Yangguang
Zhu
abd,
Chen
Ye
cd,
Xiao
Xiao
h,
Zhuang
Sun
ad,
Xiufen
Li
b,
Li
Fu
e,
Hassan
Karimi-Maleh
fg,
Jun
Chen
*h and
Cheng-Te
Lin
*acd
aKey Laboratory of Advanced Marine Materials, Ningbo Institute of Materials Technology and Engineering (NIMTE), Chinese Academy of Sciences, Ningbo 315201, P. R. China. E-mail: linzhengde@nimte.ac.cn
bLaboratory of Environmental Biotechnology, School of Environmental and Civil Engineering, Jiangnan University, Wuxi 214122, P. R. China
cCenter of Materials Science and Optoelectronics Engineering, University of Chinese Academy of Sciences, Beijing 100049, P. R. China
dQianwan Institute, Ningbo Institute of Materials Technology and Engineering (NIMTE), Chinese Academy of Sciences, Ningbo 315201, P. R. China
eCollege of Materials and Environmental Engineering, Hangzhou Dianzi University, Hangzhou 310018, P. R. China
fSchool of Resources and Environment, University of Electronic Science and Technology of China, Chengdu 611731, P. R. China
gSchool of Engineering, Lebanese American University, Byblos 1102-2801, Lebanon
hDepartment of Bioengineering, University of California, Los Angeles, Los Angeles, CA 90095, USA. E-mail: jun.chen@ucla.edu
First published on 7th October 2024
Owing to the extensive use of antibiotics for treating infectious diseases in livestock and humans, the resulting residual antibiotics are a burden to the ecosystem and human health. Hence, for human health and ecological safety, it is critical to determine the residual antibiotics with accuracy and convenience. Graphene-based electrochemical sensors are an effective tool to detect residual antibiotics owing to their advantages, such as, high sensitivity, simplicity, and time efficiency. In this work, we comprehensively summarize the recent advances in graphene-based electrochemical sensors used for detecting antibiotics, including modifiers for electrode fabrication, theoretical elaboration of electrochemical sensing mechanisms, and practical applications of portable electrochemical platforms for the on-site monitoring of antibiotics. It is anticipated that the current review will be a valuable reference for comprehensively comprehending graphene-based electrochemical sensors and further promoting their applications in the fields of healthcare, environmental protection, and food safety.
Wider impactThe key points discussed in this review include the fabrication of graphene-based electrochemical sensors for antibiotics, the theoretical underpinnings of electrochemical sensing mechanisms, and the evolution of portable electrochemical platforms for on-site monitoring. These advancements are of significant wider interest as they not only enhance our ability to detect and quantify antibiotics in various samples but also contribute to the broader fields of environmental science, public health, and materials engineering. The future is expected to be marked by further refinement of sensor technology, with a focus on improving detection limits, selectivity, and the practicality of on-site monitoring. The insights provided in this review will be instrumental in guiding future research, particularly in understanding the complex interactions at the electrode surface and the redox processes that facilitate the detection of specific antibiotics. By addressing the questions of detection performance improvement, material modification on graphene nanosheets, the underlying sensing mechanisms, and the status of equipment platforms, this review aims to clarify the current state of graphene-based electrochemical sensors for antibiotics and their readiness for real-world applications. The potential for miniaturization and portability is a critical aspect as it will enable more widespread and accessible monitoring of antibiotic residues, thus safeguarding both ecological systems and human health. |
Conventional analytical methods, such as high-performance liquid chromatography (HPLC), liquid chromatography–mass spectrometry (LC–MS), gas chromatography–mass spectrometry (GC–MS), and electrochemical detection methods, for antibiotic detection have been under development for decades.12–15 Among them, chromatography enables the simultaneous detection of multiple antibiotics and offers advantages such as high sensitivity and strong anti-interference capability.12,13 Yet, it still faces bottlenecks, such as costly equipment, complex sample pretreatments, and high detection costs, making it suitable only for sample analysis in laboratory research. Hence, constructing electrochemical sensors has been preferably accepted due to its high sensitivity, simplicity, and time efficiency.16–18 However, direct electrochemical detection of antibiotics on conventional electrodes like glassy carbon electrodes (GCEs) usually confronts low sensitivity. Thus, adopting chemical modification with various nanomaterials on electrochemical electrodes has become a widely used strategy.19–21
Graphene is a two-dimensional nanomaterial consisting of a single layer of carbon atoms arranged in a hexagonal honeycomb lattice. Its unique atomic structure and electronic properties endow it with exceptional electrical, optical, and chemical characteristics.22 When considering its electrochemical aspects, it has been regarded as a perfect electrode modification material due to its excellent properties, such as high surface-to-volume ratio, good electrical conductivity, favourable biocompatibility and rapid electron transferring rate.18,23,24 Moreover, electrochemical signals can be further enhanced through the functionalization of graphene via covalent and non-covalent modifications.25 Recently, graphene-based electrochemical sensors for antibiotics have been extensively constructed, and the detection performance including selectivity, limit of detection (LOD) and real sample analysis have been greatly improved.26–28
As far as we know, the focus of the reported reviews is on describing the status via bibliometric analysis and highlighting the research on graphene-based nanocomposite-modified electrodes in the context of materials science and their application in determining antibiotics in real samples within a laboratory environment.29–34 However, it is inadequate for comprehensively reviewing graphene-based electrochemical sensors for antibiotics. The specific function of graphene with various materials in electrode fabrication is not elaborated. The antibiotic detection mechanism involved in electrochemical sensing is not clear yet, especially referring to the redox reaction process involving electron transferring and interfacial interactions between the electrode surface and the adsorbed molecules at the atomic level. Furthermore, whether the electrochemical sensors constructed in current stage achieve the requirement for the on-site monitoring of antibiotics in real samples is unknown.
In this review, we focus on three aspects: graphene-based modifiers for electrode fabrication, theoretical analyses of electrochemical sensing mechanisms, and portable electrochemical platforms for the on-site monitoring of antibiotics in real samples. It is anticipated that our review will be a valuable reference for comprehending the whole fabrication process of graphene-based electrochemical sensors, and further promoting the development of the sensors for practical applications in trace level detection of antibiotics.
Antibiotic class | Type | Chemical formula | Mechanism of action | Side-effects | Electrodes | Ref. |
---|---|---|---|---|---|---|
Aminoglycosides | Streptomycin | C21H39N7O12 | Inhibits protein synthesis | Ototoxicity, nephrotoxicity | GO/P(NIPAm-MPTC-GMA) | 38 |
Kanamycin | C18H36N4O11 | GO/Pt–Cu alloy/aptamer | 39 | |||
Tetracyclines | Tetracycline | C22H24N2O8 | Inhibits protein synthesis | Endocrine disruption of aquatic species, | rGO/Fe3O4/aptamer | 40 |
Oxytetracycline | C22H24N2O9 | GO/Au NPs/aptamer | 41 | |||
Sulfonamides | Sulfamethoxazole | C12H14N4O4S | Inhibits folic acid synthesis | Diarrhea, vomiting | Graphene/ZnO | 42 |
Sulfadiazine | C11H11N3O2S | GO/COF | 43 | |||
Macrolides | Azithromycin | C38H72N2O12 | Inhibits protein synthesis | Decreased shelf-life | GO/CNTs | 44 |
Erythromycin | C37H67NO13 | Graphene/Au–Pt NPs | 45 | |||
Quinolones | Ciprofloxacin | C17H18FN3O3 | Inhibits DNA replication | Severe hepatic toxicity | rGO/Au NPs/aptamer | 46 |
Ofloxacin | C18H20FN3O4 | GO/ionic liquid | 47 | |||
Beta-lactams | Amoxicillin | C16H19N3O5S | Inhibits cell wall synthesis | Rashes, fever | GO/CdTe/Au NPs | 48 |
Penicillin | C16H18N2O4S | GO/Fe3O4/CNTs/aptamer | 49 | |||
Amphenicols | Chloramphenicol | C11H12Cl2N2O5 | Inhibits protein synthesis | Inhibition of bone marrow | rGO/Co3O4 | 50 |
Furans | Nitrofurantoin | C8H6N4O5 | Inhibits protein synthesis | Ecological risks, human health damage | rGO/GdFeO3 | 51 |
Furazolidone | C8H7N3O5 | GO/GeW | 52 |
Fig. 1 Development timeline of graphene-based electrochemical sensors for antibiotics. (a) Aspects such as antibiotics, modifiers and LOD performance, and all corresponding data originate from ref. 17, 27, 46 and 53–73. (b) Evolution history of equipment platforms.44,62,69,74–78 |
Aptamers or immunosensors based on graphene-modified electrodes exhibit superior detection performance towards antibiotics than other types of modifiers with a lower LOD and deserve to be further developed. The equipment platforms constructed in electrochemical sensors is essential to real applications successfully and its evolution history is depicted in Fig. 1b. Graphene-based sensors in laboratory require bulky equipment and multiple processing involved in the conventional assays, while in the contrary, potable sensors being capable of on-site monitoring indicates miniaturized device, simplified processing and wireless transmission technology, and prevails in the future.
Fig. 2a depicts the relationships between three aspects as A, B and C. Suitable modifiers designed for electrode fabrication exhibit superior detection performance, and their upper limit was determined by real-sample analysis in this work. Conversely, the interpretation of sensing mechanism provides theoretical prediction of suitable modifiers for specific antibiotics. The sensing mechanism for antibiotics includes two parts: electrochemical reaction and interfacial reaction at the atomic level, as presented in Fig. 2b. Fig. 2c illustrates the miniaturization of sensors containing an integrated platform and screen-printed electrodes for on-site monitoring in real applications. The three aspects as A, B and C will be comprehensively discussed in the further sections.
Fig. 2 Schematic of key review elements involved in graphene-based electrochemical sensors for antibiotics. (a) Relationship between the above-mentioned three aspects. (b) Electrochemical sensing mechanism in theory. Reproduced with permission.79 Copyright from 2022, American Chemical Society. (c) Miniaturized sensing devices for on-site monitoring in real-time applications. |
Fig. 3 Working mechanism of electrochemical sensors for antibiotics: (a) electrochemical operation modes. (b) and (c) Scheme of direct reaction model and indirect reaction model. (d) Scheme of electrochemical aptasensor for chloramphenicol (CAP) detection. Reproduced with permission.69 Copyright from 2021, Elsevier. |
The electrical signals produced from different electrochemical detection models show a large difference due to the different construction designs of the sensing system.51,69,83 Through the review of the above-mentioned electrochemical sensors for antibiotics, two types of electrochemical detection models can be summarized as a direct reaction model and an indirect reaction model, as shown in Fig. 3b and c. Notably, the choice of electrochemical sensors depends on the properties of the target substance, leading to varying applicability of the two principles. For antibiotics containing structural components that readily participate in redox reactions, such as furazolidone and sulfamethoxazole, the direct model is appropriate for the detection. Conversely, for target antibiotics that exhibit difficulty in catalytic redox reactions, such as streptomycin and tetracyclines, the indirect model is the alternative option.
The electrical signals produced in the direct model are based on the direct catalytic reaction of antibiotics and enhanced with the increased concentrations of antibiotics. Through the optimization of parameters such as scan rate, pH, and amounts of modified materials, the electrochemical performance of the electrode modified for antibiotic detection is adjusted to the maximum. Then, the electrochemical redox reaction for antibiotic detection can be concluded as, for example, the irreversible reduction of a nitro group (R-NO2) of nitrofurantoin (NFT) into a phenylhydroxylamine group (R-NHOH) corresponding to an equal number of electrons (4e−) and protons (4H+) transferred.51 The current responses triggered by electrochemically active GdFeO3/rGO nanocomposites were enhanced with the increasing concentrations of NFT. However, the direct catalytic reduction of antibiotics consumes more energy and causes increased difficulty in the fabrication of hybrid materials with higher sensitivity.
To further improve the detection performance of antibiotics, indirect reaction models are developed based on the redox reactions of a pair of Fe[(CN)6]3−/4− and the special design of aptasensors. Ascribed to the easier catalytic reactions of a pair of Fe[(CN)6]3−/4− and the specific binding of aptamers and antibiotics, the electrical signals toward antibiotic detection from the indirect reaction models trigger the higher, and obtain better sensitivity than direct reaction models. Chloramphenicol is taken as a typical antibiotic for illustrating the indirect reaction model, as shown in Fig. 3d, wherein polyethyleneimine-functionalized reduced graphene oxide and gold nanocubes (PEI-rGO/AuNCs) served as the modified electrode to immobilize the aptamers.69 The large size and negative charge of single-stranded DNA-binding protein hinder the redox reactions of Fe[(CN)6]3−/4− and reduce the electrical signal drastically. When chloramphenicol is present in the electrolyte, the aptamer–chloramphenicol complex is formed, which inhibits the combination of single-stranded DNA-binding protein with the aptamer, resulting in a stronger response. Herein, the modification of PEI onto the rGO surface can be achieved via hydrogen bonding and electrostatic attraction, effectively anchoring it to the electrode.
In particular, due to the advantages of large specific surface area, strong interfacial charge transfer capability, good stability, and ease of functionalization, graphene is an ideal material for electrode modification, which triggers intense electrochemical responses in electrochemical sensors.
fk+ = qk(N + 1) − qk(N) | (1) |
fk− = qk(N) − qk(N − 1) | (2) |
The next is to focus on interfacial interactions between the electrode surface and adsorbed antibiotic molecules. The molecular electrostatic potential (MEP) map depicts the electrostatic potential distribution of positive and negative regions in the electrode-antibiotic system, identifying the sites for nucleophilic and electrophilic attacks.92 Qian et al. used the MEP map of isoniazid and different oxygen contents of rGO to reveal interfacial interaction between graphene and target antibiotics.79 The primary distribution of electron density on the oxygen groups of rGO created repulsive forces that prevented isoniazid from approaching the rGO sheet, resulting in worse detection performance. Notably, MEP offers an intuitive way to estimate the best sites of the system for interfacial interactions. Density of states (DOS) provides details about the chemical bonding between graphene-based nanocomposites and antibiotics, exploring the active sites and interfacial interactions from the perspective of orbital interaction.93 A quantitative tool to evaluate the stability of antibiotics adsorbed on various graphene sensing substrates was performed by adsorption energy (Ead) calculation, where the negative value indicates a spontaneous reaction involved in the adsorption process.79 In particular, the more negative value of Ead presents a stronger bond between the antibiotic and the substrate.94 The Ead value can be calculated using eqn (3), where E(sub/atb), E(sub) and E(atb) depict the total energies of the antibiotic-adsorbed graphene material, sensor substrate and antibiotic, respectively. Ren et al. studied the influence of N-doped graphene on the adsorption of phenol ions using Ead calculations.95 It revealed that N doping can alter the electronic structure of graphene and create new active sites, thereby enhancing the adsorption of target molecules.
To further investigate the type of interfacial interactions between graphene-based nanocomposites and antibiotics, the reduced density gradient (RDG) is applied using eqn (4), where ρ(r) denotes the electron density.92 The various types of interactions are discriminated by the function Signλ2(r)·ρ(r), where Signλ2(r) indicates the sign of the electron density Hessian matrix's second eigenvalue. Herein, Signλ2(r)·ρ(r) < 0 implies the strong attractive interactions such as hydrogen bond or halogen bond, while Signλ2(r)·ρ(r) ≈ 0 depicts the lesser attractive interactions as van der Waals (vdW) attraction and Signλ2(r)·ρ(r) > 0 presents the strong repulsive interactions as the steric effect. Using eqn (3) and (4), we can interpret the interfacial interactions between graphene-based nanocomposites and antibiotics quantitatively, and predict the suitable electrode modifiers to the specific antibiotic theoretically. Adekoya et al. adopted the RDG map to determine the specific type of interfacial interactions between graphene and antibiotic cephalexin (CEX).92 Hydrogen bonding facilitates the successful binding of CEX on GO/PEG in the [Signλ2(r)·ρ(r) < 0] region, confirming a strong interaction between CEX and GO/PEG.
Ead = E(sub/atb) − E(sub) − E(atb) | (3) |
(4) |
Fig. 4 Identification of active sites and chemical reactivity of antibiotic structures. (a) Illustration of the 3D structure, the lowest energy of conformation, f− indices, (b) molecular structure, (c) HOMO distribution of levofloxacin and (d) protonation probability for C9 and C13 atoms functioned with buffer pH. (a)–(d) Reproduced with permission.90 Copyright from 2018, Wiley-VCH. (e) HOMO, LOMO and the energy gap of sulfamethazine, sulfathiazole and sulfamethoxazole. Reproduced with permission.91 Copyright from 2015, Walter de Gruyter. |
In conclusion, to assess the chemical reactivity of the target antibiotic, the HOMO–LUMO energy gap is the suitable evaluation index. A lower value indicates that the target antibiotic is highly polarizable, often exhibiting strong chemical reactivity but low stability. Fukui indexes can be further applied to identify the vulnerable sites of specific antibiotics to redox reactions.
Fig. 5 Interfacial interactions of antibiotics adsorbed on graphene-based nanocomposites. (a) Optimized structure of GO/PEG–CEX. (b) MEP distribution of CEX, GO/PEG and GO/PEG–CEX. The blue regions indicate low electron density, while the red domains depict electron enrichment. RDG isosurface map for determining the type of interfacial interactions in (c) GO, (d) GO/PEG and (e) GO/PEG–CEX. (a)–(e) Reproduced with permission.92 Copyright from 2022, American Chemical Society. |
The influences of the electronic structure of graphene regulated by doping with elements such as nitrogen (N), varying oxygen content or hybridization with other materials are vital to elaborate the interfacial interactions of antibiotics adsorbed on graphene-based nanocomposites. Ren et al. investigated the influence of N-doped graphene on the adsorption of phenol ions by DFT modelling.95 As depicted in Fig. 6a–d, the distances between the pristine rGO plane and its next to graphitic N (γ1 and γ2), pyridinic N and O atom of phenol were 2.834, 1.597, 1.508 and 2.426 Å, respectively. Compared to pristine rGO, the shorter C–O distance of N-doped rGO demonstrated that N doping can alter the electronic structure of graphene and create new active sites for enhancing the adsorption of phenol ions. The rGO plane adjacent to graphitic N presents a higher affinity for adsorbing phenol ions than pyridinic N, while pyrrolic N indicates ineffective adsorption capacity. Therefore, N doping enhances the metal-like property of graphene due to the induction of a high positive charge density on the ortho-carbon atoms, promoting its adsorption capacity.
Fig. 6 Interfacial interactions of antibiotics adsorbed on graphene-based nanocomposites. Optimal models for the adsorption of a phenol ion on (a) pristine rGO and (b)–(d) various sites of N-doped rGO including three types of N as pyridinic (α), pyrrolic (β) and graphitic (γ). (a)–(d) Reproduced with permission.95 Copyright from 2018, Elsevier. (e) Optimized structures of isoniazid adsorption on rGO1 and rGO2. (f) Molecular electrostatic map of isoniazid, rGO1, and rGO2. (e) and (f) Reproduced with permission.79 Copyright from 2021, American Chemical Society. (g) Illustration of low-energy adsorption of amoxicillin onto the Au/graphene surface. (h) Total density of states (DOS) and partial density of states (PDOS) contributed by the Au/graphene electrode and amoxicillin. (g) and (h) Reproduced with permission.93 Copyright from 2021, Elsevier. |
Qian et al. conducted a DFT calculation to reveal the roles of rGO oxygen content in developing graphene-based electrochemical detection of isoniazid with high performance.79 As shown in Fig. 6e, the positive Ead of the rGO1 sensor with a higher oxygen content and isoniazid exhibited repulsive forces that prevented isoniazid from approaching the rGO sheet. Fig. 6f shows the molecular electrostatic potential of isoniazid, rGO1 and rGO2, and indicates the primary distribution of electron density on the oxygen groups to form repulsive forces. The computational results confirmed the experimental data that rGO1 behaved with weaker voltammetric responses. Osikoya et al. investigated the adsorption behaviour of amoxicillin on the Au/graphene electrode by DFT calculation.93 As presented in Fig. 6g, amoxicillin presenting a vertical configuration with a –NH2 group preferred to be adsorbed stably onto the Au/graphene electrode surface, with an Ead value of −2.03 eV. The adsorption behaviour can be attributed to the formation of a strong X–H⋯Au (X = C or N) interaction between the Au atom and the –CH3 or –NH2 group of amoxicillin. Partial density of states results presented in Fig. 6h suggest that amoxicillin has a high affinity towards the Au/graphene nanointerface.
In summary, the adsorption of the target antibiotic preferentially occurred on graphene-based nanocomposites via hydrogen bonding, electrostatic interaction or covalent bonding. Carboxylic, epoxy or hydroxyl functional groups are the common adsorption sites on graphene-based electrodes for targeting antibiotics that contain a –CH3 or –NH2 group. The abundant oxygen functional groups on graphene preferentially facilitate the adsorption of target antibiotic stably through various interfacial interactions.
Fig. 7 Surface functionalization/decoration of graphene: (a) graphene and its derivatives. (b) Schematic of various modifiers functionalized on graphene. Reproduced with permission.100 Copyright from 2019, Multidisciplinary Digital Publishing Institute. |
As graphene's electrocatalytic activity is restricted, the functionalization of graphene is considered the effective way to facilitate its applications further and is classified as covalent and non-covalent modification, as illustrated in Fig. 7b. Covalent modification can be achieved by forming amide and carbamate ester bonds with the carboxyl and hydroxyl groups, and non-covalent modification includes weak van der Waals forces, π–π stacking, and electrostatic interaction.25,101,102 A series of materials such as metal NPs, metal oxides, enzymes, aptamers and others are applied as modifiers to combine with graphene and fabricate electrodes to form electrochemical sensors for antibiotics.64,68,93,103,104 Our reviewing in this part focuses on the prevalent methods to fabricate various graphene-based hybrid structures and the relationship between material structure and sensing performance, providing valuable references for the synthesis of graphene-based modifiers for electrode fabrication.
Fig. 8 Solution blending method for preparing graphene-based hybrids as (a) Ti3C2Tx/TiO2 NPs/rGO. Reproduced with permission.105 Copyright from 2024, Wiley-VCH. (b) rGO/PEI/TiO2. Reproduced with permission.106 Copyright from 2023, Elsevier. |
Fig. 9 Interfacial electrostatic self-assembly method for preparing films with various graphene-based hybrid structures. (a) Illustrative preparation of rGO/SiO2 nanospheres and (b) their TEM image. (a) and (b) Reproduced with permission.108 Copyright from 2021, American Chemical Society. (c) Controllable formation of electroactive nanostructures by the LBL assembly method. Reproduced with permission.110 Copyright from 2011, American Chemical Society. (d) Multilayered structure of rGO/CNTs by the LBL assembly method. Reproduced with permission.109 Copyright from 2010, American Chemical Society. (e) pH regulation on the structure of LBL-assembled multilayered films. Reproduced with permission.111 Copyright from 2016, Springer Nature. |
Furthermore, to precisely regulate the film thickness formed on substrates, the layer-by-layer (LBL) assembly method is commonly used to obtain optimal coating films.112 Multilayered nanostructures composed of graphene and CNTs can be controllably formed on electrodes via the LBL method, using electrostatic and π–π interactions, as illustrated in Fig. 9c and d.110 Typically through the π-stacking interactions, CNTs adhere to GO flakes well and disperse in water stably by the high solubility of GO as a surfactant, facilitating the formation of the films uniformly. The controlled layering process yields optimal performance for sensing applications. The thickness and structure of the LBL-assembled films could be controlled and altered by pH with electrostatic and hydrogen bonding, providing an innovative perspective to fabricate the hybrid film by the LBL method, as shown in Fig. 9e.111
In conclusion, an interfacial electrostatic self-assembly offers an optimal method for fabricating a uniform hybrid structure with an enlarged specific surface area and improved target adsorption, thereby boosting its sensing performance.
Fig. 10 In situ growth method for obtaining strong interfacial contact between graphene and modifiers. (a) Illustration of graphene deposition on the substrate at high temperatures. (b) In situ growth of graphene on HPHT diamond by thermal treatment. SEM images of HPHT diamond (c) before and (d) after treatment. (e) Raman spectrum and (f) its mapping of graphene-diamond. (g) Formation mechanism of graphene-diamond. (b)–(g) Reproduced with permission.113 Copyright from 2018, Elsevier. (h) Solvo-hydrothermal method for the in situ growth of graphene-based modifiers. (i) Electrode modified with zeolitic imidazole framework-derived Co and nitrogen co-doped carbon polyhedrons (ZIF-67C) and GO hybrids. Reproduced with permission.114 Copyright from 2019, Springer. (j) Formation of Co2+ doping and synthesis procedure of Co–Fe3O4 nanospheres/GO. Reproduced with permission. Copyright from 2021, American Chemical Society. |
Fig. 10h shows the solvo-hydrothermal method for the in situ growth of graphene-based modifiers. Chen et al. synthesized a homogeneous nanocomposite of zeolitic imidazole framework-derived Co and nitrogen co-doped carbon polyhedrons and rGO by an in situ growth method,114 as depicted in Fig. 10i. The composite material derived from MOF and rGO demonstrates excellent stability, selectivity and sensitivity towards Metronidazole detection, owing to its enhanced mass transfer, abundant electroactive sites, and superior conductivity. Nehru et al. fabricated cobalt-doped Fe3O4 nanospheres deposited on GO via a facile hydrothermal technique,66 as shown in Fig. 10j. The schematic Fe3O4 crystal structure (site A and B) represents a cubic crystal feature. Here, site A has been occupied by Fe3+ and site B has been occupied by an equal number of Fe2+ and Fe3+ ions. During the introduction of Co2+ into the Fe3O4 matrix, Fe3+ was replaced by Co2+ in the site A, while Fe2+ was transformed to Fe3+ in site-B to sustain electric neutrality. The in situ growth of Co-doped Fe3O4 on GO improved the performance shortages of the Fe3O4 system due to agglomeration and less conductivity, and enhanced its operation stability due to the strong intermolecular interactions. Thus, the hybrid boosts the electrochemical kinetics and exhibits a low detection limit of 1.04 nM towards chloramphenicol. In summary, the in situ growth of graphene-based modifiers renders superior mechanical stability, electrical conductivity, abundant electro-active sites, and electrocatalytic activity and boosts the electron-transfer kinetics for antibiotic detection.40,60,115
Fig. 11 Glove-based wearable sensors for antibiotic detection. (a) Schematic. (b) Configuration of the glove-based stretchable device, with (c) scan finger (left) containing biosensing electrodes and collection finger (right). (d) Photograph of the glove-based sensor, containing a ring bandage connecting the electrodes with a portable potentiostat. (e) and (f) Swipe sampling protocol on the glove to complete the sensing function. (b)–(f) Reproduced with permission.123 Copyright from 2017, American Chemical Society. (g) A laser-induced flexible electrochemical sensing system integrated on the finger. (h) DPV curves of antibiotic detection on flexible graphene electrodes. (g) and (h) Reproduced with permission.77 Copyright from 2022, Elsevier. (i) and (j) Photograph of the glove-based sensor for on-site monitoring in food samples. Reproduced with permission.124 Copyright from 2021, Elsevier. |
Li et al. constructed a laser-induced flexible electrochemical sensing system on fingers for rapid real-time on-site identification of chloramphenicol, clenbuterol, and ractopamine in meat.77 As depicted in Fig. 11g, flexible graphene electrodes were facilely patterned and prepared by CO2 laser and integrated with disposable blue nitrile gloves for constructing a finger-based sensing system. Through the connection with a portable electrochemical analyser, electrochemical signals could be received by the directly touching object under DPV tests with fingertips and displayed on mobile phone. The LOD of chloramphenicol and the other two feed additives, namely, clenbuterol and ractopamine was 2.70, 1.29 and 7.81 μM, respectively, enabling the successful application of the sensing platform on the finger for food security, as shown in Fig. 11h. Raymundo-Pereira et al. developed a non-enzymatic sensor system printed on three fingers of a rubber glove for the detection of carbendazim in food samples, as depicted in Fig. 11i.124 In Fig. 11j, the sensor was capable of monitoring carbendazim in a rapid and cost-effective manner by directly touching the samples with the glove, exhibiting 47 nM LOD for carbendazim by the DPV method, and was successfully applied for cabbage and juice samples.
Fig. 12 Epidermal and microneedle-based wearable sensors through sweat or interstitial fluid sampling for antibiotic detection in clinical treatment. (a) Scheme of an epidermal sensor involving the integration of skin multilayer models. Reproduced with permission.116 Copyright from 2020, American Chemical Society. (b) A wearable platform integrated into a wristband for noninvasive drug monitoring and the cross-section view of the flexible electrodes. (c) Sweat extraction through iontophoresis method and (d) DPV signals from caffeine detection. (b)–(d) Reproduced with permission.126 Copyright from 2018, Wiley-VCH. (e) Scheme of microneedle transdermal sensing via interstitial fluid access. (f) and (g) Microneedle array biosensor to the forearm with 60 s firm pressure. (h) Diagram of the penetration of microneedle array into the dermal-interstitial space. (f)–(h) Reproduced with permission.75 Copyright from 2019, Elsevier. |
To overcome the difficulty of accessing human interstitial fluids, the microneedle array has been developed to substitute hypodermic needles for transdermal drug delivery and sampling human interstitial fluids. Microneedle-based wearable sensors are constructed by the integration of multiple microneedle electrodes on a wearable patch device for real-time diagnostic evaluation of the antibiotic level.75,127Fig. 12e illustrates the microneedle transdermal sensing via interstitial fluid access. Rawson et al. developed a microneedle β-lactamase biosensor for real-time and minimally invasive monitoring of penicillin V in vivo, as shown in Fig. 12f and g.75 The open circuit potential of the working electrode versus the reference electrode was recorded as detection signals, as depicted in Fig. 12h. Penicillin V, as a common antibiotic to cure bacterial infections, diffuses from the extracellular fluid to the hydrogel layer, and hydrolysed to penicilloate and a proton by β-lactamase. Therefore, the increasing concentration of penicillin V in the tissue promoted the protons generated at the sensor surface, inducing the elevations of open circuit potential. The LOD of penicillin V by the microneedle method was estimated as 0.17 mg L−1, prospecting potential antibiotic monitoring for individualized dose optimization.
Kling et al. constructed a microfluidic platform enabling the electrochemical readout of up to eight enzyme-linked assays for the simultaneous detection of two antibiotics, namely, tetracycline and streptogramin in spiked human plasma, as presented in Fig. 13a.74 The microchannel network contained a single electrochemical cell with a three-electrode setup for the amperometric signal detection, and was connected to the substrate solution reservoir via the common to the outlet. The signal amplification of antibiotic detection was obtained using a defined stop-flow measurement technique, and the corresponding function principle is illustrated in Fig. 13b. The bound enzyme glucose oxidase generated limited H2O2 when the glucose substrate was constantly supplied in the microchannel and catalyzed the reaction to more H2O2 when the flow was stopped. When restarting the flow, the accumulated clouds were flushed over the working electrode, and the generated peak signal from H2O2 oxidation was measured amperometrically, as depicted in Fig. 13c and d. The formula of antibiotic detection in the human plasma involved a DNA–protein interaction sensing mechanism.74 The repressor protein showed a conformational change in the antibiotic presence, indicating no binding capability to their designated operator DNA. The higher concentration of antibiotics resulted in lesser proteins bound to operator DNA, demonstrating the decreased electrochemical signal. The platform concluded the LOD of 6.33 and 9.22 ng mL−1 for tetracycline and pristinamycin respectively, as shown in Fig. 13e.
Fig. 13 Microfluidic electrochemical chip proposed for the simultaneous detection of multi-antibiotics. Microfluidic biosensors comprising multiple immobilization sections: (a) system configuration; (b) measurement principle; (c) sensing device; (d) reaction mechanism of H2O2 oxidation at the working electrode; (e) amperometric responses and the resulting on-chip calibration curve. (a)–(e) Reproduced with permission.74 Copyright from 2016, American Chemical Society. (f) A microfluidic electrochemical biosensor constructed for multianalyte monitoring based on enzymatic catalysis. Reproduced with permission.128 Copyright from 2022, Wiley-VCH. (g) Paper-based microfluidic device for multianalyte monitoring based on the respiratory inhibition of E. coli. Reproduced with permission.129 Copyright from 2019, Elsevier. |
Dincer et al. constructed a microfluidic electrochemical biosensor consisting of an electrochemical cell, immobilized region, and a hydrophobic barrier in prevention of electrode fouling during detection, as depicted in Fig. 13f.128 The measurement signal could be obtained from the catalysis reaction of H2O2 by glucose oxidase, and the platform proved the feasibility for the on-site monitoring of multi-antibiotics such as piperacillin, tazobactam and meropenem in human biofluids. Zhang et al. developed a paper-based microfluidic device for multianalyte monitoring, as displayed in Fig. 13g,129 which was operated based on the respiratory inhibition of E. coli due to the interference of antibiotics in the environment, resulting in the amplification of electrochemical signal. The paper-based microfluidic platform offers a new perspective for the on-site monitoring of multi-antibiotics.
In summary, a portable electrochemical platform for the on-site monitoring of antibiotics depends on specific application requirements, leading to varying applicability of the three wearable scenarios. For the convenient monitoring of antibiotics in food samples, a wearable “lab-on-a-glove” platform shows great potential to achieve the goal. Yet, this platform still faces a bottleneck in the limit of detection of antibiotics.
For the non-invasive monitoring of antibiotics in human interstitial fluids such as sweat and blood, an epidermal and microneedle-based platform is the optimum choice for clinical monitoring. It can achieve trace-level or ultra trace-level detection limits of target antibiotics.
Moreover, the microfluidic electrochemical chip provides an opportunity for the simultaneous detection of multi-antibiotics in human biofluids with accuracy, but it suffers from the limitation of complicated device construction. The selection of graphene-based electrodes for diverse wearable platforms is contingent upon specific application scenarios, whereas graphene facilitates signal amplification during detection periods.
Graphene-based modifying electrode | Analyte | Technique | Detection range | LOD | Real sample | Ref. | |
---|---|---|---|---|---|---|---|
Noble metal NPs | rGO/Ag NPs/NiF | Amikacin | DPV | 0.05–15.0 μM | 38.0 nM | Urine | 131 |
rGO/Au NPs/Pd NPs | Lomefloxacin | SWV | 4.0–500.0 μM, 30.0–350.0 μM | 38.0 nM | Urine | 56 | |
Amoxicillin | 9.0 μM | ||||||
rGO/Au NPs | Chloramphenicol | LSV | 2.0–80.0 μM | 0.59 μM | Eye drops | 53 | |
rGO/Pd NPs | Chloramphenicol | DPV | 0.05–1.0 μM | 0.05 μM | Honey, tap water | 132 | |
GO/Au NPs | Clindamycin | SWV | 0.95–140.0 μM | 0.29 μM | Urine, river water | 73 | |
rGO/Au NPs/polypyrrole | Doxorubicin | CV | 0.02 μM–25.0 mM | 0.02 μM | Drugs | 133 | |
rGO/Cu NPs | Metronidazole | i–t | 0.002–210.0 μM | 0.6 nM | Drugs | 60 | |
GO/Ag NPs | Metronidazole | i–t | 0.09–4594.0 μM | 0.07 μM | Drugs | 134 | |
CNTs | GO/CNTs | Azithromycin | LSV | 0.1–10 μM | 0.07 μM | Urine, capsules | 44 |
rGO/VS2/CNTs | Azithromycin | DPV | 2.8–300 nM | 0.9 nM | Human serum, urine | 67 | |
rGO/CNTs | Natamycin | ASV | 0.05–2.5 μM | 0.01 μM | Red wine, beverage | 58 | |
GO-CNTs | Tetracycline | DPV | 20.0–310.0 μM | 0.36 μM | River water | 54 | |
MOF | rGO/Z-800 | Chloramphenicol | DPV | 1.0–180.0 μM | 0.25 μM | Milk, honey | 135 |
rGO/NH2-UiO-66 | Ciprofloxacin | ASV | 0.02–1.0 μM | 6.67 nM | Tap water, lake water | 62 | |
Metal oxides | rGO/Eu2O3 | Chloramphenicol | i–t | 0.02–800.3 μM | 1.32 nM | Milk, honey | 63 |
rGO/TiO2 | Furazolidone | DPV | 1.0–150.0 pM | 0.43 pM | Blood serum | 136 | |
Graphene/MnMoO4 | Ornidazole | i–t | 0.01–0.77 μM | 0.85 nM | River water | 68 | |
Graphene/ZnO | Sulfamethoxazole | DPV | 1–170 μM | 0.4 μM | Urine, serum, lake water | 42 | |
Trimethoprim | 1–170 μM | 0.3 μM | |||||
GO/Fe3O4/Co | Chloramphenicol | DPV | 0.005–152.2 μM | 1.04 nM | Food | 66 | |
Immunosensors | Graphene/prussian/chitosan | Kanamycin | DPV | 0.04–28.8 nM | 13.0 pM | Food | 137 |
GO/P(NIPAm-MPTC-GMA) | Streptomycin | DPV | 0.09–170.0 nM | 2.89 pM | Milk | 38 | |
Graphene/Zn/Ni-ZIF/Au NPs | Monensin | DPV | 0.38–150.0 nM | 0.17 nM | Milk | 138 | |
rGO/Au NPs/peroxidase | Oxytetracycline | DPV | 1.0 pM–4.0 μM | 1.0 pM | Food | 41 | |
Aptamers | GO/Ag NPs/aptamer | Chloramphenicol | DPV | 10.0 pM–0.2 μM | 3.3 pM | Milk, honey | 65 |
rGO/Au NPs/aptamer | Ciprofloxacin | SWV | 0.001–1.0 μM | 1.0 nM | Milk | 46 | |
Graphene/Pt–Cu alloy/aptamer | Kanamycin | DPV | 1.0 pM–10.3 nM | 0.87 pM | Food | 39 | |
GO/aptamer | Tetracycline | EIS | 0.1 pM–10.0 μM | 0.03 pM | Blood serum | 139 | |
GR/Fe3O4/Au NPs/aptamer | Streptomycin | DPV | 0.09–340.0 nM | 0.05 nM | Milk | 57 |
Although electrochemical sensing methodology has been successfully applied for antibiotic detection, there are still many scientific and technical challenges that need to be overcome in the future. As illustrated in Fig. 14, the current challenges can be listed as follows:
(1) The category distribution of electrochemical sensors for antibiotics is inhomogeneous. Due to the presence of multiple antibiotics in the environment, the current sensors primarily designed for common antibiotics, typically amphenicols and aminoglycosides, are not sufficient and beneficial for the trace determination of antibiotic residues in the environment. Constructing electrochemical sensors for antibiotics covering as many species as possible is necessary.
(2) The electrodes designed on graphene-based modifiers without the selective capability suffer technical difficulty in the selectivity of target antibiotics, especially in the presence of multiple antibiotic species with similar structures in real samples. While for the current immunosensors or aptasensors aimed to the selectivity issue, the enduring preservation and maintenance of bioactive molecules as immune components and aptamers pose a challenge to ensure their consistence performance over time. Thus, constructing electrochemical sensors for antibiotics that possess both high sensitivity and excellent selectivity towards various types of antibiotics is still urgent and meaningful. Particularly, ensuring the long-term stability of graphene-based sensors is essential for the sustained monitoring and practical applications.
(3) The theoretical interpretation of the antibiotic sensing mechanism is vitally important for pursuing excellent detection performance with various electrochemical techniques. Yet, the sensing mechanism of antibiotics reported in current research papers mainly refers to the analysis of electrical responses. The redox reaction process involving electron transfer and interfacial interactions between the electrode surface and the adsorbed molecules at the atomic level should be carried out for in-depth exploration, using theoretical calculation software of the quantum-mechanical method. It not only aggravates the interpretation of the antibiotic sensing mechanism theoretically, but also provides key information about the specific nanomaterials with sensitivity to every antibiotic.
(4) Various forms of antibiotics exist in environment such as in river water, food, human serum, urine, and honey, increasing the difficulty for the development of portable electrochemical platforms for the on-site monitoring of antibiotics. Thus, the platform should be automated, conforming to the scenario requirements of antibiotic detection in various real samples. The optimization of the microneedle, epidermal or microfluidic platform is hopeful for on-site therapeutic antibiotic monitoring in human biofluids towards clinical treatment. Glove-based wearable sensors show potential for the future on-site monitoring of antibiotics in the environment. Moreover, the cost of constructing graphene-based electrochemical sensors to achieve mass production should be reduced to an accepted extent to realize the commercial application.
Overall, the future development of graphene-based electrochemical sensors for antibiotics requires addressing the challenges at the technological, engineering, and application levels. The deep exploration within the scope enables the further utilization of the sensors in the fields of healthcare monitoring, environmental protection, and food safety.
This journal is © The Royal Society of Chemistry 2025 |