFA, 30% sucrose solution was fully dehydrated and embedded in optimal cutting temperature compound (OCT), which was placed into liquid nitrogen for coagulation and frozen sections were made for subsequent staining. Sections were fixed by 100% acetone at −20 °C for 20 min after rewarming at room temperature. Goat serum was blocked for 90 min at room temperature, then primary antibody Cathepsin B Polyclonal antibody (12216-1-AP, Proteintech Group, 1:500 dilution) was added dropwise and incubated overnight at 4 degrees Celsius in the refrigerator. Dropwise addition of secondary antibody (L3016, SAB, 1:500 dilution) for 1 h at room temperature, followed by DAPI staining for 5–10 min and sealing with sealing agent. Immunofluorescence experiments were washed with PBS 3 times for 5 min each step. Fluorescence signal distribution and intensity were observed using a laser confocal microscope after staining.
Biochemical analysis
Blood samples were immediately collected from both health mice and HCC model mice after injected with CyNAP (15 mg kg−1 body weight) or PBS solution in 24 h. The collected blood samples were centrifuged at 3000 r.p.m. for 15 min at room temperature. The supernatant was stored for subsequent test. ALT (alanine aminotransferase), AST (aspartate aminotransferase) were measured with biological reagent kif (C009-1-1; C010-1-1) from Nanjing Jiancheng Bioengineering Institute. BUN (blood urea nitrogen) and sCr (serum creatinine) were measured with biological reagent kid (#D799850-0100; #D799853-0096) from Sangon Biotech (Shanghai) Company. Processed samples were tested following the instructions provided with the assay kits. UV absorption values were detected by multimode microplate reader (Synergy H1 of BioTek) and calculated by reference standard.
Statistical analysis
The in vitro, in vivo and ex vivo NIFR signal intensity was quantified with region of interest analysis (ROI) analysis by Living Image 4.4 software. All the data was mean ± standard deviation (SD) unless stated otherwise. Investigators were blinded to group allocation during experiments. Statistical differences between two groups were tested with a two-tailed Student’s t test and more than three groups were determined by one-way analysis of variance followed by Tukey’s post hoc test. For all tests, P values less than 0.05 were considered statistically significant. *p < 0.05, **p < 0.01, ***p < 0.001 and ****p < 0.0001. All statistical calculations were performed using GraphPad Prism 9.5.0 and OriginPro 2024, including assumptions of tests used.
Reporting summary
Further information on research design is available in the Nature Portfolio Reporting Summary linked to this article.
Data availability
References
Zhu, F. Y. et al. Recent advances in super-resolution optical imaging based on aggregation-induced emission. Chem. Soc. Rev. 53, 3350–3383 (2024).
Bertolini, M., Wong, M. S., Mendive-Tapia, L. & Vendrell, M. Smart probes for optical imaging of T cells and screening of anti-cancer immunotherapies. Chem. Soc. Rev. 52, 5352–5372 (2023).
Wu, L., Huang, J., Pu, K. & James, T. D. Dual-locked spectroscopic probes for sensing and therapy. Nat. Rev. Chem. 5, 406–421 (2021).
Li, J. & Pu, K. Development of organic semiconducting materials for deep-tissue optical imaging, phototherapy and photoactivation. Chem. Soc. Rev. 48, 38–71 (2019).
Kenry, DuanY. & Liu, B. Recent Advances of Optical Imaging in the Second Near‐Infrared Window. Adv. Mater. 30, 1802394 (2018).
Chen, C., Zhang, X., Gao, Z., Feng, G., Ding, D. Preparation of AIEgen-based near-infrared afterglow luminescence nanoprobes for tumor imaging and image-guided tumor resection. Nat. Protoc. https://doi.org/10.1038/s41596-024-00990-4 (2024).
Zhang, X. et al. A Universal Photoactivatable Tag Attached to Fluorophores Enables Their Use for Single‐Molecule Imaging. Angew. Chem. Int. Ed. 61, e202211767 (2022).
Zhang, T. et al. Near‐Infrared Aggregation‐Induced Emission Luminogens for In Vivo Theranostics of Alzheimer’s Disease. Angew. Chem. Int. Ed. 62, e202211550 (2022).
Wang, R., Zhou, L., Wang, W., Li, X. & Zhang, F. In vivo gastrointestinal drug-release monitoring through second near-infrared window fluorescent bioimaging with orally delivered microcarriers. Nat. Commun. 8, 14702 (2017).
Zhang, Y., Zhang, G., Zeng, Z. & Pu, K. Activatable molecular probes for fluorescence-guided surgery, endoscopy and tissue biopsy. Chem. Soc. Rev. 51, 566–593 (2022).
Wu, X., Wang, R., Kwon, N., Ma, H. & Yoon, J. Activatable fluorescent probes for in situ imaging of enzymes. Chem. Soc. Rev. 51, 450–463 (2022).
Liu, Y. et al. Chemical Design of Activatable Photoacoustic Probes for Precise Biomedical Applications. Chem. Rev. 122, 6850–6918 (2022).
Zeng, Z., Liew, S. S., Wei, X. & Pu, K. Hemicyanine‐Based Near‐Infrared Activatable Probes for Imaging and Diagnosis of Diseases. Angew. Chem. Int. Ed. 60, 26454–26475 (2021).
Cheng, Z. et al. Enzyme‐Activatable Near‐Infrared Hemicyanines as Modular Scaffolds for in vivo Photodynamic Therapy. Angew. Chem. Int. Ed. 63, e202404587 (2024).
Wang, H. et al. Prediction of Early Atherosclerotic Plaques Using a Sequence‐Activated Fluorescence Probe for the Simultaneous Detection of γ‐Glutamyl Transpeptidase and Hypobromous Acid. Angew. Chem. Int. Ed. 63, e202315861 (2023).
Yao, Y., Zhang, Y., Yan, C., Zhu, W.-H. & Guo, Z. Enzyme-activatable fluorescent probes for β-galactosidase: from design to biological applications. Chem. Sci. 12, 9885–9894 (2021).
Huang, J., Li, J., Lyu, Y., Miao, Q. & Pu, K. Molecular optical imaging probes for early diagnosis of drug-induced acute kidney injury. Nat. Mater. 18, 1133–1143 (2019).
Crossley, S. W. M. et al. A Transfer Hydrogenation Approach to Activity-Based Sensing of Formate in Living Cells. J. Am. Chem. Soc. 146, 8865–8876 (2024).
Yi, N.-B. et al. Endogenous formaldehyde responsive fluorescent probe for bioimaging. Nano Res. 16, 13029–13041 (2023).
Yan, H., Wang, Y., Huo, F. & Yin, C. Fast-Specific Fluorescent Probes to Visualize Norepinephrine Signaling Pathways and Its Flux in the Epileptic Mice Brain. J. Am. Chem. Soc. 145, 3229–3237 (2023).
Zhang, Y. et al. Rational construction of a reversible arylazo-based NIR probe for cycling hypoxia imaging in vivo. Nat. Commun. 12, 2772 (2021).
Zhang, Y. et al. Inflammation-Responsive Nanoagents for Activatable Photoacoustic Molecular Imaging and Tandem Therapies in Rheumatoid Arthritis. ACS Nano 18, 2231–2249 (2024).
Liu, J. et al. Unveiling the Crucial Roles of O2•–and ATP in Hepatic Ischemia–Reperfusion Injury Using Dual-Color/Reversible Fluorescence Imaging. J. Am. Chem. Soc. 145, 19662–19675 (2023).
Liu, J. et al. Precision Navigation of Hepatic Ischemia–Reperfusion Injury Guided by Lysosomal Viscosity-Activatable NIR-II Fluorescence. J. Am. Chem. Soc. 144, 13586–13599 (2022).
Fang, Y. et al. Design, Synthesis, and Application of a Small Molecular NIR-II Fluorophore with Maximal Emission beyond 1200 nm. J. Am. Chem. Soc. 142, 15271–15275 (2020).
Chen, Z. et al. An Optical/Photoacoustic Dual-Modality Probe: Ratiometric in/ex Vivo Imaging for Stimulated H2S Upregulation in Mice. J. Am. Chem. Soc. 141, 17973–17977 (2019).
Kim, J. et al. In situ self-assembly for cancer therapy and imaging. Nat. Rev. Mater. 8, 710–725 (2023).
Xi, D. et al. Strong π–π Stacking Stabilized Nanophotosensitizers: Improving Tumor Retention for Enhanced Therapy for Large Tumors in Mice. Adv. Mater. 34, 2106797 (2021).
Liu, B., Wu, R., Gong, S., Xiao, H. & Thayumanavan, S. In Situ Formation of Polymeric Nanoassemblies Using an Efficient Reversible Click Reaction. Angew. Chem. Int. Ed. 59, 15135–15140 (2020).
Wang, L. et al. PKM2 allosteric converter: A self-assembly peptide for suppressing renal cell carcinoma and sensitizing chemotherapy. Biomaterials 296, 122060 (2023).
Li, H. et al. Self-assembly of peptide nanocapsules by a solvent concentration gradient. Nat. Nanotechnol. 19, 1141–1149 (2024).
Zhou, J. et al. Enzymatic Self-Assembly Confers Exceptionally Strong Synergism with NF-κB Targeting for Selective Necroptosis of Cancer Cells. J. Am. Chem. Soc. 140, 2301–2308 (2018).
Zhang, P. et al. Target-Amplified Drug Delivery of Polymer Micelles Bearing Staudinger Ligation. ACS Appl. Mater. Interface 11, 32697–32705 (2019).
Bednarek, C., Wehl, I., Jung, N., Schepers, U. & Bräse, S. The Staudinger Ligation. Chem. Rev. 120, 4301–4354 (2020).
Jin, Y. et al. Ligation to Scavenging Strategy Enables On-Demand Termination of Targeted Protein Degradation. J. Am. Chem. Soc. 145, 7218–7229 (2023).
Vázquez, A. et al. Mechanism‐Based Fluorogenic trans‐Cyclooctene–Tetrazine Cycloaddition. Angew. Chem. Int. Ed. 56, 1334–1337 (2016).
Hu, X. et al. CBT‐Cys click reaction for optical bioimaging in vivo. View 4, 20220065 (2023).
Zhu, Y., Zhang, X., You, Q. & Jiang, Z. Recent applications of CBT-Cys click reaction in biological systems. Bioorgan. Med. Chem. 68, 116881 (2022).
Zhang, M. & Liang, G. Applications of CBT-Cys click reaction: past, present, and future. Sci. China Chem. 61, 1088–1098 (2018).
Wang, Y., An, R., Luo, Z. & Ye, D. Firefly Luciferin‐Inspired Biocompatible Chemistry for Protein Labeling and In Vivo Imaging. Chem. Eur. J. 24, 5707–5722 (2017).
Xu, L. et al. Dual Aggregations of a Near-Infrared Aggregation-Induced Emission Luminogen for Enhanced Imaging of Alzheimer’s Disease. J. Am. Chem. Soc. 145, 27748–27756 (2023).
Chen, Z. et al. Exploring the Condensation Reaction between Aromatic Nitriles and Amino Thiols To Optimize In Situ Nanoparticle Formation for the Imaging of Proteases and Glycosidases in Cells. Angew. Chem. Int. Ed. 59, 3272–3279 (2020).
Liu, X. & Liang, G. Dual aggregation-induced emission for enhanced fluorescence sensing of furin activity in vitro and in living cells. Chem. Commun. 53, 1037–1040 (2017).
Ye, D. et al. Bioorthogonal cyclization-mediated in situ self-assembly of small-molecule probes for imaging caspase activity in vivo. Nat. Chem. 6, 519–526 (2014).
Yuan, Y. et al. Detection of Glutathione in Vitro and in Cells by the Controlled Self-Assembly of Nanorings. Anal. Chem. 85, 1280–1284 (2013).
Ye, D., Liang, G., Ma, M. L. & Rao, J. Controlling Intracellular Macrocyclization for the Imaging of Protease Activity. Angew. Chem. Int. Ed. 50, 2275–2279 (2011).
Ullrich, S., George, J., Coram, A. E., Morewood, R. & Nitsche, C. Biocompatible and Selective Generation of Bicyclic Peptides. Angew. Chem. Int. Ed. 61, e202208400 (2022).
Abdelkader, E. H. et al. Genetic Encoding of Cyanopyridylalanine for In‐Cell Protein Macrocyclization by the Nitrile–Aminothiol Click Reaction. Angew. Chem. Int. Ed. 61, e202114154 (2022).
Zheng, X. et al. Condensation of 2-((Alkylthio)(aryl)methylene)malononitrile with 1,2-Aminothiol as a Novel Bioorthogonal Reaction for Site-Specific Protein Modification and Peptide Cyclization. J. Am. Chem. Soc. 142, 5097–5103 (2020).
Thombare, V. J. et al. Advancing Nitrile−Aminothiol Strategy for Dual and Sequential Bioconjugation. Chem. Eur. J. 30, e202401674 (2024).
Patil, N. A. et al. 2-Cyanoisonicotinamide Conjugation: A Facile Approach to Generate Potent Peptide Inhibitors of the Zika Virus Protease. ACS Med. Chem. Lett. 12, 732–737 (2021).
Verdoes, M. et al. Improved Quenched Fluorescent Probe for Imaging of Cysteine Cathepsin Activity. J. Am. Chem. Soc. 135, 14726–14730 (2013).
Liu, Y. et al. Molecular Engineering of Activatable NIR-II Hemicyanine Reporters for Early Diagnosis and Prognostic Assessment of Inflammatory Bowel Disease. ACS Nano 18, 8437–8451 (2024).
Zhou, Y. et al. Tailored Zwitterionic Hemicyanine Reporters for Early Diagnosis and Prognostic Assessment of Acute Renal Failure. Angew. Chem. Int. Ed. 62, e202315457 (2023).
Ruan, B. et al. Size‐Transformable Superoxide‐Triggered Nanoreporters for Crosstalk‐Free Dual Fluorescence/Chemiluminescence Imaging and Urinalysis in Living Mice. Angew. Chem. Int. Ed. 62, e202305812 (2023).
Huang, J. et al. A Renally Clearable Activatable Polymeric Nanoprobe for Early Detection of Hepatic Ischemia‐Reperfusion Injury. Adv. Mater. 34, 2201357 (2022).
Huang, J. et al. Renal clearable polyfluorophore nanosensors for early diagnosis of cancer and allograft rejection. Nat. Mater. 21, 598–607 (2022).
Huang, J. et al. A Renal‐Clearable Macromolecular Reporter for Near‐Infrared Fluorescence Imaging of Bladder Cancer. Angew. Chem. Int. Ed. 59, 4415–4420 (2020).
Sun, Q. et al. Coumarin-hemicyanine-based far-red to near-infrared fluorescent probes: A new generation of fluorescent probe design platform. TrAC Trends Anal. Chem. 167, 117272 (2023).
Liu, Y., Teng, L., Lou, X.-F., Zhang, X.-B. & Song, G. Four-In-One” Design of a Hemicyanine-Based Modular Scaffold for High-Contrast Activatable Molecular Afterglow Imaging. J. Am. Chem. Soc. 145, 5134–5144 (2023).
Li, H. et al. Activity-based NIR fluorescent probes based on the versatile hemicyanine scaffold: design strategy, biomedical applications, and outlook. Chem. Soc. Rev. 51, 1795–1835 (2022).
Tian, X. et al. A Molecular‐Splicing Strategy for Constructing a Near‐Infrared Fluorescent Probe for UDP‐Glucuronosyltransferase 1A1. Angew. Chem. Int. Ed. 60, 24566–24572 (2021).
Griffith, O. W. & Meister, A. Potent and specific inhibition of glutathione synthesis by buthionine sulfoximine (S-n-butyl homocysteine sulfoximine). J. Bio. Chem. 254, 7558–7560 (1979).
Ma, Y. et al. Rational Design of a Double-Locked Photoacoustic Probe for Precise In Vivo Imaging of Cathepsin B in Atherosclerotic Plaques. J. Am. Chem. Soc. 145, 17881–17891 (2023).
Richard, J.-A. De novo synthesis of phenolic dihydroxanthene near-infrared emitting fluorophores. Org. Biomol. Chem. 13, 8169–8817 (2015).
Okoh, O. A. et al. Synthesis and photophysical properties of meso-aminophenyl-substituted heptamethine dyes as potential leads to new contrast agents. Color. Technol. 135, 305–311 (2019).
Yuan, L. et al. Approach to Development of Near-Infrared Fluorescent Sensors for in Vivo Imaging. J. Am. Chem. Soc. 134, 13510–13523 (2012).
Sestito, S. et al. Design and synthesis of H2S-donor hybrids: A new treatment for Alzheimer’s disease. Eur. J. Med. Chem. 184, 111745 (2019).
Cwalinski, T. et al. Methylene Blue−Current Knowledge, Fluorescent Properties, and Its Future Use. J. Clin. Med. 9, 3538 (2020).
Acknowledgements
J.H. thanks the National Key R&D Program of China (2022YFA0912700), National Natural Science Foundation of China (22207130), the Key Research and Development Plan of Guangzhou City (202206080007), the Sun Yat-sen University Funded Program (2022_76220_B21127), the Fundamental Research Funds for the Central Universities (22lgqb35), the introduced innovative R&D team project of “The Pearl River Talent Recruitment Program” (2021ZT09Y544), and Guangdong-Hongkong-Macao research team project of the Guangdong basic and applied basic research foundation (2022b1515130008) for the financial support. J.H. acknowledges the award of “The Recruitment Program of Global Youth Experts” and the support from the start-up funding of Sun Yat-Sen University.
Author information
Authors and Affiliations
Contributions
W.X., S.Y. and J.H. conceived and designed the study. W.X. and S.Y. performed the probe synthesis and in vitro experiments. S.Y. performed in vivo experiments, histology experiments and cell imaging experiments. J. L. performed computer calculation. W.X., S.Y. and J.H. analyzed the data. W.X., S.Y. and J.H. drafted the manuscript. All authors contributed to the writing of this article.
Corresponding author
Ethics declarations
Competing interests
The authors declare no competing interests.
Peer review
Peer review information
Nature Communications thanks Nitin Patil, and the other, anonymous, reviewer(s) for their contribution to the peer review of this work. A peer review file is available.
Additional information
Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Supplementary information
Source data
Rights and permissions
Open Access This article is licensed under a Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 International License, which permits any non-commercial use, sharing, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if you modified the licensed material. You do not have permission under this licence to share adapted material derived from this article or parts of it. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by-nc-nd/4.0/.
About this article
Cite this article
Xu, W., Yi, S., Liu, J. et al. Nitrile-aminothiol bioorthogonal near-infrared fluorogenic probes for ultrasensitive in vivo imaging. Nat Commun 16, 8 (2025). https://doi.org/10.1038/s41467-024-55452-y
Received:
Accepted:
Published:
DOI: https://doi.org/10.1038/s41467-024-55452-y