Introduction

Energy transfer in biological processes is achieved through the transfer of electrons and protons1,2,3. In essence, the transfer of energy-carrying electrons and protons is the basis of life processes, and it is inseparable from the electron transport chain4,5,6. Organisms can pass messages between biomolecules via electron transfer, thereby achieving a variety of life processes7,8,9,10. After signaling molecule binding, receptor or adapter protein kinases become activated by phosphorylation, thereby turning on downstream signaling pathways11,12,13. Activation of signaling pathways is directly related to electron transfer between the signaling and receptor molecules since the phosphorylation of receptor molecules is accompanied by the formation of chemical bonds and further electron transfer5,14. Taking tumor development as an example, compared with normal cells, tumor cells achieve extremely active electron transfer to support their proliferation, differentiation, and migration15,16,17,18. If the monitoring of tumor proliferation-associated signal pathway activity could be achieved, it would be possible to evaluate therapeutic behavior on this basis. This possibility has motivated the need for develo** novel strategies to implement in vivo monitoring of electron transfer processes that occur in signaling pathways.

There is a variety of signaling pathways in a cell that switch on and off nimbly, and influence each other constantly, thus forming a complex network13,19,20. Because the activity of these signaling pathways changes rapidly, a rapid, sensitive monitor of dynamic changes in signal transduction is required. Techniques have been well developed to measure electron transfer between biological molecules, such as nuclear magnetic resonance spectroscopy and electrochemical methods21,22,23. Optical methods are mostly characterized by high temporal and spatial resolution, non-contact detection, and high-sensitive visualization based on machine vision11,24,25,26,27,28. Specifically, the development of quantum biological electron tunneling spectroscopy enables imaging of electron transfer in live cells, while it is not easily applicable for in vivo testingFull size image

The electrochemical activity of ZnGa2O4 was further characterized in different concentrations of glucose solution to determine the relationship between the afterglow signal and the electron transfer. ZnGa2O4 and GOD were sequentially coated on the electrode surface, and the electrode was employed for cyclic voltammetry in a solution containing different concentrations of glucose (Fig. 2b, c). While the GOD coating resulted in total passivation (Supplementary Fig. 1), the ZnGa2O4/GOD coating exhibited obvious oxidation and reduction peaks of GOD (Supplementary Figs. 2 and 3), in part due to ZnGa2O4-induced electron transfer. The electrochemical performance of ZnGa2O4/GOD coating was detected in glucose solution (Fig. 2d), the reduction peak currents gradually decreased with glucose addition, that is, the consumption of oxidized GOD provides routes for the electronic transduction occurring at the ZnGa2O4/GOD surface. The afterglow signal was demonstrated to be proportional to the reduction peak current by quantitative analysis, leading to a better understanding of the essential mechanisms associated with the afterglow signal and the electron transfer (Fig. 2e).

Mechanism of electron transfer-triggered imaging

The electrical signals during the reaction were characterized by electrochemical methods, and the voltametric peak current as a function of the scan rate was evaluated (Fig. 3a). As schematized in Fig. 3b, the peak current is proportional to the scan rate, indicating that the redox process occurring on the ZnGa2O4/GOD coating is surface adsorption-limited rather than diffusion-limited41. Thus, it is reasonable to speculate on the electron transfer between GOD and ZnGa2O4 adsorbed on the surface. The corresponding band diagrams of ZnGa2O4, glucose, and FAD, a cofactor of GOD, were calculated to investigate the electron transfer from an electron donor to an acceptor (Fig. 3c and Supplementary Figs. 47)42,43. For glucose with high electrochemical energy levels, electrons in its LUMO energy level will tunnel into ZnGa2O4 with the assistance of FAD-dependent enzymes. The Fermi level of ZnGa2O4 will increase via electron transfer and lead to the increment of electrons in the defect level by charge separation and trap**, thus eventually enhancing the afterglow emission. As a consequence, the electrons transferred from glucose to ZnGa2O4 can recombine to the electronic ground state via radiative pathways, leading to an enhanced emission that is highly consistent with the observation in afterglow images.

Fig. 3: Electron transfer between ZnGa2O4 and target molecules.
figure 3

a Voltammograms of GOD/ZnGa2O4-coated electrodes of a PBS solution at different scan rates (10–100 mV·s−1). b Extracted oxidation peak current (ip) mean values (circles) from the voltammograms shown in a plotted versus the scan rate (n = 3 independent electrodes). Error bars in all graphs represent the s.d. of the mean values. c The energetic structure of the system, as determined by density functional theory (DFT) calculations (Supplementary Figs. 36). FAD, flavin adenine dinucleotide; CB, conduction band; DL, defect level; VB, valence band; LUMO, lowest unoccupied molecular orbital; HOMO, highest occupied molecular orbital. d EPR spectrum of glucose and GOD mixture observed at 295 K for the 5,5-dimethyl-1-pyrroline-N-oxide (DMPO)/alkyl radical adducts. The dashed line is a simulated spectrum using parameters reported in the text. The inhomogeneous line widths for the spin-adduct have been adjusted to 0.4 G to take into account the broadening caused by the relaxation effect. e EPR spectrum of glucose, GOD, and ZnGa2O4 mixture observed at 295 K for the DMPO/peroxyl radical adducts. The dashed line is a simulated spectrum using parameters reported in the text. The inhomogeneous line widths for the spin-adduct have been adjusted to 0.8 G to take into account the broadening caused by the relaxation effect. f A higher alkyl radical EPR intensity is seen when the glucose is treated with GOD d. When the glucose is treated by GOD with ZnGa2O4, a higher peroxyl radical EPR intensity is observed (e).

The catalytic reaction between GOD and glucose was also characterized by electron paramagnetic resonance (EPR) spectroscopy to investigate the performance of ZnGa2O4-mediated electron transfer in this reaction (Fig. 3d, e). The AN values of DMPO/R radical adducts (Fig. 3d) and DMPO/ROO radical adducts (Fig. 3e) were 15.66 G and 13.70 G, and their AH values were 24.21 G and 10.82 G, comparable to spin adducts of DMPO/alkyl and DMPO/peroxyl, respectively44,45. The EPR spectrum of the glucose and GOD mixture displayed large numbers of alkyl radicals, while peroxyl radicals were generated with the addition of ZnGa2O4 (Fig. 3e), suggesting the transition of radical signals from alkyl radicals to peroxyl radicals after exposure of ZnGa2O4 (Supplementary Figs. 810). It can be concluded that with both ZnGa2O4 and GOD present, glucose will be oxidized to peroxyl radicals rather than alkyl radicals, further indicating that electron transfer occurs between GOD and ZnGa2O4.

Optimization of the ETTE nanoprobe

In an ideal solution environment, it was demonstrated that ZnGa2O4 can produce obvious afterglow signal, varying with the number of transferred electrons. By calculating the photoluminescence efficiency and the electron-hole recombination rate of the probe (see Supplementary Information, section ‘Electron tunneling probability calculation’), it can be concluded that the efficiency of photoluminescence is proportional to the number of electron-hole pairs of the material. Since ZnGa2O4 is a p-type semiconductor material, where the hole concentration is much higher than the electron concentration, it can be considered that the effect of hole concentration change on the radiative recombination rate is negligible. Thus, the photoluminescence intensity of ZnGa2O4 is positively correlated with its electron concentration. To optimize the electron transfer-induced luminescence efficiency of the probe, it is necessary to increase the number of transferred electrons between the target molecule and ZnGa2O4 in unit time. For tunneling through a one-dimensional barrier in a weakly coupled donor-acceptor system, the electron transfer rate kET is shown in formula (1).

$${k}_{{ET}}=\frac{2\pi }{{{\hslash }}}{\left|{T}_{{DA}}\right|}^{2}\left(F.C\right)={A}^{2}{e}^{{-R}_{{DA}}\beta }(F.C)$$
(1)

where ℏ is the reduced Planck constant, ℏ = h/2π, TDA is the electronic tunneling matrix element between donor and acceptor localized states. (F.C.) is the Franck-Condon factor which is determined by the activated or tunneling nuclear processes coupled to the transfer46. Formula (1) gives the calculation method of electron tunneling probability TAD2, where RDA is the distance between donor and acceptor, and β is the barrier height. It can be seen that the probability of electron tunneling can be greatly enhanced by reducing the barrier height and the distance between the target molecule and the probe47.

To induce electron transfer from the signaling pathway to the probe, molecular “wires” with molecular targeting ability were assembled on the surface of the material to close the distance between the target molecule and the probe. Because ferrocene structure exhibits high electrical conductivity and low energy barrier35, an aptamer-ferrocene polymer was a promising candidate for molecular wires (see Supplementary Information, section ‘Synthesis of Fe-base’, and Supplementary Figs. 1120). An increase in the number of ferrocene moieties can reduce the barrier height of the aptamer-ferrocene polymer by extending the conjugate field, and, on the other hand, lengthen the distance between the target molecule and the probe. Here, ETTE nanoprobes with different lengths of molecular wires were constructed via the phase transfer reaction to establish how the number of Fe-bases affects the electron transfer (Supplementary Table 1, Supplementary Fig. 2132). Energy level distributions of ZnGa2O4 and ferrocene moieties in the ETTE nanoprobe were calculated (Supplementary Fig. 33), confirming that electrons in the defect level of ZnGa2O4 will transfer to the ferrocene moieties and quench the afterglow signal48. It can be concluded that within the ETTE nanoprobe, the higher the electron tunneling probability, the lower the afterglow emission. Fluorescence spectrum of ETTE nanoprobe with varying Fe-base modifications was obtained and the probe with nine Fe-bases exhibited the lowest afterglow signal (Supplementary Figs. 3437). As a result, the ETTE nanoprobe with nine Fe-bases was selected as the optimized system that balances the reduction of barrier height with the retention of a close distance.

Electron transfer-triggered imaging of EGFR activity in living cells

EGFR signaling pathway activity is associated with tumor cell proliferation, angiogenesis, tumor invasion, metastasis, and inhibition of apoptosis17,18. Key molecular EGFR can self-phosphorylate for activating downstream signal pathways located in cells, such as mitogen-activated protein kinase (MAPK) and c-Jun N-terminal kinase (JNK) pathways. Hence, selective detection of the activity of EGFR-related signal pathways can help to understand the progression of tumor development. To map EGFR activity in complex biological samples, it is necessary to achieve the specific detection of target signal pathway activity in numerous active signal pathway networks. Thus, a molecular wire with target recognition was needed to satisfy the requirements. Known as a chemical antibody, the nucleic acid aptamer was a promising candidate for target recognition, and an EGFR-targeting aptamer was modified with 9 Fe-bases to form a specific molecular wire. This molecular wire was then functionalized on the surface of ZnGa2O4 to achieve specific electron transfer imaging of EGFR activity.

The imaging of EGFR signaling pathway activity was performed in A549 lung cancer cells using the modified ETTE nanoprobes as imaging agents (Supplementary Fig. 38). The cultured lung cancer cells were first preincubated with ETTE nanoprobes in PBS, then exposed to EGF to activate the EGFR signaling pathway. In EGFR-activated cells and in EGFR-inactivated cells, comparable fluorescence signals could be observed (Supplementary Figs. 3940), referring to the similar EGFR-targeting capability of ETTE nanoprobes. In the case of the afterglow signal from the EGFR-targeting ETTE nanoprobe (Fig. 4a), the pseudo-color image showed a weak afterglow signal in PBS buffer, due to the feeble life activity and the inactivated EGFR signaling pathway in the cells. After activation of the EGFR pathway, the afterglow signal displayed a significant enhancement, confirming that the probes have intensive afterglow-signal-generation efficiency in EGFR-activated cells and show high sensitivity to the EGFR activity. To simulate the complex in vivo environment, the above tests were repeated in the serum, and the enhancement of the afterglow signals could also be observed in EGFR-activated cells (Fig. 4b). In both PBS and FBS environment, the EGFR signaling pathway in cells was activated with the addition of EGF, and the intracellular afterglow signal was also enhanced obviously. Statistically, the relative signal intensity of these images was calculated. The relative intensity of the afterglow signal rose from 20 to 60% after EGF exposure under PBS condition, while 80% of relative intensity value could be obtained at EGF stimulation in the serum (Fig. 4c). These observations demonstrate that the ETTE probe could achieve electron transfer-triggered imaging for the EGFR signaling activity. EGFR and phosphorylated EGFR (P-EGFR) in A549 cells were characterized using western blot analysis. As observed in western blot results, the band for P-EGFR was clearly displayed in the EGF-treated cells (Supplementary Fig. 41), fitting well with the afterglow signal of the ETTE probe. The sensitivity of ETTE probes was evaluated in PBS and FBS environment. To measure the sensitivity of ETTE probes, EGF was progressively increased, which increased the afterglow intensity in EGFR-expressing A549 cells, yielding an EC50 (concentration for 50% maximal effect) of about 20 ng·mL−1 both in PBS and in FBS condition (Supplementary Figs. 4245). The ETTE probe achieves comparable sensitivity with other strategies such as fluorescence resonance energy transfer49 and surface-enhanced Raman scattering50. The longest timescales of the ETTE probe can reach almost 10 min, making it possible to obtain kinetic information (Supplementary Fig. 46). Furthermore, molecular level validation of the ETTE probe for EGFR dependent signal map** was performed on the EGFR-knocked out A549 cells. For EGFR-knocked out A549 cells, ETTE probes were found insensitive to the stimulation of EGF (Supplementary Figs. 4750), suggesting that the afterglow signal of the ETTE probe is EGFR dependent. To confirm that the signal of the ETTE probe is indeed EGFR specific, EGFR signaling activity in some EGFR-expressing cells, such as A549 cell, CAL-27 cell, and MDA-MB-231 cell, were tested with EGF stimulation and EGFR tyrosine kinase inhibition51,52 (Supplementary Fig. 51). As a result, stimulation of EGF was found to activate the signals of ETTE probes significantly, while the afterglow signals in EGFR-expressing cells were completely blocked by co-application of afatinib, an EGFR tyrosine kinase inhibitor, indicating specific responses of the ETTE probe53,54 (Supplementary Figs. 5255). Altogether, the results demonstrate that the afterglow signal from the ETTE probe is specific to EGFR and can be used as a reliable, non-invasive method to measure EGFR-related signaling activity in live cells. Flow cytometry has been used to analyze cellular signaling events, however, this platform always gives relatively static results and cannot map the signaling events dynamically in living cells55,56. The ETTE nanoprobe can possess a dynamic monitoring capability of the EGFR signaling pathway during cancer cell division, as depicted in Supplementary Fig. 56 and Supplementary Movie 1. These phenomena are possibly due to the electron transfer from activated EGFR to the probe (Fig. 4d). During a fluorescence emission process, charging of the probe can be achieved by light irradiation, enabling bright fluorescent signals57. Drowned by the strong fluorescent signals, the slight optical signal change caused by electron transfer in organisms can thus not be observed. For the ETTE probe, some of the electrons/holes created by optical excitation would be trapped at specific defects and enable the probe to store the excitation light32. Different from fluorescent signals, afterglow signals from the ETTE probe were produced by the slow release and radiative recombination of carriers from trap** defects. As a time-dependent decayed signal, afterglow intensity mostly relies on the stored electrons that would be affected by the electron transfer between biomolecules via electron acception/donation. As a consequence, the map** of EGFR activity in living cells can be achieved by evaluating the afterglow emission of this optimized ETTE nanoprobe.

Fig. 4: EGFR signaling pathway-related electron transfer imaging in living cells.
figure 4

All cells were incubated with ETTE nanoprobes and DAPI for imaging (Supplementary Fig. 25). Afterglow images and corresponding pseudo-color images of ETTE nanoprobes from A549 cells incubated without or with EGF in PBS (a) and in 10% fetal bovine serum (FBS) culture medium (b). Higher intensities reflect higher EGFR signaling activity. Color scale (ΔF/F0): 0–1.5. Scale bar, 20 μm. The laser of 635 nm and 365 nm were used as the excitation sources, afterglow signal from the ETTE probe was obtained at 1.0 s after excitation. The imaging experiments were repeated independently three times and similar results were obtained. c The bar graph shows the relative afterglow intensities of ETTE nanoprobes. Data are presented as the mean values ± s.d.; unpaired two-tailed Student’s t-test; n = 4 independent experiments. d Schematic of EGFR signaling pathway-related electron transfer-triggered emission of ETTE nanoprobes (not to scale).

Electron transfer-triggered imaging of EGFR activity in vivo

To further clarify the EGFR monitoring capacity of the ETTE probes in vivo, mice from the EGFR-noninhibited and EGFR-inhibited group were investigated (Fig. 5a), respectively. As a clinical used EGFR inhibitor, cetuximab can specifically bind to the EGFR with high affinity and block the normal function of the receptor58. The in vivo electron transfer-triggered imaging of EGFR was therefore performed using a murine model of lung cancer that was susceptible to cetuximab in vivo59 (see Supplementary Information, section ‘EGFR Inhibition Using Cetuximab’). The inhibition of EGFR was performed using cetuximab in tumor-bearing mice and was characterized by ETTE probe at day 0 and day 14, respectively (Supplementary Figs. 5761). Both real-time afterglow results showed that the afterglow intensity of the EGFR-inhibited group was significantly decreased after injection of cetuximab compared with the EGFR-noninhibited group (Fig. 5b, c). The correlation curve of afterglow signal intensity versus injection time showed that the relative signal intensity fluctuated above 100% for the PBS-treated group (Fig. 5d, e gray line), while it decreased to 60% for the cetuximab-treated group (Fig. 5d, e red line). For the EGFR-inhibited group, the afterglow signal was dimmer after 1 h post injection of cetuximab and reached a minimum at 2 h at the tumor site within the extended observation time window. The electron transfer-triggered images of EGFR-noninhibited and EGFR-inhibited mice suggest that the inhibition of EGFR could be vividly observed using our ETTE probe. To ensure the activity of EGFR signaling pathway, the expression of EGFR and P-EGFR in tumor tissues was characterized by traditional western blot and immunohistochemistry (IHC). It was observed in the western blot result that the band for P-EGFR was clearly displayed in the PBS-treated tumor tissues and weakened in the cetuximab-treated ones (Fig. 5f). When it came to IHC images, the PBS-treated group exhibited an extremely vivid yellow signal of P-EGFR while the cetuximab-treated group displayed a little yellow region (Fig. 5g). These observations suggested the low expression of P-EGFR and the low activity of the EGFR signaling pathways in the EGFR-inhibited group. In the electron transfer-triggered images, the afterglow signals of cetuximab-treated tumor tissues were apparently weaker than those of the PBS-treated group, revealing the inhibition of EGFR in the cetuximab-treated group (Fig. 5h and Supplementary Fig. 62). The statistical results of western blot and IHC demonstrated that the P-EGFR expression of the cetuximab-treated group was significantly lower than that of the PBS-treated group (Fig. 5i–j, Supplementary Figs. 6364), matching well with the EGFR activity evaluated by the ETTE probe in tumor tissues (Fig. 5k). Consequently, the electron transfer-triggered images are highly consistent with the western blot and IHC results, verifying the feasibility of the ETTE probe for imaging the EGFR signaling activity in vivo.

Fig. 5: EGFR signaling pathway-related electron transfer-triggered imaging in vivo during EGFR inhibition.
figure 5

a Schematic of electron transfer-triggered imaging of EGFR-inhibited mice with a single established A549 tumor. Mice with 50 mm3 subcutaneous tumors were administered with PBS or cetuximab intraperitoneally twice per week. b, c Afterglow images of A549 xenograft mice treated with PBS or cetuximab at day 0 (b) and day 14 (c). The afterglow imaging was performed after 1.0 s of white light excitation. The afterglow signal intensities are displayed as color maps, color scale: (b) 3.5 × 105–3.0 × 106, (c) 6.0 × 104–1.0 × 106. d, e Corresponding quantitative data analysis of relative total radiance in tumor area of mice shown in b (d) and c (e). Data presented as mean values ± s.d., n = 5 mice. f Western blot analysis on expression of EGFR, P-EGFR, and β-catenin protein in tumors after the final treatment. A representative image of five biologically independent samples from each group is shown. g IHC analysis of tumor sections. EGFR and P-EGFR in the tumor sections of each group were observed by IHC analysis. Scale bar: 50 μm. A representative image of five biologically independent samples from each group is shown. h Afterglow images of harvested tumors from the tumor-bearing mice treated with PBS or cetuximab after 14 days treatment. Color scale: 6.0 × 104–3.0 × 106. i, j Quantitative analysis of P-EGFR in tumor sections observed by western blot analysis (i) and IHC analysis (j). k Quantitative analysis of afterglow intensity in the tumor site. Data are presented as the mean values ± s.d.; unpaired two-tailed Student’s t-test; n = 5 mice.

Moreover, the biosafety of the ETTE nanoprobe was evaluated by measuring its tissue and blood compatibilities on female athymic BALB/c mice. Histological analysis of major organs, hematology analysis, and blood biochemical analysis were performed on healthy mice treated by intravenous injection with PBS or ETTE probe. The body weight of mice was monitored for 14 days after the injection, and no apparent body weight loss was observed (Supplementary Fig. 65). The histological analysis of major organs showed that the treatment of ETTE nanoprobe did not cause visible damage to all the tested organs after treatment (Supplementary Fig. 66), suggesting negligible organ toxicity of the probe. Furthermore, the general hematology parameters and standard blood biochemical indexes were assayed. Compared with the PBS-treated group, the ETTE probe-treated group displayed no statistically significant differences in both hematology parameters and blood biochemical indexes (Supplementary Figs. 6768), demonstrating good blood compatibility and no obvious toxicity to liver and kidney. Altogether, the ETTE nanoprobe displays satisfiable biosafety in the mice model, laying a good foundation for further in vivo application.

Map** EGFR activity for efficacy assessment in vivo

During tumor treatment, the activity of EGFR relevant signal pathways keeps changing. When a tumor grows, tumor cells need to proliferate and differentiate rapidly, EGFR-related signaling pathways are activated in the cells. During treatment, the growth of tumor cells will be retarded, leading to the decrease of EGFR activity. To investigate whether electron transfer-triggered imaging can be used as a more general strategy to probe signaling events in vivo, it was further explored whether the EGFR activity evaluated by ETTE probe can reflect the therapeutic effect of EGFR non-targeted therapeutic agents.

Using MTH1 siRNA as a non-targeted therapeutic agent60 (see Supplementary Information, section ‘siRNA Transfection’, and Supplementary Figs. 6971), whole-body EGFR activity was evaluated by ETTE probe in mice, and the afterglow signal in the tumor area was quantitatively calculated (Supplementary Figs. 72 and 73). Considering that the metabolism in each mouse is unlikely the same, the relative afterglow signal intensity of the tumor portion relative to the spinal symmetry region was also calculated. In the untreated group, the afterglow signal after treatment in the tumor site was apparently higher than the initial signal (Fig. 6a, b), while the treated group showed a significant decay (Fig. 6c, d). After treatment, the correlation curve of afterglow signal versus injection time showed that the relative signal intensity fluctuates raised from 100 (Fig. 6b gray line) to 120% (Fig. 6b red line) for the untreated group, while it decreased from about 100 (Fig. 6d gray line) to 60% (Fig. 6d red line) for the treated group, suggesting that the EGFR activity in the tumor region was significantly weakened. Compared with the untreated group, the treated mice have statistically significant low tumor volume and tumor weight (Fig. 6e, f), the same as the trend of afterglow signal (Supplementary Figs. 74 and 6g), indicating that the efficacy of treatment can be evaluated by the ETTE probe. Finally, the expression of EGFR and P-EGFR in tumor tissues was characterized by IHC to show the activity of the relevant signaling pathways. In the untreated group, both EGFR and P-EGFR were highly expressed (Fig. 6h), while the treated group showed a high EGFR content and a low P-EGFR content (Fig. 6i), revealing that most EGFR molecules stay inactivated in the treated group. The statistical results further demonstrated that the EGFR activity of the treated group was significantly lower than that of the untreated group (Fig. 6j, k), in agreement with the ETTE images (Fig. 6e, g). These results are highly consistent with the widely accepted view that the EGFR signaling pathway in tumors stays activated to satisfy the needs of proliferation, angiogenesis, invasion, and metastasis during tumor growth17,18. Overall, this ETTE imaging method may offer valuable guidelines for the map** of signaling activity in vivo and will inspire new strategies to evaluate therapeutic efficacy.

Fig. 6: EGFR signaling pathway-related electron transfer-triggered imaging in vivo during tumor treatment.
figure 6

Time-dependent whole-body fluorescence images and corresponding quantitative data analysis of the tumor portion relative to the spinal symmetry region of A549 xenograft mice after intravenous injection of ETTE nanoprobe without (a, b) and with (c, d) treatment with MTH1 siRNA as a therapeutic agent for 14 days and before and after treatment. Color scale: 2.5 × 103–1.5 × 104. Data presented as mean values ± s.d., n = 3 mice. During the treatment, the afterglow images of mice treated with PBS or MTH1 siRNA were obtained at day 0 (Before treatment) and day 14 (After treatment). The afterglow imaging was performed after 1.0 s of white light excitation. The afterglow signal intensities are displayed as color maps. The white circles indicated the location of the tumor. The in vivo afterglow imaging was performed with triplicates. e A representative afterglow image of harvested tumor from the tumor-bearing mouse with ETTE nanoprobes, showing that the afterglow signal mostly appeared in large tumors. Color scale: 1.0 × 104–5.0 × 104. f The average tumor weight of each group at the experimental endpoint. Tumors were resected after the initial treatment. Data presented as mean values ± s.d., n = 4 mice. g Quantitative analysis of afterglow intensity in the tumor site. Data presented as mean values ± s.d., n = 4 mice. h, i IHC analysis of primary tumor sections. EGFR (h) and P-EGFR (i) in the tumor sections of each group were observed by IHC analysis. Scale bar: 50 μm. Quantitative analysis of EGFR (j) and P-EGFR (k) in tumor site observed by IHC analysis. Data are presented as the mean values ± s.d.; unpaired two-tailed Student’s t-test; n = 3 mice.