Next Article in Journal
CO2 Capture and Release in Amine Solutions: To What Extent Can Molecular Simulations Help Understand the Trends?
Next Article in Special Issue
Strategies and Applications of Graphene and Its Derivatives-Based Electrochemical Sensors in Cancer Diagnosis
Previous Article in Journal
The Protective Effects of Water-Soluble Alginic Acid on the N-Terminal of Thymopentin
Previous Article in Special Issue
Facile Synthesis of Iron and Nitrogen Co-Doped Carbon Dot Nanozyme as Highly Efficient Peroxidase Mimics for Visualized Detection of Metabolites
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

N-Doped Graphene Quantum Dots Confined within Silica Nanochannels for Enhanced Electrochemical Detection of Doxorubicin

1
Key Laboratory of Surface & Interface Science of Polymer Materials of Zhejiang Province, Department of Chemistry, Zhejiang Sci-Tech University, Hangzhou 310018, China
2
The First Affiliated Hospital of Guangxi University of Chinese Medicine, Nanning 530200, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Molecules 2023, 28(18), 6443; https://doi.org/10.3390/molecules28186443
Submission received: 3 August 2023 / Revised: 29 August 2023 / Accepted: 1 September 2023 / Published: 5 September 2023

Abstract

:
Herein, we describe a fast and highly sensitive electrochemical sensor for doxorubicin (DOX) detection based on the indium tin oxide (ITO) modified with a binary material consisting of vertically-ordered mesoporous silica films (VMSFs) and N-doped graphene quantum dots (NGQDs). VMSFs, with high permeability and efficient molecular transport capacity, is attached to the ITO electrode via a rapid and controllable electrochemical method, which can serve as a solid template for the confinement of numerous NGQDs through facile electrophoresis. By virtue of the excellent charge transfer capacity, π-π and electrostatic preconcentration effects of NGQDs, as well as the electrostatic enrichment ability of VMSF, the presented NGQDs@VMSF/ITO shows amplified electrochemical signal towards DOX with a positive charge, resulting in good analytical performance in terms of a wide linear range (5 nM~0.1 μM and 0.1~1 μM), high sensitivity (30.4 μA μM−1), and a low limit of detection (0.5 nM). Moreover, due to the molecular sieving property of VMSF, the developed NGQDs@VMSF/ITO sensor has good selectivity and works well in human serum and urine samples, with recoveries of 97.0~109%, thus providing a simple and reliable method for the direct electrochemical analysis of DOX without complex sample pretreatment procedures.

Graphical Abstract

1. Introduction

Doxorubicin (DOX) is a widely used anthracycline anticancer drug that is effective in treating breast, lung, liver, and ovarian cancer [1,2,3]. Its mechanism of action involves inhibiting DNA replication and transcription in cancer cells [4,5]. However, DOX has several adverse effects on the human body, including nausea, vomiting, liver failure, local tissue necrosis, and potential heart failure [6,7]. Therefore, it is crucial to closely monitor the dosage of DOX in human biological fluids and further regulate its concentration. Various methods have been developed to quantify DOX in biological samples, such as high-performance liquid chromatography [8], fluorometry [9,10], electrophoresis [11], and electrochemical sensors [12]. Among these methods, electrochemical techniques offer a simple, sensitive, and cost-effective approach without the need for a professional operator or large-scale instrument, which has been widely used for trace analysis of DOX in biological samples [12,13]. DOX bearing hydroquinone and quinone moieties can give rise to the redox signals at the electrode surface, resulting in the design of various electrochemical detection strategies using many different nanomaterials [14,15]. However, the presence of biomolecules in biological fluids and other electrochemically active compounds often interferes with the effective electron transfer between DOX and electrode, compromising the sensitivity, repeatability, and reliability of electroanalytical sensors.
Nanostructured materials with their unique features (high surface area and nanoconfinement effects) have received a great deal of interest compared to the bulk counterparts. [16,17,18,19]. To date, vertically-ordered mesoporous silica films (VMSFs) have gained increasing attention due to their unique permselectivity and anti-fouling abilities in complex samples [20,21,22,23]. Benefiting from perpendicularly ordered channel structure, the ultrasmall and adjustable pore size (2~11 nm), as well as the high density of silanol groups on the walls, VMSFs can serve as a preconcentrated material on the electrode surface for analytes of interest through various interactions (e.g., electrostatic, size exclusion, and lipophilicity) between analytes or probe and silica nanochannels, effectively increasing the analytical performance [24,25,26]. Owing to the uniform pore size, as well as large specific surface area and good mechanical stability, VMSFs have emerged as superior protective layer on the electrode surface and simultaneously allows the effective diffusion of small molecules (analytes or probes), greatly minimizing contamination of the complex matrix on the electrode and maintaining the electrode performance [27,28,29,30,31]. Moreover, VMSFs with many tiny nanochannels can confine various nanomaterials, such as metals [32,33,34], polymers [35], graphene quantum dots (GQDs) [36], and electrochemical probes [37,38,39], to fabricate a functional sensing interface. These exceptional characteristics of VMSFs provide a universal anti-fouling electroanalytical platform for direct determination in complicated real matrix [40,41].
GQDs are a kind of zero-dimensional materials and composed of small graphene fragments, and have been widely employed for various fluorescent or colorimetric sensors due to their unique characteristics in terms of large specific surface area, good water solubility, excellent electron mobility, and high biocompatibility [42,43,44,45,46]. Since metal ions can induce the aggregation of GQDs through the coordination interaction between metal ions and functional groups of GQDs, a great deal of fluorescent sensors based on GQDs have been reported for the detection of Hg2+, Cu2+, and Fe3+ using fluorescence quenching responses [47,48,49]. In addition, GQDs have been utilized to modify the electrode surface for enhanced electrochemical detection. As our group reported previously, GQDs can be confined in an ultrasmall space of perpendicularly aligned silica nanochannels through electrophoresis and display ultrasensitive and rapid detection for metal ions and dopamine [36].
In this work, we report a simple and highly sensitive electrochemical method for DOX detection in human serum and urine samples based on the indium tin oxide (ITO) electrode modified with VMSF confining N-doped graphene quantum dots (NGQDs) into the silica nanochannels. VMSFs, possessing high permeability and efficient molecular transport capacity, can act as a solid template for the physical confinement of NGQDs via facile electrophoresis. The achieved NGQDs@VMSF/ITO sensor combines the advantages of VMSF and NGQDs (Scheme 1): (1) electrostatic preconcentration effects of NGQDs and VMSF, as well as the π-π interaction of NGQDs for the enrichment of DOX with positive charge to the electrode surface; (2) excellent charge transfer capacity of NGQDs for facilitating the electron transfer between DOX and the underlying electrode; (3) tiny spaces in nanochannels of VMSF for amplifying the electrochemical signals of DOX. The constructed NGQDs@VMSF/ITO sensor shows good analytical performance for DOX determination. Furthermore, thanks to the molecular sieving property of VMSF and cost-effective electrode fabrication, the designed NGQDs@VMSF/ITO sensor enables the direct analysis of DOX in a complex matrix (human serum and urine), making it a promising candidate for reliably monitoring DOX content in clinical diagnosis.

2. Results and Discussion

2.1. Characterization of NGQDs

NGQDs used in this work were synthesized by a one-step hydrothermal method, which was firstly studied using transmission electron microscopy (TEM) and X-ray photoelectron spectroscopy (XPS). As shown in Figure 1a,b, NGQDs have a uniform size of approximately 1.5 nm on average. Their lattice spacing is 0.23 nm, which is assigned to the graphene (100) plane (inset of Figure 1a). The survey XPS and high-resolution N 1s spectra of NGQD show three distinct peaks of C 1s, N 1s, and O 1s (Figure 1c,d), confirming the presence of amino and hydroxyl groups.

2.2. Characterization of VMSF/ITO before and after Physical Confinement of NGQDs

VMSF grown on the ITO surface was characterized by TEM, showing a uniform pore size of 2~3 nm and perpendicularly ordered structure with a thickness of 96 nm (Figure 2a,b). The cross-section of VMSF contains three layers from top to bottom, namely VMSF, ITO, and glass, which is also confirmed by scanning electron microscopy (SEM) (Figure 2c). The size of NGQDs (~2.0 nm) is similar to the pore size of the silica nanochannels, allowing the confinement of NGQDs into the nanochannels of VMSF to form NGQDs@VMSF/ITO. XPS and energy dispersive spectrometer (EDS) analyses further prove the confinement of NGQDs in the nanochannels through the detection of the characteristic peak of nitrogen element (Figure 3a and Figure S1). In addition, electrochemical technique was used to investigate the as-prepared NGQDs@VMSF/ITO electrode. In the cyclic voltammetry (CV) curve of bare ITO, displayed in Figure 3b,c, well-defined reversible redox peaks are observed, which is assigned to the electrochemical reaction of K3[Fe(CN)6] and Ru(NH3)6]Cl3 probes, respectively. Interestingly, the VMSF/ITO electrode exhibits enhanced electrochemical signal for cationic Ru(NH3)6]Cl3 but decreased the signal for anionic K3[Fe(CN)6] compared with bare ITO. This case indicates that VMSF is a good electrode-modified material for the detection of positively charged analytes because of the negative charges of silanol groups on the silica walls and apparent electrostatic effect in the ultrasmall pore size of VMSF in the measured experimental conditions. After the confinement of NGQDs into the nanochannels, NGQDs@VMSF/ITO exhibits further remarkable electrostatic permselectivity, which is attributed to the negative charge of NGQDs in the experimental conditions [29], and also indicates the successful preparation of the NGQDs@VMSF/ITO sensor.

2.3. Enhanced Electrochemical Response of DOX on NGQDs@VMSF/ITO

The electrochemical behavior of DOX during potential scan from −0.2 to −0.9 V at the bare ITO, VMSF/ITO, and NGQDs@VMSF/ITO electrodes was investigated, respectively, and the results were recorded in Figure 4. These three types of electrode exhibit no Faradic signals in blank PBS (inset of Figure 4a), but show an apparent irreversible cathodic peak for DOX (Figure 4a), indicating the occurrence of the electrochemical reduction process of DOX at the electrode surface. The cathodic peak current of DOX at the VMSF/ITO electrode is significantly higher than that at the bare ITO electrode, which is thanks to the electrostatic enrichment of DOX (pKa = 8.22) by the negatively charged properties of VMSF. Although the insulating property, VMSF remains highly permeable for the accessible diffusion of DOX due to the high porosity and ultrathin features. In addition, NGQDs@VMSF/ITO gives more sensitive cathodic peak current for DOX and the obtained current value is approximately three times higher than that obtained at the VMSF/ITO electrode (Figure 4b), which is due to the proconcentration effect in terms of electrostatic attraction and π-π interaction between NGQDs and DOX, as well as the enhanced conductivity and electron mobility of NGQDs. Therefore, the fabricated NGQDs@VMSF/ITO electrode combines the advantages of VMSF and NGQDs, displaying a good electrochemical response for the detection of DOX.

2.4. Optimized Conditions for DOX Determination

To achieve excellent performance for DOX analysis, detection conditions, including electrophoresis time for fabrication of NGQDs@VMSF/ITO, enrichment time, and pH of the supporting electrolyte, were optimized. Figure 5a shows the cathodic current responses of 1 μM DOX at the NGQDs@VMSF/ITO electrodes prepared with different electrophoresis times. In the range of 5~10 min, the obtained current value increases with increasing electrophoresis time. When the electrophoresis time is 10 min, the obtained signal value is the largest and the current response gradually decreases after 10 min. This reduced signal under longer electrophoresis time may be due because there are too many NGQDs in the nanochannel to compete with DOX in the space. Therefore, 10 min is selected as the optimal electrophoresis times for the fabrication of NGQDs@VMSF/ITO in this study. As shown in Figure 5b, the cathodic peak current increases with the enrichment time, finally reaching the equilibrium at 120 s. Thus, 120 s is selected as the optimal preconcentration time in the following experiments. The pH of PBS was also optimized; it can be seen from Figure 5c that the magnitude of cathodic peak current increases gradually with the increasing pH value at first and is maximal at pH 7.0, which is due to the pronounced negative charges on the silica wall. However, the obtained current value decreases slightly at pH 8.0, which is attributed to the instability of VMSF in an alkaline environment. Therefore, pH 7 is the optimal condition for the quantification of DOX. In addition, Figure 5d reveals the LSV responses of the NGQDs@VMSF/ITO electrode towards 1 μM DOX at different pH values of 0.01 M PBS in the range from 4.0 to 8.0, yielding a linear relationship between the pH values of PBS and cathodic peak potential (Epc) (inset of Figure 5d). Epc gradually shifts negatively with an increase in pH and displays a good linear relationship with pH (Epc = −0.0541pH − 0.247, R2 = 0.994). The slope (ΔEpc/ΔpH) of the linear regression equation was −54.1 mV/pH, demonstrating that the number of protons and electrons involved in the electrochemical reaction of DOX at the NGQDs@VMSF/ITO electrode was equal.

2.5. Electrochemical Detection of DOX Using NGQDs@VMSF/ITO

Figure 6a illustrates the LSV responses of DOX with different concentrations at the NGQDs@VMSF/ITO electrode in 0.01 M PBS. The results show that the cathodic peak current gradually increases with the addition of DOX in the range from 5 nM to 0.1 μM. As presented in Figure 6b, the as-prepared NGQDs@VMSF/ITO sensor demonstrates a good linear relationship between the cathodic peak current (I, μA) and DOX concentration (CDOX, μM), with two linear ranges, namely 5 nM to 0.1 μM and 0.1 to 1 μM. The resulting linear fitting equations in the low and high concentration ranges are I (μA) = 30.4 CDOX (μM) −0.0530 (R2 = 0.995) and I (μA) = 10.7 CDOX (μM) −1.68 (R2 = 0.993), respectively. The calculated limit of detection (LOD) was 0.5 nM (S/N = 3). The analytical capacity of VMSF/ITO for DOX detection was also evaluated under the same experimental conditions and the results are compared in Figure 6b. Our NGQDs@VMSF/ITO sensor has several apparent advantages over VMSF/ITO in terms of linear range (5 nM~1 μM vs. 0.1~1 μM), sensitivity (30.4 μA μM−1 vs. 4.63 μA μM−1), and much lower LOD. In addition, Table 1 provides a comparison of the analytical performance of NGQDs@VMSF/ITO for DOX detection with other electrochemical sensors, showing that the developed NGQDs@VMSF/ITO sensor has higher sensitivity, wider linear range, and lower LOD. The fabrication of the NGQDs@VMSF/ITO sensor is cheap and simple, without the need of many materials, along with the disposable property, making the sensor more suitable for a range of practical application.

2.6. Selectivity and Anti-Fouling Ability of NGQDs@VMSF/ITO and Real Sample Analysis

Selectivity of the developed NGQDs@VMSF/ITO sensor for the detection of DOX was then investigated. Interfering species, including sodium ion (Na+), potassium ion (K+), glutathione (GSH), glucose (Glu), uric acid (UA) dopamine (DA), and ascorbic acid (AA), may co-exist and interfere with DOX detection in practical analysis, which are similarly mixed with DOX, respectively, and detected by the fabricated NGQDs@VMSF/ITO electrode under optimal experimental conditions. Figure 7a records the cathodic peak current variation (I/I0) of 1 μM DOX at the NGQDs@VMSF/ITO electrode with (I) or without (I0) 5 μM interfering species. It can be found that the obtained cathodic peak current of DOX in the presence of five-fold concentration of distractors is almost identical to the signal of DOX alone, demonstrating the good selectivity of NGQDs@VMSF/ITO sensor. Note that the electroactive species (AA, DA, and UA) can enter the nanochannels of VMSF but do not cause a severe impact on the determination of DOX. This is due to that these species have anodic peaks at the ITO electrode and their potentials are far from the signal of DOX in this work. Additionally, four representative simulated fouling substances—DNA (protamine), peptides (Nisin), macromolecular proteins (BSA), and polysaccharides (amylum)—were selected to evaluate the antifouling ability of NGQDs@VMSF/ITO. Figure 7b presents the ratio of the current responses of the bare ITO and NGQDs@VMSF/ITO electrodes in the presence of 1 μM DOX before (I0) and after (I) 5 min of interaction with the simulated fouling agents. It can be clearly observed that the signal at the bare ITO electrode decreases significantly in the presence of different macromolecular substances, indicating that electrode passivation severely reduced the sensitivity of the bare ITO electrode in the complex matrix. However, the current signal at the NGQDs@VMSF/ITO electrode only slightly decreases, showing the protective effect of NGQDs@VMSF against biofouling. Non-electroactive substances in fouling medium can indeed non-specifically adsorb on the most sensitive electrode surface and cause interferences with the detection of analyte. As for our developed NGQDs@VMSF/ITO sensor, similar non-specific adsorption also occurs, but the transport of DOX, which has a small size, to the underlying ITO electrode through silica nanochannels is effective, eventually guaranteeing the sensitivity of electrochemical sensor. All the above results demonstrate that the designed NGQDs@VMSF/ITO has excellent molecular sieving properties and remains an active electrode surface for the electrochemical analysis of DOX in the presence of inorganic ions, electroactive biological small molecules, and biological macromolecules. Such outstanding analytical performance of NGQDs@VMSF/ITO is attributed to the unique features of VMSF. On the one hand, VMSF, with abundant ultrasmall channels and uniform pore size (2~3 nm), can block the transport of biological macromolecules through a size-exclusion effect. On the other hand, VMSF is rich in silanol groups on the pore walls and provides an electrostatic exclusion effect for positively charged interfering substances.
Excellent anti-fouling and anti-interference effects of VMSF and amplified signal capacity of NGQDs contribute to the potential of NGQDs@VMSF/ITO in the direct analysis of real samples. Considering that real samples are more complex and produce more matrix effects, we selected human serum and urine as the real samples to examine the analytical capacity of as-prepared NGQDs@VMSF/ITO in fouling medium. The received human serum and urine samples were directly diluted samples using 0.01 M PBS (pH = 7.0) and added with several known concentrations of DOX to obtain mimic detection samples, which were further analyzed by our fabricated NGQDs@VMSF/ITO sensor using the standard addition method. By comparing the detected concentration of DOX achieved from the NGQDs@VMSF/ITO and known concentrations of added DOX, recovery can be obtained to evaluate the accuracy of NGQDs@VMSF/ITO in practical application. As listed in Table 2, satisfactory recoveries (97.0~109%) and small relative standard deviation (RSD, <4.3%) values were obtained, suggesting the developed NGQDs@VMSF/ITO sensor is accurate enough and highly feasible for detecting DOX in actual samples.

3. Materials and Methods

3.1. Chemicals

Tetraethoxysilane (TEOS, Aladdin, Shanghai, China), cetyltrimethylammonium bromide (CTAB, Aladdin, Shannghai, China), sodium nitrate (NaNO3, Hangzhou Gaojing Fine Chemical Reagent, Hangzhou, China), hydrochloric acid (HCl, Hangzhou Shuanglin Chemical Reagent, Hangzhou, China), and ethanol (Aladdin, Shanghai, China) were employed to grow VMSF on the ITO-coated glass surface. 1-Aminopyrene (Aladdin, Shanghai, China) and ammonia water (Hangzhou Gaojing Fine Chemical Reagent (China)) were used for synthesis of NGQDs. Potassium ferricyanide (K3[Fe(CN)6], Aladdin, Shanghai, China), potassium phthalate monobasic (KHP, Aladdin), and hexaammineruthenium (III) chloride (Ru(NH3)6Cl3, Macklin China, Shanghai, China) were used for electrochemical characterization of NGQDs@VMSF/ITO and VMSF/ITO electrodes. Doxorubicin (DOX), sodium phosphate monobasic dihydrate (NaH2PO4), glucose (Glu), glutathione (GSH), dopamine (DA), ascorbic acid (AA), uric acid (UA), amylum, bovine serum albumin (BSA), and acetone were received from Aladdin (Shanghai, China). Sodium chloride (NaCl) and potassium chloride (KCl) were bought from the Hangzhou Gaojing Fine Chemical Reagent (Hangzhou, China). Sodium phosphate dibasic dodecahydrate (Na2HPO4), sodium hydroxide (NaOH), nisin, and fish sperm DNA were obtained from Macklin (Shanghai, China).

3.2. Quantitative Detection of DOX Using the Developed NGQDs@VMSF/ITO Electrode

A series of DOX with known concentrations was added to the 0.01 M PBS solution (pH 7.0) and then the NGQDs@VMSF/ITO electrode was used to record the electrochemical signal of DOX under optimal experimental conditions. The enrichment procedure was performed under stirring for 120 s and LSV method was employed to record the electrochemical signals of DOX at room temperature. The cathodic peak currents displayed at −0.6 V were used for the determination of DOX.
The other experimental details about the used materials and instruments, and fabrication procedures of VMSF/ITO and NGQDs@VMSF/ITO electrodes are seen in Supplementary Materials.

4. Conclusions

In summary, we have developed a fast and facile electrochemical sensor for the quantitative detection of DOX in human serum and urine samples based on the NGQDs@VMSF/ITO sensor. VMSF bears ultrasmall nanochannels, rendering a solid template for the physical confinement of NGQDs through electrophoresis. Owing to the excellent charge transfer capacity, π-π, and electrostatic preconcentration effects of NGQDs, as well as the electrostatic enrichment and anti-fouling abilities of VMSF, our proposed NGQDs@VMSF/ITO sensor can not only display superior analytical performance for DOX detection with a high sensitivity (30.4 μA μM−1) and a rather low limit of detection (0.5 nM), but also demonstrates great potential for pretreatment-free and direct analysis of human serum and urine samples. Moreover, the fabrication of NGQDs@VMSF/ITO sensor is convenient, cost-effective, and requires less-complex modification steps, making it a promising candidate for the monitoring of DOX in biological fluids and showing potential application for portable on-site analysis in clinical diagnosis.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules28186443/s1, Figure S1: SEM and EDS characterizations of NGQDs@VMSF/ITO electrode.

Author Contributions

Investigation, C.Z. and X.Z.; data curation, C.Z. and X.Z.; writing—original draft preparation, C.Z. and X.Z.; writing—review and editing, F.Y. and J.L.; conceptualization and supervision, F.Y. and J.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (21904117), the Zhejiang Provincial Natural Science Foundation of China (LY21B050003), the Fundamental Research Funds of Zhejiang Sci-Tech University (22062310-Y).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Not applicable.

References

  1. Er, E.; Erk, N. Construction of a sensitive electrochemical sensor based on 1T-MoS2 nanosheets decorated with shape-controlled gold nanostructures for the voltammetric determination of doxorubicin. Microchim. Acta 2020, 187, 223. [Google Scholar] [CrossRef] [PubMed]
  2. Rahimi, M.; Bagheri Gh, A.; Fatemi, S.J. A new sensor consisting of bird nest-like nanostructured nickel cobaltite as the sensing element for electrochemical determination of doxorubicin. J. Electroanal. Chem. 2019, 848, 113333. [Google Scholar] [CrossRef]
  3. Ehsani, M.; Soleymani, J.; Mohammadalizadeh, P.; Hasanzadeh, M.; Jouyban, A.; Khoubnasabjafari, M.; Vaez-Gharamaleki, Y. Low potential detection of doxorubicin using a sensitive electrochemical sensor based on glassy carbon electrode modified with silver nanoparticles-supported poly(chitosan): A new platform in pharmaceutical analysis. Microchem. J. 2021, 165, 106101. [Google Scholar] [CrossRef]
  4. Shams, A.; Yari, A. A new sensor consisting of Ag-MWCNT nanocomposite as the sensing element for electrochemical determination of Epirubicin. Sens. Actuators B Chem. 2019, 286, 131–138. [Google Scholar] [CrossRef]
  5. Behravan, M.; Aghaie, H.; Giahi, M.; Maleknia, L. Determination of doxorubicin by reduced graphene oxide/gold/polypyrrole modified glassy carbon electrode: A new preparation strategy. Diam. Relat. Mater. 2021, 117, 108478. [Google Scholar] [CrossRef]
  6. Alizadeh, P.M.; Hasanzadeh, M.; Soleymani, J.; Gharamaleki, J.V.; Jouyban, A. Application of bioactive cyclic oligosaccharide on the detection of doxorubicin hydrochloride in unprocessed human plasma sample: A new platform towards efficient chemotherapy. Microchem. J. 2019, 145, 450–455. [Google Scholar] [CrossRef]
  7. Selvi, S.V.; Prasannan, A.; Chen, S.M.; Vadivelmurugan, A.; Tsai, H.C.; Lai, J.Y. Glutathione and cystamine functionalized MoS2 core-shell nanoparticles for enhanced electrochemical detection of doxorubicin. Microchim. Acta 2021, 188, 35. [Google Scholar] [CrossRef]
  8. Alhareth, K.; Vauthier, C.; Gueutin, C.; Ponchel, G.; Moussa, F. HPLC quantification of doxorubicin in plasma and tissues of rats treated with doxorubicin loaded poly(alkylcyanoacrylate) nanoparticles. J. Chromatogr. B 2012, 887, 128–132. [Google Scholar] [CrossRef]
  9. Martínez Ferreras, F.; Wolfbeis, O.S.; Gorris, H.H. Dual lifetime referenced fluorometry for the determination of doxorubicin in urine. Anal. Chim. Acta 2012, 729, 62–66. [Google Scholar] [CrossRef]
  10. El-Kimary, E.I.; El-Yazbi, A.F. An eco-friendly stability-indicating spectrofluorimetric method for the determination of two anticancer stereoisomer drugs in their pharmaceutical preparations following micellar enhancement: Application to kinetic degradation studies. Spectrochim. Acta A 2016, 163, 145–153. [Google Scholar] [CrossRef]
  11. Lu, H.; Yuan, G.; He, Q.; Chen, H. Rapid analysis of anthracycline antibiotics doxorubicin and daunorubicin by microchip capillary electrophoresis. Microchem. J. 2009, 92, 170–173. [Google Scholar] [CrossRef]
  12. Yan, F.; Chen, J.; Jin, Q.; Zhou, H.; Sailjoi, A.; Liu, J.; Tang, W. Fast one-step fabrication of a vertically-ordered mesoporous silica-nanochannel film on graphene for direct and sensitive detection of doxorubicin in human whole blood. J. Mater. Chem. C 2020, 8, 7113–7119. [Google Scholar] [CrossRef]
  13. Soleymani, J.; Hasanzadeh, M.; Eskandani, M.; Khoubnasabjafari, M.; Shadjou, N.; Jouyban, A. Electrochemical sensing of doxorubicin in unprocessed whole blood, cell lysate, and human plasma samples using thin film of poly-arginine modified glassy carbon electrode. Mater. Sci. Eng. C 2017, 77, 790–802. [Google Scholar] [CrossRef] [PubMed]
  14. Lv, X.; Ge, W.; Li, Q.; Wu, Y.; Jiang, H.; Wang, X. Rapid and ultrasensitive electrochemical detection of multidrug-resistant bacteria based on nanostructured gold coated ITO electrode. ACS Appl. Mater. Interfaces 2014, 6, 11025–11031. [Google Scholar] [CrossRef] [PubMed]
  15. Guo, H.; Jin, H.; Gui, R.; Wang, Z.; Xia, J.; Zhang, F. Electrodeposition one-step preparation of silver nanoparticles/carbon dots/reduced graphene oxide ternary dendritic nanocomposites for sensitive detection of doxorubicin. Sens. Actuators B Chem. 2017, 253, 50–57. [Google Scholar] [CrossRef]
  16. Singh, J.; Soni, R.K.; Nguyen, D.D.; Kumar Gupta, V.; Nguyen-Tri, P. Enhanced photocatalytic and SERS performance of Ag nanoparticles functionalized MoS2 nanoflakes. Chemosphere 2023, 339, 139735. [Google Scholar] [CrossRef]
  17. Singh, J.; Rishikesh; Kumar, S.; Soni, R.K. Synthesis of 3D-MoS2 nanoflowers with tunable surface area for the application in photocatalysis and SERS based sensing. J. Alloy Compd. 2020, 849, 156502. [Google Scholar] [CrossRef]
  18. Singh, J.; Manna, A.K.; Soni, R.K. Sunlight driven photocatalysis and non-enzymatic glucose sensing performance of cubic structured CuO thin films. Appl. Surf. Sci. 2020, 530, 147258. [Google Scholar] [CrossRef]
  19. Gong, J.; Tang, H.; Wang, M.; Lin, X.; Wang, K.; Liu, J. Novel three-dimensional graphene nanomesh prepared by facile electro-etching for improved electroanalytical performance for small biomolecules. Mater. Design 2022, 215, 110506. [Google Scholar] [CrossRef]
  20. Zhou, P.; Yao, L.; Chen, K.; Su, B. Silica nanochannel membranes for electrochemical analysis and molecular sieving: A comprehensive review. Crit. Rev. Anal. Chem. 2020, 50, 424–444. [Google Scholar] [CrossRef]
  21. Ma, K.; Yang, L.; Liu, J.; Liu, J. Electrochemical sensor nanoarchitectonics for sensitive detection of uric acid in human whole blood based on screen-printed carbon electrode equipped with vertically-ordered mesoporous silica-nanochannel film. Nanomaterials 2022, 12, 1157. [Google Scholar] [CrossRef] [PubMed]
  22. Deng, X.; Lin, X.; Zhou, H.; Liu, J.; Tang, H. Equipment of vertically-ordered mesoporous silica film on electrochemically pretreated three-dimensional graphene electrodes for sensitive detection of methidazine in urine. Nanomaterials 2023, 13, 239. [Google Scholar] [CrossRef] [PubMed]
  23. Zhou, H.; Ding, Y.; Su, R.; Lu, D.; Tang, H.; Xi, F. Silica nanochannel array film supported by ß-cyclodextrin-functionalized graphene modified gold film electrode for sensitive and direct electroanalysis of acetaminophen. Front. Chem. 2022, 9, 812086. [Google Scholar] [CrossRef] [PubMed]
  24. Wei, X.; Luo, X.; Xu, S.; Xi, F.; Zhao, T. A flexible electrochemiluminescence sensor equipped with vertically ordered mesoporous silica nanochannel film for sensitive detection of clindamycin. Front. Chem. 2022, 10, 872582. [Google Scholar] [CrossRef]
  25. Lv, N.; Qiu, X.; Han, Q.; Xi, F.; Wang, Y.; Chen, J. Anti-biofouling electrochemical sensor based on the binary nanocomposite of silica nanochannel array and graphene for doxorubicin detection in human serum and urine samples. Molecules 2022, 27, 8640. [Google Scholar] [CrossRef]
  26. Ma, N.; Luo, X.; Wu, W.; Liu, J. Fabrication of a disposable electrochemical immunosensor based on nanochannel array modified electrodes and gated electrochemical signals for sensitive determination of C-reactive protein. Nanomaterials 2022, 12, 3981. [Google Scholar] [CrossRef]
  27. Wang, K.; Yang, L.; Huang, H.; Lv, N.; Liu, J.; Liu, Y. Nanochannel array on electrochemically polarized screen printed carbon electrode for rapid and sensitive electrochemical determination of clozapine in human whole blood. Molecules 2022, 27, 2739. [Google Scholar] [CrossRef]
  28. Yan, Z.; Zhang, S.; Liu, J.; Xing, J. Homogeneous electrochemical aptamer sensor based on two-dimensional nanocomposite probe and nanochannel modified electrode for sensitive detection of carcinoembryonic antigen. Molecules 2023, 28, 5186. [Google Scholar] [CrossRef]
  29. Huang, J.; Zhang, T.; Zheng, Y.; Liu, J. Dual-mode sensing platform for cancer antigen 15-3 determination based on a silica nanochannel array using electrochemiluminescence and electrochemistry. Biosensors 2023, 13, 317. [Google Scholar] [CrossRef]
  30. Zhu, X.; Xuan, L.; Gong, J.; Liu, J.; Wang, X.; Xi, F.; Chen, J. Three-dimensional macroscopic graphene supported vertically-ordered mesoporous silica-nanochannel film for direct and ultrasensitive detection of uric acid in serum. Talanta 2022, 238, 123027. [Google Scholar] [CrossRef]
  31. Su, R.; Tang, H.; Xi, F. Sensitive electrochemical detection of p-nitrophenol by pre-activated glassy carbon electrode integrated with silica nanochannel array film. Front. Chem. 2022, 10, 954748. [Google Scholar] [CrossRef]
  32. Ding, L.; Li, W.; Sun, Q.; He, Y.; Su, B. Gold nanoparticles confined in vertically aligned silica nanochannels and their electrocatalytic activity toward ascorbic acid. Chem. Eur. J. 2014, 20, 12777–12780. [Google Scholar] [CrossRef] [PubMed]
  33. Li, X.; Zhou, L.; Ding, J.; Sun, L.; Su, B. Platinized silica nanoporous membrane electrodes for low-fouling hydrogen peroxide detection. ChemElectroChem 2020, 7, 2081–2086. [Google Scholar] [CrossRef]
  34. Huang, L.; Su, R.; Xi, F. Sensitive detection of noradrenaline in human whole blood based on Au nanoparticles embedded vertically-ordered silica nanochannels modified pre-activated glassy carbon electrodes. Front. Chem. 2023, 11, 1126213. [Google Scholar] [CrossRef]
  35. Ding, L.; Li, W.; Wang, Q.; Sun, Q.; He, Y.; Su, B. Vertically oriented silica mesochannels as the template for electrodeposition of polyaniline nanostructures and their electrocatalytic and electroanalytical applications. Chem. Eur. J. 2014, 20, 1829–1833. [Google Scholar] [CrossRef] [PubMed]
  36. Lu, L.; Zhou, L.; Chen, J.; Yan, F.; Liu, J.; Dong, X.; Xi, F.; Chen, P. Nanochannel-confined graphene quantum dots for ultrasensitive electrochemical analysis of complex samples. ACS Nano 2018, 12, 12673–12681. [Google Scholar] [CrossRef]
  37. Zhou, X.; Han, Q.; Zhou, J.; Liu, C.; Liu, J. Reagentless electrochemical detection of tumor biomarker based on stable confinement of electrochemical probe in bipolar silica nanochannel film. Nanomaterials 2023, 13, 1645. [Google Scholar] [CrossRef]
  38. Gong, J.; Zhang, T.; Luo, T.; Luo, X.; Yan, F.; Tang, W.; Liu, J. Bipolar silica nanochannel array confined electrochemiluminescence for ultrasensitive detection of SARS-CoV-2 antibody. Biosens. Bioelectron. 2022, 215, 114563. [Google Scholar] [CrossRef]
  39. Chen, D.; Luo, X.; Xi, F. Probe-integrated electrochemical immunosensor based on electrostatic nanocage array for reagentless and sensitive detection of tumor biomarker. Front. Chem. 2023, 11, 1121450. [Google Scholar] [CrossRef]
  40. Luo, X.; Zhang, T.; Tang, H.; Liu, J. Novel electrochemical and electrochemiluminescence dual-modality sensing platform for sensitive determination of antimicrobial peptides based on probe encapsulated liposome and nanochannel array electrode. Front. Nutr. 2022, 9, 962736. [Google Scholar] [CrossRef]
  41. Yan, L.; Xu, S.; Xi, F. Disposal immunosensor for sensitive electrochemical detection of prostate-specific antigen based on amino-rich nanochannels array-modified patterned indium tin oxide electrode. Nanomaterials 2022, 12, 3810. [Google Scholar] [CrossRef]
  42. Li, Y.; Gu, X.; Zhao, J.; Xi, F. Fabrication of a ratiometric fluorescence sensor based on carbon dots as both luminophores and nanozymes for the sensitive detection of hydrogen peroxide. Molecules 2022, 27, 7379. [Google Scholar] [CrossRef] [PubMed]
  43. Cui, Y.; Duan, W.; Jin, Y.; Wo, F.; Xi, F.; Wu, J. Ratiometric fluorescent nanohybrid for noninvasive and visual monitoring of sweat glucose. ACS Sens. 2020, 5, 2096–2105. [Google Scholar] [CrossRef]
  44. Xu, S.; Zhang, S.; Li, Y.; Liu, J. Facile synthesis of iron and nitrogen co-doped carbon dot nanozyme as highly efficient peroxidase mimics for visualized detection of metabolites. Molecules 2023, 28, 6064. [Google Scholar] [CrossRef] [PubMed]
  45. Wan, Y.; Zhao, J.; Deng, X.; Chen, J.; Xi, F.; Wang, X. Colorimetric and fluorescent dual-modality sensing platform based on fluorescent nanozyme. Front. Chem. 2021, 9, 774486. [Google Scholar] [CrossRef] [PubMed]
  46. Deng, X.; Zhao, J.; Ding, Y.; Tang, H.; Xi, F. Iron and nitrogen co-doped graphene quantum dots as highly active peroxidases for the sensitive detection of l-cysteine. New J. Chem. 2021, 45, 19056–19064. [Google Scholar] [CrossRef]
  47. Ge, S.; He, J.; Ma, C.; Liu, J.; Xi, F.; Dong, X. One-step synthesis of boron-doped graphene quantum dots for fluorescent sensors and biosensor. Talanta 2019, 199, 581–589. [Google Scholar] [CrossRef]
  48. Mao, Y.; Zhao, C.; Ge, S.; Luo, T.; Chen, J.; Liu, J.; Xi, F.; Liu, J. Gram-scale synthesis of nitrogen doped graphene quantum dots for sensitive detection of mercury ions and l-cysteine. RSC Adv. 2019, 9, 32977–32983. [Google Scholar] [CrossRef]
  49. Shen, C.; Ge, S.; Pang, Y.; Xi, F.; Liu, J.; Dong, X.; Chen, P. Facile and scalable preparation of highly luminescent N,S co-doped graphene quantum dots and their application for parallel detection of multiple metal ions. J. Mater. Chem. B 2017, 5, 6593–6600. [Google Scholar] [CrossRef]
  50. Zhao, H.; Shi, K.; Zhang, C.; Ren, J.; Cui, M.; Li, N.; Ji, X.; Wang, R. Spherical COFs decorated with gold nanoparticles and multiwalled carbon nanotubes as signal amplifier for sensitive electrochemical detection of doxorubicin. Microchem. J. 2022, 182, 107865. [Google Scholar] [CrossRef]
  51. Rezvani Jalal, N.; Madrakian, T.; Afkhami, A.; Ahmadi, M. Ni/Co bimetallic metal–organic frameworks on nitrogen-doped graphene oxide nanoribbons for electrochemical sensing of doxorubicin. ACS Appl. Nano Mater. 2022, 5, 11045–11058. [Google Scholar] [CrossRef]
  52. Rong, S.; Zou, L.; Meng, L.; Yang, X.; Dai, J.; Wu, M.; Qiu, R.; Tian, Y.; Feng, X.; Ren, X.; et al. Dual function metal-organic frameworks based ratiometric electrochemical sensor for detection of doxorubicin. Anal. Chim. Acta 2022, 1196, 339545. [Google Scholar] [CrossRef] [PubMed]
Scheme 1. Schematic diagram of the preparation process of the NGQDs@VMSF/ITO electrode (a) and highly sensitive sensing mechanism of NGQDs@VMSF/ITO electrode for DOX (b).
Scheme 1. Schematic diagram of the preparation process of the NGQDs@VMSF/ITO electrode (a) and highly sensitive sensing mechanism of NGQDs@VMSF/ITO electrode for DOX (b).
Molecules 28 06443 sch001
Figure 1. (a) TEM image of the as-prepared NGQDs. Inset is high-resolution TEM image with indicated lattice parameter and (b) size distribution (bottom) of NGQDs. (c) XPS survey spectrum of NGQDs. (d) High-resolution N 1s of the synthesized NGQDs.
Figure 1. (a) TEM image of the as-prepared NGQDs. Inset is high-resolution TEM image with indicated lattice parameter and (b) size distribution (bottom) of NGQDs. (c) XPS survey spectrum of NGQDs. (d) High-resolution N 1s of the synthesized NGQDs.
Molecules 28 06443 g001
Figure 2. Electron microscopy measurements of VMSF/ITO electrode: (a) top-view TEM, (b) cross-sectional view TEM, and (c) cross-sectional view SEM.
Figure 2. Electron microscopy measurements of VMSF/ITO electrode: (a) top-view TEM, (b) cross-sectional view TEM, and (c) cross-sectional view SEM.
Molecules 28 06443 g002
Figure 3. (a) XPS survey spectrum of NGQDs. CV responses of 0.5 mM K3[Fe(CN)6] (b) and [Ru(NH3)6]Cl3 (c) at the bare ITO (black curve), VMSF/ITO (red curve), and NGQDs@VMSF/ITO (blue curve) electrodes in 50 mM KHP solution.
Figure 3. (a) XPS survey spectrum of NGQDs. CV responses of 0.5 mM K3[Fe(CN)6] (b) and [Ru(NH3)6]Cl3 (c) at the bare ITO (black curve), VMSF/ITO (red curve), and NGQDs@VMSF/ITO (blue curve) electrodes in 50 mM KHP solution.
Molecules 28 06443 g003
Figure 4. CV (a) and LSV (b) plots of 1 μM DOX at the bare ITO, VMSF/ITO, and NGQDs@VMSF/ITO electrodes in 0.01 M PBS (pH = 7.0) solution. Inset in (a) is these three electrodes in 0.01 M PBS (pH = 7.0) solution in the absence of DOX. The scan rate for CV tests was 50 mV s–1.
Figure 4. CV (a) and LSV (b) plots of 1 μM DOX at the bare ITO, VMSF/ITO, and NGQDs@VMSF/ITO electrodes in 0.01 M PBS (pH = 7.0) solution. Inset in (a) is these three electrodes in 0.01 M PBS (pH = 7.0) solution in the absence of DOX. The scan rate for CV tests was 50 mV s–1.
Molecules 28 06443 g004
Figure 5. The cathodic current signal of 1 μM DOX at the NGQDs@VMSF/ITO electrode in 0.01 M PBS (pH = 7.0) solution under different electrophoresis time (a) and preconcentration time (b). (c,d) Effect of pH values of 0.01 M PBS on the electrochemical signal of 1 μM DOX at the NGQDs@VMSF/ITO electrode, including cathodic peak currents (c) and cathodic peak (d) in 0.01 M PBS (pH = 7.0) solution. Inset in (d) shows the dependence of cathodic peak potential on the pH value.
Figure 5. The cathodic current signal of 1 μM DOX at the NGQDs@VMSF/ITO electrode in 0.01 M PBS (pH = 7.0) solution under different electrophoresis time (a) and preconcentration time (b). (c,d) Effect of pH values of 0.01 M PBS on the electrochemical signal of 1 μM DOX at the NGQDs@VMSF/ITO electrode, including cathodic peak currents (c) and cathodic peak (d) in 0.01 M PBS (pH = 7.0) solution. Inset in (d) shows the dependence of cathodic peak potential on the pH value.
Molecules 28 06443 g005
Figure 6. (a) LSV response of DOX at the NGQDs@VMSF/ITO electrode in 0.01 M PBS solution (pH 7.0). The concentrations of DOX are 0.005 μM, 0.01 μM, 0.03 μM, 0.05 μM, 0.07 μM, 0.1 μM, 0.3 μM, 0.5 μM, 0.7 μM, and 1 μM, respectively. (b) The linear relationship between the cathode peak currents and DOX concentration obtained at both NGQDs@VMSF/ITO and VMSF/ITO electrodes. The concentrations of DOX at the VMSF/ITO electrode are 0.1 μM, 0.3 μM, 0.5 μM, 0.7 μM, and 1 μM, respectively.
Figure 6. (a) LSV response of DOX at the NGQDs@VMSF/ITO electrode in 0.01 M PBS solution (pH 7.0). The concentrations of DOX are 0.005 μM, 0.01 μM, 0.03 μM, 0.05 μM, 0.07 μM, 0.1 μM, 0.3 μM, 0.5 μM, 0.7 μM, and 1 μM, respectively. (b) The linear relationship between the cathode peak currents and DOX concentration obtained at both NGQDs@VMSF/ITO and VMSF/ITO electrodes. The concentrations of DOX at the VMSF/ITO electrode are 0.1 μM, 0.3 μM, 0.5 μM, 0.7 μM, and 1 μM, respectively.
Molecules 28 06443 g006
Figure 7. (a) The obtained current ratio (I/I0) at the NGQDs@VMSF/ITO electrode for the determination of 1 μM DOX without (I0) and with (I) the added interfering species (Na+, K+, Glu, GSH, AA, DA, and UA). The concentration of interfering species is 5 μM. (b) The peak current ratio (I/I0) on the NGQDs@VMSF/ITO (without shadow) and bare ITO (with shadow) in the presence of 1 μM DOX. I and I0 represent the cathodic peak currents obtained in the present and absence of 10 μg/mL of DNA (green), Nisin (blue), amylum (red), or BSA (yellow).
Figure 7. (a) The obtained current ratio (I/I0) at the NGQDs@VMSF/ITO electrode for the determination of 1 μM DOX without (I0) and with (I) the added interfering species (Na+, K+, Glu, GSH, AA, DA, and UA). The concentration of interfering species is 5 μM. (b) The peak current ratio (I/I0) on the NGQDs@VMSF/ITO (without shadow) and bare ITO (with shadow) in the presence of 1 μM DOX. I and I0 represent the cathodic peak currents obtained in the present and absence of 10 μg/mL of DNA (green), Nisin (blue), amylum (red), or BSA (yellow).
Molecules 28 06443 g007
Table 1. Comparison of the analytical performance of NGQDs@VMSF/ITO electrode with other reported electroanalytical methods for DOX determination.
Table 1. Comparison of the analytical performance of NGQDs@VMSF/ITO electrode with other reported electroanalytical methods for DOX determination.
Sensor MaterialMethodLinear Range (μM)Sensitivity
(μA μM1)
LOD
(nM)
Ref.
AgNPs/Chitosan/GCESWV0.103–8.60.8613103[3]
AuNPs@COFs-MWCNTs/GCEDPV0.08–3.0, 3.0–251.18516[50]
NiCo-BTC MOFs/N-GONRs/GCESWV0.01–1.0, 1.0–8095.0246[51]
MB@MWCNTs/UiO-66-NH2/GCEDPV0.1–750.0182951[52]
VMSF/ErGO/GCEDPV0.001–207.8150.77[12]
AuNRDs/1T-MoS2/SPEDPV0.01–9.500.89522.5[1]
NGQDs@VMSF/ITOLSV0.005–0.1, 0.1–130.40.5This work
AgNPs: silver nanoparticles; GCE: glassy carbon electrode; SWV: square wave voltammetry; DPV: differential pulse voltammetry; AuNPs: gold nanoparticles; MWCNTs: multi-walled carbon nanotubes; NiCo-BTC MOFs: Ni/Co bimetallic metal−organic frameworks; N-GONRs: nitrogen-doped graphene oxide nanoribbons; MB: methylene blue; ErGO: electrochemically reduced graphene oxide; AuNRDs: homogenous and uniform single-crystal gold nanorods; 1T-MoS2: metastable 1T phase molybdenum disulfide; SPE: screen-printed electrode.
Table 2. Determination of DOX in serum and urine samples.
Table 2. Determination of DOX in serum and urine samples.
Sample aAdded (μM)Found (μM)RSD (%)Recovery (%)
Serum0.1000.1010.199.0
0.5000.4991.1100
1.000.9902.699.9
Urine0.1000.1092.4109
0.5000.5143.8103
1.000.9704.397.0
a Samples with the addition of known concentrations of DOX were diluted 50 times using PBS (0.01 M, pH = 7). The concentration of DOX was the concentration after dilution.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zhang, C.; Zhou, X.; Yan, F.; Lin, J. N-Doped Graphene Quantum Dots Confined within Silica Nanochannels for Enhanced Electrochemical Detection of Doxorubicin. Molecules 2023, 28, 6443. https://doi.org/10.3390/molecules28186443

AMA Style

Zhang C, Zhou X, Yan F, Lin J. N-Doped Graphene Quantum Dots Confined within Silica Nanochannels for Enhanced Electrochemical Detection of Doxorubicin. Molecules. 2023; 28(18):6443. https://doi.org/10.3390/molecules28186443

Chicago/Turabian Style

Zhang, Chaoyan, Xiaoyu Zhou, Fei Yan, and Jing Lin. 2023. "N-Doped Graphene Quantum Dots Confined within Silica Nanochannels for Enhanced Electrochemical Detection of Doxorubicin" Molecules 28, no. 18: 6443. https://doi.org/10.3390/molecules28186443

Article Metrics

Back to TopTop