Sharmila
Prashanth
a,
Raifa Abdul
Aziz
b,
Shamprasad Varija
Raghu
bc,
Yoon-Bo
Shim
d,
K.
Sudhakara Prasad
*a and
Airody Vasudeva
Adhikari
*ae
aNanomaterial Research Laboratory (NMRL), Smart Materials And Devices, Yenepoya Research Centre and Centre for Nutrition Studies, Yenepoya (Deemed to be University), Deralakatte, Mangalore 575018, India. E-mail: avachem@gmail.com; ksprasadnair@yenepoya.edu.in
bNeurogenetics Lab, Department of Applied Zoology, Mangalore University, Mangalagangothri, 574199, Karnataka, India
cDivision of Neuroscience, Yenepoya Research Centre, Yenepoya (Deemed to be University), Deralakatte, Mangalore 575018, India
dDepartment of Chemistry and Institute of Biophysio Sensor Technology, Pusan National University, Busan 46241, Republic of Korea
eDepartment of Chemistry, National Institute of Technology Karnataka, Surathkal, Mangalore 575025, India
First published on 11th December 2023
Serotonin, also known as 5-hydroxytryptamine (5-HT), is an important neurotransmitter that regulates many physiological processes. Both low and high concentrations of 5-HT in the body are associated with several neurological disorders. Hence, there is an urgent need to develop fast, accurate, reliable, and cost-effective disposable sensors for 5-HT detection. Herein, we report the sensing of 5-HT using a disposable paper-based electrode (PPE) modified with a ternary nanocomposite comprising poly(pyrrole) (P(py)), reduced graphene oxide (rGO), and iron oxide (Fe2O3). The sensor material was well characterized in terms of its structural, morphological, and chemical attributes using electron microscopy, spectral techniques, and electrochemical studies to prove the robust formation of the electroactive ternary nanocomposite and its suitability for 5-HT detection. The developed sensor exhibited an impressive limit of detection (LOD) of 22 nM with a wide linear range of 0.01 to 500 μM, which falls in the recommended clinically relevant range. The analytical recovery, spike sample analysis, and interference studies with ascorbic acid (AA), uric acid (UA), and epinephrine (E) showed satisfactory results, wherein the sensor could detect simultaneously both 5-HT and dopamine (DA). The potential practical utility of the developed sensor was further assessed by quantifying the concentration of 5-HT in the brain samples of Drosophila melanogaster, a versatile genetic model organism employed for modeling different neural disorders in humans, and validated by gold-standard HPLC-UV experiments. The as-fabricated single-run disposable sensor with a ternary nanocomposite exhibits excellent stability with good reproducibility and is a promising platform for identifying clinically relevant concentrations of 5-HT.
Electrochemical methods comprise a collection of extremely useful measurement tools, and the right choice of electrode materials plays a pivotal role, i.e., achieving electrodes with the best possible properties can lead to products with a wide range of potential applications.12,13 Paper-based electrodes (PPEs) have recently gained much attention as prospective alternatives to traditional three-electrode systems because of their low cost, easy disposability, portability, ion accessibility, flexibility, and facile manufacturing methods.11,14 Normally, the PPE design consists of hydrophobic and hydrophilic regions, along with 3-in-1 electrode patterns incorporating working, counter, and reference electrodes.15 To the best of our knowledge, there have been limited studies on disposable sensors for the detection of 5-HT, which can detect 5-HT in brain fluids and can be validated using gold standard HPLC-UV methods. In view of this, in the present study, we have chosen an inexpensive PPE for developing an electrochemically active ternary nanocomposite for detecting 5-HT levels.
Polymer-modified electrodes (PMEs) have gained significant attention for their ability to enhance sensitivity and selectivity. Conducting polymers are potential candidates for such electrodes, as they possess features responsible for their anomalous electronic properties like high electrical conductivity, low electronegativity, and high electron affinity due to their π-electron conjugated backbones, making them excellent materials for the immobilization of biomolecules.16,17 P(py), a biocompatible polymer, synthesized via electro-polymerization, has emerged as a promising electronic sensing material, garnering attention in the sensor field. Owing to its inherent electrical conductivity, favourable stability, redox characteristics, and extensive surface area, P(py) is a promising electronic sensing material, particularly for use in neutral pH environments.17–19 However, P(py) exhibits certain limitations, including poor mechanical properties, relatively poor thermal stability in air, non-biodegradability, and lack of mechanical stability, which have hindered a more comprehensive range of applications.20 Subsequently, previous research endeavours have proposed various strategies to enhance the properties of P(py). One such strategy involves incorporating appropriate materials like reduced graphene oxide (rGO) and metal nanoparticles to achieve the desired characteristics of the electrode.21
Electrochemical detection of 5-HT is essentially based on the oxidation on a suitable electrocatalyst surface. Recently, it has been demonstrated that P(py)–rGO, a binary composite electrode, performs better when compared to the P(py) electrode alone, exhibiting higher capacitance and lower charge transfer resistance with excellent cyclability and renewability.22 rGO is a highly versatile material, effective for various applications, including electrocatalysis and electrochemical sensing. Its outstanding properties, such as high conductivity, large surface area, and tunable electronic structure, make it an attractive material for these applications. However, the primary drawback associated with binary nanocomposites is their inadequate stability during catalysis. Therefore, developing a straightforward method for creating nanocomposites that demonstrate superior catalytic performance is crucial.23
Transition metal nanoparticles are widely recognized for their exceptional conductivity and catalytic properties, which increase the electron transfer between the redox center in the target analyte and the electrode surface.24 It is well-established that nanocomposites comprising conducting polymers and transition metal nanoparticles are advantageous for achieving enhanced sensitivity and stability. Also, the incorporated metal nanoparticles serve as redox mediators, facilitating efficient electron transfer. Simultaneously, the polymer component acts as an adsorbent for biomolecules, further improving the performance of the nanocomposite.23,24 Over the last decade, considerable attention has been given to transition metal oxide nanoparticles due to their notable advantages, which include their ability to be functionalized, exceptional thermal and chemical stability, tunable oxidation states, unique optical properties, and large surface area. These characteristics make them highly appealing for various scientific and technological applications.25,26 Several research groups have investigated the applications of ternary nanocomposites, such as Ag–P(py)–Cu2O,25 rGO–Ag2Se,27 SPCE–ZnONR–PMB(DES),28 poly(bromocresol l green)–Fe3O4,29 P(py)–Fe3O4,30 and FeC–AuNPs–MWCNTs31 for 5-HT sensing. Even though ternary nanocomposites have been explored for 5-HT sensing, most of the reported articles failed to validate the sensors using gold standard methods, and also the real sample analysis was conducted with spiked blood or urine samples in the absence of 5-HT. Hence, it is imperative to develop a sensor that has the potential to detect 5-HT directly in clinical samples or in brain fluids.
Drosophila melanogaster, commonly referred to as the fruit fly, is a significant and simplified model organism employed for modelling different neural disorders in humans.32 The utilization of Drosophila in neuroscience research is favoured due to its genetic tractability, complex behavioural patterns, well-documented and straightforward neuroanatomy, and possession of many orthologous genes to humans.33 So, in our experimental design, we used the brain samples of Drosophila melanogaster for the quantitative and sensitive detection of 5-HT employing the newly fabricated electrode.
In the present work, a simple and cost-effective method for determining 5-HT was proposed using a P(py)–rGO–Fe2O3 ternary nanocomposite decorated PPE. The fabrication was achieved using a step-by-step procedure, where layer-by-layer assembly of the ternary nanocomposite is established. At first by depositing P(py) as a base coat using amperometry, a conductive platform is established followed by the electro-reduction of drop-coated graphene oxide (GO), and finally, the electro-deposition of Fe2O3 was carried out through cyclic voltammetry. The fabricated PPEs were characterized using electron microscopy (FE-SEM) and spectral techniques (FT-IR and XPS) and subjected to electrochemical studies such as impedance spectroscopy (EIS) and cyclic voltammetry (CV) to understand the robust formation of the sensor. The sensor was further tested with different concentrations of 5-HT and investigated to satisfy the requirements of a real-world sensor for clinical applications. The newly developed sensor is further validated using the gold standard HPLC-UV method and explored to detect the limited quantity of 5-HT in the Drosophila melanogaster brain samples.
Electrochemical experiments, including electrochemical impedance spectroscopy (EIS), cyclic voltammetry (CV), and amperometry, were conducted using a CHI 708E electrochemical workstation from CH Instruments and a Zensor simulator potentiostat (Zensor®, Ecas 100 from Taiwan). Surface morphological analyses of the fabricated sensor were performed using Fourier transform infrared-attenuated total reflectance (ATR-FTIR) spectroscopy with a Shimadzu instrument from Japan and a field emission scanning electron microscope (FE-SEM) equipped with an Oxford energy-dispersive X-ray spectrometer (EDS). The FE-SEM utilized was a model EVO MA18 with a magnification range of 1× to 100000×. X-ray photoelectron spectroscopy (XPS) experiments were carried out using a K-alpha X-ray photoelectron spectrometer manufactured by Thermo Scientific in the United Kingdom. Curve fitting of XPS data was performed using the XPS PEAK 41 system software.
Scheme 1 Schematic illustration of the synthesis of the P(py)–rGO–Fe2O3 ternary nanocomposite on the surface of the PPE, and the inset shows the drop-coating method. |
Furthermore, to elucidate the alterations in surface functionality during the fabrication, we conducted a thorough analysis employing Fourier-Transform Infrared Spectroscopy (ATR-FTIR), and X-ray Photoelectron Spectroscopy (XPS). Fig. S2 (ESI†) presents the outcomes derived from the AT-FTIR spectrum of the modified electrodes. The PPE modified with P(py) exhibits additional peaks at 1577 cm−1 and 1438 cm−1, which indicate the characteristic vibrations associated with C–C and C–N in the P(py) ring.42 The peaks observed at 2967 cm−1 and 3347 cm−1 correspond to stretching vibrations of C–H and O–H bonds, respectively, suggesting the incorporation of P(py) on the electrode surface. PPE–P(py)–rGO displays a slight shift in the peak, indicating a change in the chemical environment. In summary, the FTIR analysis provided compelling evidence supporting the existence of nanocomponents in the modified paper electrodes. In addition, XPS analysis was carried out to understand each modified electrode's elemental electronic states and surface composition. The XPS spectra of the newly formed nanocomposites are depicted in Fig. 3. The investigation involved wide scan and deconvoluted XPS spectra of C 1s, O 1s, N 1s, and Fe 2p for PPE–P(py) (Fig. S1A, ESI†), PPE–P(py)–rGO (Fig. S1B, ESI†), and PPE–P(py)–rGO–Fe2O3 (Fig. 3). The atomic percentages of carbon (C), oxygen (O), nitrogen (N), and iron (Fe) were determined from the XPS spectra to gain insights into the compositional changes induced by the surface modifications.
The PPE exhibited a predominant composition of carbon (90.29%) and oxygen (9.71%), with no detectable nitrogen. This composition suggests that the PPE mainly consists of a carbonaceous material, likely having oxygen-containing functional groups. Upon the electrodeposition of P(py) onto the PPE, the atomic composition revealed a decrease in carbon content (87.7%), a slight increase in oxygen content (10.01%), and the appearance of nitrogen (2.29%). The presence of nitrogen indicates the successful incorporation of P(py) molecules on the PPE surface, forming nitrogen-containing functional groups. In the PPE–P(py)–rGO–Fe2O3 sample, the carbon content further decreased (81.29%) compared to the previous sample, while the oxygen content increased (16.7%). The nitrogen content also decreased to 0.81%. These observations suggest that the interaction between P(py) molecules and the PPE extends to include rGO in the system, resulting in surface chemistry modifications. Furthermore, the PPE–P(py)–rGO–Fe2O3 ternary nanocomposite exhibited a notable decrease in carbon content (75.31%), accompanied by a significant increase in oxygen content (21.38%). Minor amounts of iron (0.32%) and nitrogen (0.72%) were also detected. These findings strongly indicate the successful integration of iron oxide nanoparticles (Fe2O3) into the system, likely interacting with both rGO and P(py) on the PPE surface. The observed changes in carbon, oxygen, nitrogen, and iron contents suggest the formation of new chemical bonds and functional groups, resulting in altered surface properties. The XPS results thus provide valuable insights into the layer-by-layer modification of the PPE.
Notably, an increase in the O1/C1 atomic ratio was consistently observed throughout the layer modifications, signifying a higher proportion of oxygen atoms relative to carbon atoms. The increasing oxygen content and the elevated O1/C1 ratios observed for PPE–P(py), PPE–P(py)–rGO, and PPE–P(py)–rGO–Fe2O3 (measured as 0.114, 0.205, and 0.285, respectively) demonstrated the progressive formation of oxygen–carbon bonds or oxygen-containing functional groups, such as C–O/CO bonds, and the incorporation of oxygen into the carbon-based materials through interactions with the oxygen atoms present in Fe2O3 in each modification step. These findings suggest that oxygen-rich groups are generated on the PPE surface during layering. Conversely, the N1/C1 ratios showed a decreasing trend with each modification, with values of 0.026, 0.01, and 0.0095 for PPE–P(py), PPE–P(py)–rGO, and PPE–P(py)–rGO–Fe2O3, respectively. This decrease in the N1/C1 ratio indicated a reduction in the nitrogen content on the PPE surface during each successive modification step.15,43,44
The deconvoluted C1s spectrum of the PPE–P(py) sample exhibited binding energy peaks at 289 and 291 eV, indicating the presence of C–C/C–H bonds and π–π transitions.45 However, in the case of PPE–P(py)–rGO and PPE–P(py)–rGO–Fe2O3, the deconvoluted C1s peaks shifted to 292 eV and 294 eV, respectively. The observed higher binding energy shifts are associated with functional groups on rGO and the charge distribution on P(py), which influence the electron density of the carbon atoms.46 The deconvoluted O1s spectra for all three modifications (PPE–P(py), PPE–P(py)–rGO, and PPE–P(py)–rGO–Fe2O3) showed peaks at 537 and 538 eV, indicating the presence of –COOH and –C–O–C groups. However, in the O1s spectrum of PPE–P(py)–rGO, additional binding energy peaks at 535 eV corresponded to –C–OH groups.15 In the deconvoluted N1s spectra, binding energy peaks at 404 eV and 405 eV were observed, which correspond to nitrogen defects/graphitic nitrogen and oxidized nitrogen species.45,47,48 Additionally, the N1s spectrum of PPE–P(py)–rGO exhibited an additional binding energy peak at 402 eV, indicating the presence of imine groups. Regarding the Fe (iron) spectrum, the XPS binding energy peaks observed at 716 eV, 719 eV, and 729 eV indicated the presence of iron in different oxidation states, including Fe2+ and Fe3+.47,49 The peak at 716 eV is characteristic of the Fe3+ form of Fe2O3. These binding energy peaks arise from interactions between iron and other elements or functional groups, which influence the electron density and shift in the binding energy. The presence of both Fe2+ and Fe3+ states, as indicated by the Fe 2p XPS spectrum and the peak at 729 eV (corresponding to the Fe 2p1/2 core level binding energy), supports the identification of iron(III) oxide (Fe2O3) in the analysed sample.50 The increase in the O1/C1 ratio and the decrease in the N1/C1 ratio confirmed the successful layer-by-layer modification of the PPE. These changes in elemental composition and surface chemistry indicate the successful incorporation of P(py), rGO, and Fe2O3 nanoparticles on the PPE, leading to potentially enhanced properties and new applications for the resulting nanocomposite materials. These combined EDS and XPS results have provided strong evidence for the successful PPE–P(py)–rGO–Fe2O3 sensor fabrication.
To determine the electroactive area of the electrode, we employed the Randles–Sevcik equation for quasi-reversible electrochemical systems,15,54
Furthermore, PPE–P(py)–rGO–Fe2O3 was evaluated for its electrochemical behaviour using cyclic voltammetry with 1 mM 5-HT in 0.1 M PB solution at pH 7.4 and a scan rate of 0.05 V s−1. The cyclic voltammograms obtained for PPE, PPE–P(py), PPE–P(py)–rGO, and PPE–P(py)–rGO–Fe2O3 are summarized in Fig. 4(B). The unmodified PPE exhibited an ill-defined peak meant for 5-HT oxidation, and on the other hand, PPE–P(py) displayed a remarkable oxidation peak with increased peak current. The oxidation peak current intensity was further enhanced for PPE–P(py)–rGO, which was further increased for PPE–P(py)–rGO–Fe2O3, indicating the importance of rGO and Fe2O3 in the electrochemical detection of 5-HT.55 Combining P(py) with rGO and Fe2O3 nanoparticles resulted in a 2.4-fold increase in oxidation peak current intensity compared to PPE–P(py) and a 10.8-fold increase compared to the PPE, attributed to the good electron transport behaviour of P(py) and rGO. Moreover, the 5-HT response of PPE–P(py)–rGO–Fe2O3 was 1.3 times higher than that of PPE–P(py)–rGO, demonstrating the strong electrocatalytic behaviour of Fe2O3. The anodic peak potential (Epa) of 5-HT at PPE–P(py)–rGO–Fe2O3 was observed to be 0.28 V. The peak current is due to the oxidation reaction of 5-HT to 5-HT quinone imine. Here, 5-HT, an electro-active neurotransmitter, can undergo electrochemical oxidation in a solvent with physiological pH.
In general, the oxidation process involves a two-step mechanism, characterized by two-electron, two-proton transfer,56 wherein 5-HT is initially oxidized to form a carbocation, followed by a subsequent oxidation step leading to the formation of quinone imine, as depicted in Scheme 3.56
The remarkable enhancement of the electrocatalytic current observed at the PPE–P(py)–rGO–Fe2O3 electrode can be attributed to the catalytic properties exhibited by Fe2O3. Fe2O3 facilitates electron transfer between the electrode and 5-HT molecules, facilitating their oxidation. By accepting electrons from the electrode, Fe2O3 promotes the transfer of electrons to the 5-HT molecules, thereby promoting their oxidation process. Moreover, the paper electrode modified with P(py)–rGO indeed plays an important role in the electrode performance towards 5-HT. The P(py)–rGO composite provides a conductive pathway for electron transfer from the electrode to the Fe2O3 nanoparticles and a high surface area for the adsorption of serotonin molecules and their subsequent oxidation. The electrochemical oxidation of 5-HT at the electroactive sites of Fe3+ results in the formation of quinone imine, generating Fe2+ ions and resulting in an increased anodic peak current.57 The increase in oxidation current for 5-HT observed at the Fe2O3-modified electrode (rGO and P(py)), serving as a catalyst, promotes the electrochemical oxidation reaction and facilitates electron transfer. Iron(III) oxide (Fe2O3) is a good catalyst for the oxidation of 5-hydroxytryptamine (5-HT) to quinone imine. This is because the Fe(III)/Fe(II) ion center in Fe2O3 can catalyse the oxidation of 5-HT.58 The observed oxidation peak for 5-HT is due to the two-electron oxidation of 5-HT. The electrochemical response of the 5-HT electrode involves a two-step process: an initial electrochemical oxidation followed by a subsequent chemical reaction. In the first step, ferrous ions (Fe(II)) undergo electrochemical oxidation to form ferric ions (Fe(III)) (eqn (1)). This electrochemical process is responsible for the generation of the electrical signal. In the second step, the chemically reactive Fe(III) ions act as an oxidant, causing the oxidation of 5-hydroxytryptamine (5-HT) to form the corresponding quinone derivative (eqn (2)).
Fe(II) + 2e− → Fe(III) | (1) |
5-HT + Fe(III) → 5-HT quinone imine + Fe(II) + H2O | (2) |
A scan rate study was conducted to gain a deeper understanding of the electrochemical behaviour of 5-HT on the surface of PPE–P(py)–rGO–Fe2O3. The study aimed to investigate the influence of the scan rate on the oxidation peak current of 5-HT in a 0.1 M PB solution. The results, presented in Fig. 4(C), revealed a gradual increase in the anodic oxidation peak current (Ipa) as the scan rate ranged from 0.02 V s−1 to 0.12 V s−1. The plot of the scan rate against anodic peak current intensity (Fig. 4(D)) yielded a linear relationship. This observation suggests a surface-controlled electrochemical reaction60 on the PPE–P(py)–rGO–Fe2O3 sensor. From the above information, the concentration of the adsorbed analytes (C) on PPE–P(py)–rGO–Fe2O3 was calculated using the Brown–Anson model equation Ipa = (n2F2CAϑ)/4RT and was found to be 0.137 × 10−9 mole cm−2.61 Furthermore, the electronic transport properties of individual layers were validated by electrochemical impedance spectroscopy (EIS) in a 0.1 M KCl solution containing 5 mM [Fe(CN)6]3−/4− at open circuit potential (OCP) over a frequency range of 100 mHz to 1 MHz (Fig. 5(A)). The Randles-equivalent circuit is a model that consists of four components: solution resistance (Rs), double-layer capacitance (Cdl), charge transfer resistance (Rct), and Warburg impedance (Zw). The initial point of the semicircle in the Nyquist plot corresponds to solution resistance (Rs), while the diameter of the semicircle represents the charge transfer resistance (Rct) (Fig. 5(A)). The bare electrode displayed a higher charge transfer resistance (1317.48 Ω), while PPE–P(py)–rGO–Fe2O3 showed a lower Rct (34.61 Ω), indicating its excellent conductivity due to high electron transfer efficiency. Rct is mathematically related to the thermodynamic gas constant (R), the absolute temperature (T), the standard heterogeneous electron transfer rate constant (K0), the number of transferred electrons (N), Faraday's constant (F), the electrode surface area (A), and the concentration of the redox species in the solution (C), as expressed by the equation, Rct = RT/K0N2F2AC.61 The calculated K0 values of PPE and PPE–P(py)–rGO–Fe2O3 were found to be 0.005 × 10−4 and 0.6 × 10−4, respectively. The observed combination of a low charge-transfer resistance (Rct) value and a high standard heterogeneous electron transfer rate constant (K0) value can be attributed to a synergistic effect that enhances the sensor's performance, making it highly suitable for 5-HT detection.38,62
Evaluating the anti-interference capability of PPE–P(py)–rGO–Fe2O3 is essential for its successful application in the electrochemical analysis of real samples. Optimization experiments were conducted to investigate the effects of potential interferences on the measurement of 5-HT at a concentration of 50 μM, including 50 μM concentration of ascorbic acid (AA), uric acid (UA), dopamine (DA), and epinephrine (E). The performance of the modified electrode was assessed, and the resulting data were carefully analysed and summarized in Fig. 6(B) and Fig. S3A (ESI†). Interestingly, adding interfering chemicals had a minimal impact on the anodic peak current response of the modified electrode, indicating its excellent anti-interference properties. However, the presence of interfering E showed a slight variation in peak current responses. Interestingly, the sensor exhibited dual detection behaviour when tested with 5-HT and DA, where simultaneous detection of both biomolecules was possible without any substantial change in the oxidative peak current for 5-HT (Fig. S3A, ESI†).65,66,67 Despite the slight effect, the modified electrode still demonstrated remarkable anti-interference properties, indicating its potential for accurate and reliable electrochemical measurements of 5-HT in the presence of interfering species. It should be noted that Fe2O3 nanoparticles are known to have a high affinity for serotonin.57 Fe2O3 has a strong electrostatic interaction with the positively charged amino group on the serotonin molecule. The modified electrode surface exhibited electrostatic repulsive behaviour towards other negative analytes such as AA and UA. Unlike DA, epinephrine showed slight interference with 5-HT detection (Fig. S3A, ESI†). Interestingly the DA oxidation occurred at a much lower potential than the common oxidation potential known for DA. This could be attributed to the chemical modification of the electrode surface with P(py)–rGO and Fe2O3. The selective electrochemical oxidation process of 5-HT at the modified electrode is similar to the oxidation of folic acid reported at the Fe2O3 modified electrode.58,67
Furthermore, the findings of the modified electrode were validated by comparing them with the results obtained from the high-performance liquid chromatography (HPLC) method (Fig. S6, ESI†). The concentration of the analyte, in this case, was determined to be 32 μM. The close agreement between the results obtained for the modified electrode and the HPLC method showcased the accuracy and reliability of the modified electrode in quantifying 5-HT in real samples (Table 1). In order to check the analytical feasibility, the spike sample analysis in Table 2 depicts the mean obtained recoveries of 101.05%, 98.22%, and 100.78% for 5-HT analysis, accompanied by the respective relative standard deviation (RSD) of 0.054%, 0.99%, and 1.32%. These findings confirmed the exceptional precision of the newly developed electrode in quantifying 5-HT. As a result, the developed sensor exhibits significant potential for precisely identifying 5-HT in human samples, rendering it an appealing candidate for potential applications in analytical chemistry and biosensing.
Real sample | Determined concentration by the EC method | Determined concentration by the HPLC-UV method | Recovery |
---|---|---|---|
Brain sample of Drosophila melanogaster | 33.3 μM | 31.84 μM | 95.6% |
Sample | Added amount (μM) | Found amount (μM) | Recovery (%) | RSD (%) |
---|---|---|---|---|
Drosophila brain sample | — | 33.3 | 95.6 | 0.13 |
1st spike | 10 | 42.1 | 97.2 | 0.054 |
2nd spike | 20 | 53.2 | 99.8 | 0.99 |
3rd spike | 30 | 64.7 | 102.2 | 1.32 |
PPE | Paper electrode |
P(py) | Poly(pyrrole) |
rGO | Reduced graphene oxide |
GO | Graphene oxide |
5-HT | 5-Hydroxytryptamine |
DPV | Differential pulse voltammetry |
HPLC-UV | High-performance liquid chromatography with ultra-violet spectroscopy |
SS | Serotonin syndrome |
5-HT-1A | Serotonin 1A receptor |
5-HT-2A | Serotonin 2A receptor |
EG-Au-gate-FET | An extended-gold-gate-field effect transistor |
PME | Polymer-modified electrode |
ZnONR | Zinc oxide nanorods |
PMB (DES) | Polymethylene blue (deep eutectic solvent) |
FeC | Ferrocene |
MWCNTs | Multiwalled carbon nanotubes |
AuNPs | Gold nanoparticles |
Ag2Se | Silver selenide |
Fe2O3 | Ferric oxide |
EIS | Electrochemical impedance spectroscopy |
CV | Cyclic voltammetry |
ATR-FTIR | Attenuated total reflectance-Fourier transform infrared spectroscopy |
FE-SEM | Field emission scanning electron microscopy |
EDS | Energy dispersive X-ray spectroscopy |
XPS | X-ray photoelectron spectroscopy |
PB | Phosphate buffer |
ECE | Electron transfer-chemical reaction-electron transfer mechanism. |
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d3ma00777d |
This journal is © The Royal Society of Chemistry 2024 |