Open Access Article
Juanhua
Zhou
and
Yang
Liu
*
Department of Chemistry, Beijing Key Laboratory for Analytical Methods and Instrumentation, Key Lab of Bioorganic Phosphorus Chemistry and Chemical Biology of Ministry of Education, Tsinghua University, Beijing, 100084, China. E-mail: liu-yang@mail.tsinghua.edu.cn
First published on 20th August 2024
Monitoring metabolites in situ at the single-cell scale is important for revealing cellular heterogeneity and dynamic changes of cell status, which provides new possibilities for disease research. Benefiting from the advantages of both electrochemical and optical methods, electrochemiluminescence (ECL) has great potential in this field. However, developing real-time in situ imaging methods is full of challenges. In this study, an ECL imaging method for formaldehyde (FA), a kind of cellular metabolite, was developed based on the in situ generation of co-reactants at the electrode interface and was successfully applied to the monitoring of single-cell FA release. Amino groups can undergo a rapid nucleophilic addition reaction with FA to form amino alcohol intermediates, which can be used as co-reactants for tris(2,2′-bipyridyl)ruthenium(II) [Ru(bpy)32+] to significantly enhance the strength of ECL. Poly(amidoamine) (PAMAM), with a large number of amino groups, and reduced graphene oxide (rGO), with excellent electrical conductivity and electrocatalytic properties, were introduced as the modification layer on the electrode surface to realize the “turn on” detection of FA. This sensing method also eliminated the use of the classic toxic co-reactant tripropylamine (TPrA) and was further applied to in situ imaging of single-cell FA release. It successfully obtained ECL images at different time points after the stimulation of HeLa cells with thapsigargin (TG), revealing the change pattern in drug efficacy over time. This work proposes a new ECL imaging approach for real-time in situ monitoring of FA release from single cells, further broadening the application of ECL imaging in single-cell analysis.
To date, various highly sensitive FA sensing methods have been developed, including chromatography,13 electrochemistry,14 and piezoresistive15 or semiconductor sensors.16 However, due to the inherent difficulties in obtaining location information, these methods still face significant challenges in real-time in situ monitoring of dynamic changes in extracellular FA content. In recent years, optical imaging methods have been widely used to monitor dynamic changes in signalling molecules in vivo due to their ability to obtain real-time in situ information. Among them, various fluorescence imaging methods for FA analysis in live samples have been reported.17–19 Although fluorescence imaging has advantages in spatial resolution, its inherent need for external light sources brings problems such as photobleaching and background interference, limiting its further development.20 Electrochemiluminescence (ECL), as a combination of electrochemical and optical methods, avoids the background light interference caused by the spontaneous bioluminescence from external light sources in fluorescence methods, showing great potential in biological analysis.21–25 Recently, there have been some reports on ECL imaging of metabolic molecules at the single-cell level.26 For example, Jiang's team was the first to develop an ECL imaging method for hydrogen peroxide efflux at the single-cell scale.27 To further improve the sensitivity of ECL imaging methods, various single-cell extracellular hydrogen peroxide ECL imaging methods based on chitosan-modified electrodes or single nanoparticle modified electrodes have also been reported successively.28–30 In addition to continually improving the sensitivity of methods, the analytical scope of ECL imaging is expanding. For example, Wang et al. developed a confined ECL imaging microarray for high-throughput sensing of dopamine released from single cells.31 Despite extensive research in the field of ECL imaging of single-cell metabolites, ECL imaging methods for single-cell FA release remain to be developed. Moreover, tris(2,2′-bipyridyl)ruthenium(II) (Ru(bpy)32+) is one of the most widely used luminophores due to its electrochemical reversibility, high quantum yield and high stability. However, the volatility and toxicity of its classic co-reactant tri-n-propylamine (TPrA) have limited the application of this ECL system to cellular process research.32 Therefore, the use of biofriendly co-reactants helps to improve the biocompatibility of this ECL system to expand its application field.33,34 Consequently, it is necessary to develop a biocompatible ECL imaging strategy to achieve real-time in-situ monitoring of FA release from single cells.
To solve the above problems, an ECL imaging method was designed to convert analyte FA into biocompatible co-reactants based on the in situ reaction at the electrode interface. FA is a reactive carbonyl compound that has been reported to undergo a rapid nucleophilic addition reaction with primary amine groups to form amino alcohol intermediates.35,36 The amino alcohol intermediates can be oxidized by electrogenerated Ru(bpy)33+ to produce stronger ECL emission.37 Poly(amidoamine) (PAMAM) dendrimers are widely used in the field of biomedicine owing to their high biocompatibility and multifunctionality.38,39 Because of the abundant primary amine groups on the terminals of the polymer, they are suitable for modification on the electrode surface to provide numerous reaction sites for FA molecules. Simultaneously, modifying reduced graphene oxide (rGO) with high conductivity, transparency, and excellent electrocatalytic performance on an ITO electrode can provide a highly transparent catalytic electrode interface.40,41 On this basis, this work developed an in-situ electrode interface reaction enhanced ECL sensing method for FA based on a PAMAM/rGO composite nanomaterial modified ITO electrode interface, as demonstrated in Scheme 1. By in situ conversion of metabolites to co-reactants on the electrode surface, it can effectively capture how to fix metabolites to reduce their free diffusion in the extracellular space, which contributes to visualizing the spatial distribution of FA. At the same time, the method also avoided the interference of additional toxic co-reactants on cells, and was applied to the dynamic monitoring of FA secretion by HeLa cells after thapsigargin (TG) stimulation, which further expanded the application of ECL imaging in the field of pathology research.
![]() | ||
| Scheme 1 Schematic illustration of the ECL imaging setup and mechanism of the ECL visualization of FA release from HeLa cells by TG stimulation. | ||
For ECL imaging, the ECL imaging system was assembled with an inverted microscope from Olympus, an electron-multiplying CCD camera, an objective (NA 0.5, 50×, Olympus, Japan) and a translation stage to adjust the sample position. A three-electrode system was used, including the modified ITO electrode as the WE, a Pt wire as the AE and an Ag/AgCl electrode as the RE. The ECL experiments were carried out in PBS (0.1 M, pH 7.4) containing 1 mM Ru(bpy)32+, with the potential applied to the electrodes using an SP-150e potentiostat electrochemical workstation (BioLogic, France). The ECL image was analysed using ImageJ software.
O stretching, and C–O stretching, respectively.44,45 The Raman spectrum in Fig. S1D† showed two bands at 1350 cm−1 (D band) and 1590 cm−1 (G band), respectively, attributed to the first-order scattering of the E2g phonon from sp2 carbon bonds and structural defects. The rGO obtained by chemical reduction of GO exhibited a layered morphology similar to GO (Fig. S1C†), and its Raman spectrum also displayed two characteristic bands, the D band and G band, caused by sp2 carbon bonds, as shown in Fig. S1D.† However, the significant reduction or even disappearance of the peaks at 3401, 1722, and 1024 cm−1 in the FTIR spectrum of rGO indicated the disappearance of OH bonds, C
O bonds, and C–O bonds, implying the successful synthesis of rGO.43
After sequentially modifying rGO and PAMAM on the surface of the ITO electrode using the drop-casting method, TEM was used to characterize the morphology of the PAMAM/rGO composite layer, as shown in Fig. 1A. The morphology was similar to that of rGO, indicating that the composite of PAMAM did not significantly change the morphology of rGO. In situ FTIR and Raman spectra were obtained to demonstrate the successful modification of PAMAM/rGO on the electrode surface. Raman spectroscopy was performed on bare ITO, rGO/ITO, and PAMAM/rGO/ITO to demonstrate the presence of rGO on ITO, as shown in Fig. 1C. The D and G bands of PAMAM/rGO/ITO indicated successful modification of rGO. To further evaluate the presence of PAMAM on the modified electrode, in situ FTIR was performed on PAMAM/rGO/ITO, rGO/ITO and PAMAM/ITO. As depicted in Fig. 1B, ITO modified with only rGO did not exhibit significant characteristic peaks in the infrared spectrum, while PAMAM/ITO and PAMAM/rGO/ITO both displayed several peaks at 3270, 2916, 2842, 1643, 1548, 1463, and 1191 cm−1. The peak near 3270 cm−1 can be attributed to NH and NH2 stretching, while the peak around 1548 cm−1 is related to N–H bending vibrations. In addition, the peak at 1191 cm−1 is associated with C–N stretching vibrations. The peak at 1643 cm−1 originates from C
O stretching vibrations, and the two peaks at 2916 and 2842 cm−1 are attributed to C–H vibrations. All the above peaks indicate the presence of NH2, C
O, and –CH2– groups, consistent with the structure of PAMAM,46,47 indicating the presence of PAMAM on the electrode surface. In summary, these results demonstrate the successful modification of the PAMAM/rGO nanomaterial layer on the electrode surface.
To further characterize the electrochemical performance of PAMAM/rGO/ITO, its conductivity was tested in a 0.1 M KCl solution containing 5.0 mM [Fe(CN)6]3−/4−via electrochemical impedance spectroscopy (EIS) and cyclic voltammetry (CV), as illustrated in Fig. S2.† Fig. S2A† provides the Nyquist plots of the impedance spectra for bare ITO, rGO/ITO, and PAMAM/rGO/ITO. As shown in Fig. S2A,† the semicircle diameter from rGO/ITO (blue line) was smaller than that from ITO (black line), while it was nearly identical to that from PAMAM/rGO/ITO (red line), suggesting an enhancement in conductivity owing to the modification of rGO.
To further investigate the mechanism by which FA significantly enhances ECL at the PAMAM/rGO/ITO electrode interface, the CV and ECL curves of bare ITO, rGO/ITO, and PAMAM/rGO/ITO were plotted in 10 mM PBS solution containing 1 mM Ru(bpy)32+, with and without FA, as shown in Fig. 2A and B, and S4.† By comparing Fig. 2A and B, it can be observed that the peak potential of ECL matches that of CV, corresponding to the oxidation potential of Ru(bpy)32+, indicating that ECL emission originates from Ru(bpy)32+. Fig. S4† indicates that the introduction of FA enhances the ECL of Ru(bpy)32+ on bare ITO, rGO/ITO, or PAMAM/rGO/ITO, which can be attributed to FA directly acting as a co-reactant of Ru(bpy)32+, consistent with previous reports.48 However, as seen in Fig. 2A, in the presence of FA, the ECL intensity on the PAMAM/rGO/ITO electrode was significantly higher than that on bare ITO and rGO/ITO, implying an additional ECL enhancement mechanism at the PAMAM/rGO/ITO electrode interface, where PAMAM plays a crucial role in ECL enhancement. Considering that the aldehyde group in FA can easily undergo a nucleophilic addition reaction with the amine groups at the ends of PAMAM, forming amino alcohols,36 to explore the changes in the terminal amine groups of PAMAM on the electrode surface before and after the addition of FA, Raman and IR spectra of the electrode surface were tested, as shown in Fig. 2C and D. In Fig. 2C, the Raman spectrum of PAMAM (black line) shows peaks at 1334 cm−1 and 3400 cm−1, corresponding to the deformation and stretching vibrations of saturated amine –NH2, 1443 cm−1 corresponding to –CH2– bending vibration, and 1672 cm−1 corresponding to C
O. Additionally, peaks near 2860 cm−1 and 3030 cm−1 are related to C–H stretching vibrations, confirming the presence of carbonyl, amine, and methylene groups in PAMAM.49,50 After the reaction with FA (red line), peaks at 1400 cm−1 and 1450 cm−1 represent secondary/tertiary amine vibrations, 1650 cm−1 corresponds to C–N stretching, and 2830 cm−1 and 2944 cm−1 correspond to C–H stretching, indicating the presence of methyl and methylene. Notably, the disappearance of the peak near 3400 cm−1 in the post-reaction Raman spectrum suggests the absence of –NH2, indicating the disappearance of the terminal primary amine group. Furthermore, the peak near 3280 cm−1 in Fig. 2D can be correlated with –OH out-of-plane bending, suggesting the transformation of the terminal amine groups of PAMAM to amino alcohols, thereby indicating a significant role of amino alcohols in the enhancement of ECL. Considering previous studies indicating that amino alcohols can be oxidized by Ru(bpy)33+ (produced by the electrooxidation of Ru(bpy)32+),37 which then participates in the ECL reaction, the potential enhancement mechanism is assumed to be as follows (Scheme 2).
In order to further evaluate the FA sensing performance of the electrode, the ECL stability at the electrode interface was studied. As shown in Fig. 3A, at the potential switching between 1.3 V and 0 V, stable ECL emission was observed in 1× PBS containing 1 mM Ru(bpy)32+ and 10 mM FA using PAMAM/rGO/ITO, with a RSD of 0.44%, indicating the excellent stability of the ECL biosensor and potential for further application in ECL detection and imaging of FA.
Additionally, sensitivity is an important indicator of FA sensing. To investigate the sensitivity of the biosensor to FA, ECL signals were recorded at various FA concentrations ranging from 0.01 mM to 20 mM. As shown in Fig. 3B, the ECL intensity increased with the increase of FA concentration. Meanwhile, Fig. 3C shows that the linear range of the FA biosensor is 0.01–3 mM. When the FA concentration exceeds 3 mM, the enhancement of ECL intensity slowed down. And when the FA concentration is greater than 10 mM, the change in ECL intensity tended to be flat. The detection limit of the method was calculated to be 0.002 mM, and the linear regression equation is I = 1524.7c(FA) + 562.9, indicating the potential applicability of this sensing approach in biological samples. Furthermore, the selectivity of the sensor was also investigated, and the ECL intensity of the PAMAM/rGO/ITO electrode was tested in the presence of a series of substances that may exist intracellularly and extracellularly, such as cations (K+, Ca2+, Mg2+), anions (NO3−, Cl−, SO42−), amino acids (glycine, glutathione), other ketone compounds (glutaraldehyde, acetone, cyclohexanone) and other reagents (reactive oxygen species, dimethyl sulfoxide, etc.). As shown in Fig. 3D, the impact of these interferences can be ignored, thus demonstrating the satisfactory specificity of the sensor. Also, this sensing interface was applied to the detection of FA in serum samples. After diluting the serum to 0.1% (v/v) with 1× PBS, different concentrations of FA were added, and the measured FA concentration was calculated using the linear regression equation. As shown in Table S1,† within the concentration range of 0.05–0.5 mM, the recovery rate of FA in diluted serum samples ranged from 94% to 101%, demonstrating the sensing ability of this interface in complex environments. In summary, the above results indicate that this sensing interface has the potential to be further applied in in situ imaging methods for monitoring single-cell release of FA.
To more intuitively compare the change in brightness over time between different single cells, Fig. 5A shows the distribution of relative gray values along the white line in Fig. 4C2–F2. In Fig. 5A, the regions of 10–40 μm and 40–75 μm correspond to the areas where cell 1 and cell 2 adhere to the electrode, respectively. From Fig. 5A, it can be seen that there was no significant difference in the relative gray values between the two cell regions at the 2nd and 10th min of drug stimulation. However, when the drug action time reached 20th min or longer, the increase in relative gray values in both cell regions indicates that the cells release FA. Moreover, comparing the cell 1 region (10–40 μm) and the cell 2 region (40–75 μm) after 20 minutes of drug action, the brightness of the cell 1 region was slightly higher than that of the cell 2 region, indicating that different cells have different responses to the same concentration of drug stimulation.
![]() | ||
| Fig. 5 (A) The relative grayscale value across the white line in Fig. 4C2–F2. (B) The trend of the relative grayscale value of individual cells over time after stimulation with TG of different concentrations. | ||
Furthermore, to evaluate the stimulating effect of different concentrations of TG on individual cells, Fig. 5B shows the trend of relative gray values per unit area of a single cell area over 30 minutes of stimulation at different drug concentrations (5 μM, 15 μM and 30 μM). From Fig. 5B, it can be seen that the stimulation of cells with the three concentrations of TG led to an increasing trend in the gray value of individual cells over time, indicating that the outflow of FA increases with the duration of TG action. However, the release amount of FA within 30 minutes did not show a significant increasing trend with higher concentrations of TG. In order to further demonstrate the correlation between the increase in grayscale values of cell regions in Fig. 4C2–F2 and the release of extracellular FA, a control experiment was set up. Specifically, DMSO was used instead of TG to image non-stimulated cells, and the results are shown in Fig. S7.† Fig. S7† demonstrates that in the absence of TG stimulation, the gray values of the cell regions did not increase over time. The correlation between luminescence and FA release indicates that this sensing interface has successfully visualized the extracellular release of FA from individual cells, confirming the feasibility of using ECL imaging to monitor drug efficacy. At present, TG is widely used to induce the production of endoplasmic reticulum stress in live cells to study the pathogenesis of various metabolic diseases. However, a standard model for endoplasmic reticulum stress for different application targets has yet to be established. The above results indicate that the ECL imaging method is expected to become a potential tool for rapid screening of stimulus drug concentrations and action times, and also has great potential in drug molecule screening.
Footnote |
| † Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4sd00177j |
| This journal is © The Royal Society of Chemistry 2024 |