Next Article in Journal
High-Performance Sulfur Dioxide Gas Sensor Based on Graphite-Phase Carbon-Nitride-Functionalized Tin Diselenide Nanorods Composite
Next Article in Special Issue
Recent Advance in Cortisol Immunosensing Technologies and Devices
Previous Article in Journal
Recent Developments in Rhodamine-Based Chemosensors: A Review of the Years 2018–2022
Previous Article in Special Issue
Effects of Acidic Solution on the One-Step Electrodeposition of Prussian Blue Nanocrystals on Screen-Printed Carbon Electrodes Modified with Magnetite Nanoparticles
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Novel Electrochemical Sensing Platform for the Detection of the Antidepressant Drug, Venlafaxine, in Water and Biological Specimens

1
Department of Chemistry, Quaid-i-Azam University, Islamabad 45320, Pakistan
2
National Centre of Excellence in Physical Chemistry, University of Peshawar, Peshawar 25120, Pakistan
3
Institute of Chemical Sciences, Bahauddin Zakaryia University, Multan 6100, Pakistan
4
Department of Chemistry, College of Science, United Arab Emirates University, Al Ain P.O. Box 15551, United Arab Emirates
*
Author to whom correspondence should be addressed.
Chemosensors 2022, 10(10), 400; https://doi.org/10.3390/chemosensors10100400
Submission received: 27 August 2022 / Revised: 26 September 2022 / Accepted: 30 September 2022 / Published: 4 October 2022
(This article belongs to the Collection Electrochemical Biosensors for Medical Diagnosis)

Abstract

:
A stable bimetallic catalyst composed of Co–Pd@Al2O3 was synthesized using a wet impregnation method, followed by calcination and H2 reduction. The synthesized catalyst was thoroughly characterized using XRD, BET, SEM, EDX, and TPR techniques. The catalyst was then drop-casted on a glassy carbon electrode (Co–Pd@Al2O3/GCE) and applied for the sensitive and selective electrochemical determination of a common antidepressant drug, venlafaxine (VEN). The proposed sensor (Co–Pd@Al2O3/GCE) demonstrated a remarkable catalytic activity for the electro-oxidation of VEN, with a decent repeatability and reproducibility. The pH dependent responsiveness of the electro-oxidation of VEN helped in proposing the redox mechanism. A linear relationship between the peak current and concentration of VEN was observed in the range of 1.95 nM to 0.5 µM, with LOD and LOQ of 1.86 pM and 6.20 pM, respectively. The designed sensor demonstrated an adequate selectivity and significant stability. Moreover, the sensor was found to be quite promising for determining the VEN in biological specimens.

1. Introduction

Venlafaxine (VEN), chemically known as 1-[2-(dimethylamino)-1-(4-methoxyphenyl) ethyl] cyclohexanol, is a member of a class of drugs known as selective serotonin and norepinephrine reuptake inhibitors (SSNRIs). It is extensively prescribed as an antidepressant to adults and children with nightmares, cataplexy, and narcolepsy [1]. The commonly recommended dosage of VEN ranges from 75 to 375 mgs per day. Over dosages may lead to cardiac arrhythmias, serotonin syndrome, depression, coma, hypertension, or hypotension, and may even cause death [2]. Thus, assessment of the toxicity, interactions, and remedial proficiency, as well as observation of the level of the drug in bio fluid, has become clinically and medically necessary [3]. A survey of the literature has revealed chromatography to be a leading technique for the trace level sensing of VEN [4,5,6,7,8,9,10]. The broadly used approaches for VEN studies in bio-specimens include high-performance liquid chromatography (HPLC), along with ultraviolet detection [11,12], liquid chromatography/mass spectrometry (LC/MS) [13], gas chromatography (GC) [14], solid-phase extraction [15], capillary electrophoresis [16], and electrochemical methods [17,18,19,20]. Although chromatographic approaches are quite sensitive and proficient, these approaches are quite time-consuming and thus cannot be used for efficient drug analysis [21]. Furthermore, chromatographic approaches have a higher on-column packing of the analytes at the LOQ level, which might decrease the column efficiency and shorten the lifetime of the column. In contrast, electro-analytical methods are very efficient, sensitive, and selective, and do not require any pre-treatment or any type of pre-separation [22,23]. The accessibility of several chemically fabricated electrodes at a comparatively lower cost provides a significant incentive for the determination of drug molecules using electroanalytical techniques. Thus, we took interest in developing a somewhat simple, sensitive, and accurate alternate scheme for the determination of VEN.
The modification of electrode surfaces plays a great role in electron transfer kinetics and electrocatalytic studies. Modified electrodes are widely applied in electrocatalytic studies, electrochemical devices, energy conversion, and chemical analysis [24]. Electrode modification with metal nanoparticles (NPs) has presented an excellent electrochemical activity owing to the optimum surface area, maximum metal dispersion, controlled morphology, excellent durability, chemical inertness, and good conductive properties [25,26,27]. In addition, the preparation of electrochemical sensors by modifying the electrode surface with NPs has been of great interest for the sensing of trace amount of analytes, which can be attributed to the amplification of electrochemical signals, leading to ultrasensitive assessments. NP-based electrochemical sensors have been developed either through the direct use of NPs as electroactive species or in the form of semiconductors to enhance electrochemical detection [28,29]. NPs are deposited on the electrode surface in order to load the electroactive species for improving the detection limit of molecules [30,31,32,33]. In this scenario, the modification of electrode surfaces with metal catalysts has shown an excellent electroactivity. Metal NPs are highly beneficial for electrocatalytic studies due to their high surface area, maximum metal dispersion, controlled particle size, morphology of the attached nanoparticles, good electrical characteristics, and stability [34,35]. The synergy of two metals in bimetallic catalysts sometimes results in a better activity, good electrical conductance, excellent stability, and better mass/electron transfer properties [36,37]. Among the various supports, gamma alumina (γ-Al2O3) offers a large surface area, mesoporous character, narrow pore-size distribution, and good mechanical characteristics [38]. The oxide groups of alumina have a proton-donating nature. Hence, alumina-coated glassy carbon electrodes (GCE) can overcome the overpotential and improve the sensitivity of many materials [39,40]. The modified electrodes with metal NPs such as Pd-Ni/C [41], Au/C [42,43,44,45], Ni/graphite [46,47], and Pt–Co/CNT [48] are widely used for various electrochemical applications. In recent years, most studies have focused on VEN electrochemical sensing using various metal oxides [49,50,51,52], whereas metal NP-based electrochemical sensors for VEN studies are rather scarce. Hence, in the current article, we report the synthesis of Co–Pd@Al2O3 bimetallic catalysts and their deposition on GCE for designing a voltammetric sensor to detect venlafaxine in human serum samples. To the best of our knowledge, no previous reports have been carried out on the quantitation of venlafaxine using a Co–Pd@Al2O3 bimetallic catalyst. The designed sensing platform presented an excellent performance towards venlafaxine oxidation, with a wide linear range and low detection limit, which might be due to the high electrical conductivity and synergistic effect of the bimetallic NPs.

2. Experiment

2.1. Reagent and Apparatus

Electrochemical studies were carried out on a Metrohm Autolab PGSTAT302N (Herisau, Switzerland) using NOVA 1.11 and FRA software. The conventional three-electrode system was used for electrochemical testing. A bare and modified glassy carbon electrode was used as a working electrode, Ag/AgCl (3M KCl) as a reference electrode, and Pt wire as a counter electrode. pH measurements were carried out using the INOLAB pH meter.
Venlafaxine hydrochloride (Fluka), cobalt acetate tetrahydrate (Co (C2H3O2)2·4H2O, Aldrich, 99.9%), and palladium chloride (PdCl2, Aldrich, 99%) were used as received, without any purification. Al2O3 powder and Nafion were purchased from Sinopharm Chemical Reagent Co., Ltd., (Shanghai, China). Deionized water was used in the preparation of all aqueous solutions.

2.2. Catalyst Preparation

The alumina support was prepared using the precipitation method, as described earlier [53] and as shown in Scheme 1. After calcination at 550 °C for 3 h, alumina was ground finely into a fine powder form. The Co–Pd@Al2O3 bimetallic catalyst was prepared via the impregnation method with a total metal loading of 10 wt%, as presented in Scheme 2. For this purpose, Al2O3 powder was mixed with the aqueous solution of cobalt acetate and palladium chloride. The concentration of both metal precursors was adjusted to obtain the desired metal loading of the catalyst. The solution was stirred continuously for 5 h at 25 °C for thorough metal loading. Then, the suspension was concentrated under vacuum conditions. Finally, the H2 reduction of the catalyst was conducted at 550 °C for 4 h.

2.3. Electrode Modification

First, the GCE surface was thoroughly cleaned. The modification of bare GCE was performed by coating Co–Pd@Al2O3 NPs according to the following procedure: 1 mg of catalyst powder was dispersed in 5 μL ethanol via ultra-sonication for 20 min. Then, 5 μL of the prepared suspension was drop-casted onto the electrodes, followed by two to three drops of Nafion solution. The modified electrode was ready for use after being completely dried at room temperature.

2.4. Characterization

The elemental analysis was performed by the atomic absorption spectrophotometry (AAS) technique by running the samples on the Perkin Elmer AA400. X-ray powder diffraction (XRD) analysis was performed on an analytical diffractometer (Rigaku D/Max 2500) with a graphite monochromatic radiation source Cu Kα1 (λ = 1.54056 Å). The morphology of the catalyst was examined using a scanning electron microscope (SEM) (MIRA3, TESCAN) equipped with an EDX (energy dispersive spectroscopy) unit for the chemical analysis. The surface area analysis was measured on the basis of the BET (Brunauer, Emmett, and Teller) method by using the Sorptometer Kelvin 1042 instrument after degassing the sample at 120 °C for 1 h.

3. Results and Discussion

Elemental analysis of the synthesized catalyst by AAS revealed a Co metal content of 14.11 wt% and Pd content of 5.82 wt%. These results were found to be in good agreement with the theoretical metal loadings (Co = 15 wt%, Pd = 5 wt%), thus presenting an effective procedure for catalyst preparation.
Moreover, EDX analysis was conducted to ensure the elemental composition of the catalyst. Figure 1a shows the EDX spectrum, in which the co-existence of Co and Pd metals ensured the formation of Co–Pd alloy NPs. The EDX analysis of the prepared catalyst detected Co = 13.34 wt%, Pd = 4.74 wt%, O = 47.97 wt%, and Al = 33.95 wt%. The EDX composition was found to be close to the AAS results, thus confirming the metallic stoichiometric contents. The surface morphology of the prepared Co–Pd nanoparticles dispersed on alumina was analyzed by SEM. The SEM image (Figure 1b) of the prepared catalyst presented the formation of spherical Co–Pd alloy NPs. The metal particles were found uniformly distributed over the surface of the Al2O3. The size of the bimetallic NPs varied, ranging from 45 to 85 nm. In addition, EDX mapping was recorded (Figure 1c), where the uniform distribution of the metal NPs was observed over the alumina support.
The XRD analysis was conducted to study the crystalline features of the prepared sample. Figure 2 presents the XRD profile of the Co–Pd@Al2O3 catalyst. The significant peaks appeared at 37.3°, 45.7°, and 67.1°, corresponding to the hkl values of (311), (400), and (440), respectively. These values are consistent with the standard pattern of Al2O3 (ICDD card no. 00-001-1303), showing a cubic crystal structure. Moreover, two additional broad peaks appeared at 32.9° and 59.6°, which were consistent with the standard pattern of Co (JCPDS:15-0806), with hkl values of (210) and (411), respectively. Small peaks were noticed at 39.2° and 55.7° for the Pd metal, which may either be due to its low content or the metal particles being too small, with corresponding hkl plans of (111) and (200), matching the reference card (JCPDS: 46-1043). Broad diffraction peaks showed the nanosized nature of the product, with an average crystallite size of 18 nm, as computed by the Debye–Scherrer formula.
Figure 3a presents the adsorption–desorption isotherms of Co–Pd/Al2O3 with a type IV hysteresis loop, which represented the mesoporous nature of the material [54,55,56]. The BET area of the catalysts was found to be 112 m2 g−1, which was much lower than Al2O3 (205 m2 g−1). The decrease in the BET area can be related to the filling of Al2O3 pores with metal particles as well as the sintering process. Figure 3b presents the distribution of pores in a narrow range with a maximum pore size of 3.3 nm in diameter.
Temperature programmed reduction (TPR) measurements were performed to study the reduction behavior of metal alloy NPs in bimetallic systems. Figure 4 presents the TPR profile of the Co–Pd@Al2O3 bimetallic catalyst. The Co/Al2O3 monometallic catalyst had a strong interaction with the Al2O3 support, and therefore showed reduction peaks in the high-temperature region. Upon the incorporation of Pd metal, the reduction peaks were shifted towards a lower temperature region. The alloying of Pd and Co metals resulted in an improvement in the catalytic activity that could be related to an easing of reducibility, better metal dispersion, and smaller particle size.

3.1. Electrochemical Characterization

The proficient electron transfer capability of Co–Pd@Al2O3-modified GCE was verified in the electrochemical impedance spectroscopy (EIS) studies using a 5.0 mM solution of K3[Fe(CN)6] in 1.0 M KCl solution. The Nyquist plots are shown in Figure 5, the data of which have been analyzed and fitted using the analogous Randles circuit (displayed in the inset). Rs is the electrolyte resistance, Rct is the charge (electron) transfer resistance, and W is the Warburg impedance. A semicircle at a higher frequency is linked with an interfacial charge transfer process and its diameter refers to the charge transfer resistance (Rct), while the linear part at lower frequency is related to diffusion processes [57,58]. Rct values for bare GCE and modified GCE were recorded as 6.41 kΩ and 3.33 kΩ, respectively. The decrease in Rct values suggests an enhancement in the charge transfer process for Co–Pd@Al2O3-modified GCE.
The electron transfer process of Co–Pd@Al2O3-modified GCE was also studied by cyclic voltammetry (CV) using a 0.1 M KCl solution containing 5.0 mM K3[Fe(CN)6] as a model redox couple at a scan rate of mV s−1. Figure 6 presents the cyclic voltammograms with a reasonably reversible redox behavior. For Co–Pd@Al2O3-modified GCE, the sharp peak current with a narrow peak potential difference and an obvious enhancement in the oxidation–reduction was observed in comparison with bare GCE, signifying a rapid electron transfer process. In order to demonstrate that Co–Pd@Al2O3 improves the surface area and conductivity of the electrode, the Randles–Sevcik equation was used (Ip = 2.6 × 105 n3/2 D1/2 ν1/2 AC). The surface area was calculated to be 0.02 and 0.04 cm2 for bare GCE and Co–Pd@Al2O3-modified GCE, respectively. A two-fold increase in the surface area compared with the unmodified electrode led to an enhanced current intensity of the redox probe.

Square Wave Anodic Stripping Voltammetric Analysis of the Targeted Analyte

By using SWASV, an electro-oxidation response of VEN was obtained at a potential of 0.65 V, as presented in Figure 7. A clearly enhanced current response for the electro-oxidation of VEN was obtained at Co–Pd@Al2O3/GCE compared with the bare GCE. This much boosted current response can be ascribed to the decent conductivity of Co–Pd@Al2O3/GCE towards the facile oxidation of VEN oxidation. An observation of Figure 6 reveals that the active surface area increased two times, while Figure 7 shows that the peak current of the analyte at the modified electrode increased by about three times compared with the bare electrode. Hence, the higher surface area of the modified electrode accompanied with the catalytic role of the modifier is responsible for the peak current amplification. The obtained outcomes confirmed that the proposed sensor is a suitable platform for VEN detection.

3.2. Optimization of the Experimental Parameters

3.2.1. Influence of the Supporting Electrolyte and pH

The effects of several buffer solutions, including the Britton–Robinson buffer with a pH ranging from 6 to 11, acetate buffer solution, and phosphate buffer solutions, were studied. Among the selected buffer solutions, BRB produced the finest response in terms of a higher peak current signal and well-defined peaks for the oxidation of VEN. The pH effect on the oxidation peak current of VEN on the Co–Pd@Al2O3/GCE was also studied at a pH ranging from 6 to 11. The best current signal was found at a pH of 7, as shown in Figure 8A.
As shown in Figure 8B, by increasing the pH, the peak shifted toward a more negative potential. As the Nernstian dependency of the peak potential on the pH was observed and the electrooxidation of VEN is well known to involve a two-electron transfer [33], the number of involved protons could be estimated as two, which is consistent with the following proposed mechanism (Scheme 3).

3.2.2. Influence of the Deposition Potential and Deposition Time

The effect of the deposition potential and deposition time at Ip of 2 μM VEN was also analyzed via SWASV in order to understand the adsorption process of VEN at the electrode surface. The highest Ip was obtained at a deposition potential of −0.5 V and at a deposition time of 60 s, as shown in Figure 9. Thus, further studies were carried out at this optimized deposition potential and deposition time.

3.3. Analytical Characterization

The proposed Co–Pd@Al2O3/GCE was used for the determination of various concentrations of VEN in order to find the lowest detection limit using SWASV. The voltammograms obtained for the different concentrations of VEN are presented in Figure 10. A linear calibration plot was obtained for a concentration between 1.95 nM and 0.5 µM, as shown in Figure 10 (inset). The calibration plot was then used to determine the limit of detection (LOD = 3 s/m) and the limit of quantification (LOQ = 10 s/m) of VEN. The current values (n = 11) of the blank solution at the peak position of VEN were used to compute the standard deviation [48]. The computed detection and quantification limits of VEN were 1.86 pM and 6.20 pM, respectively.
A comparison of the proposed Co–Pd@Al2O3/GCE electrode with the previously reported modified electrodes for the determination of VEN is shown in Table 1 [4,49,59,60,61,62,63,64]. The data show that our designed electrochemical sensor had a lower limit of detection and good linearity range, which are considered to be two important figures of merit for an effective sensing platform. Thus, the proposed sensor was found to be efficient for the detection of VEN.

3.3.1. Repeatability, Stability, and Reproducibility of Co–Pd@Al2O3/GCE

The repeatability of the fabricated electrode was examined under the optimized conditions by five successive determination runs of the same Co–Pd@Al2O3/GCE in the BRB with a pH of 7 and the 0.5 µM VEN solution, and at the end of each determination, the current was recorded. The results obtained showed the relative standard deviation (%RSD) in an acceptable range of less than 1%. The reproducibility of the electrode was also examined by employing four separately fabricated electrodes for the detection of 0.5 µM VEN under the same conditions, and the %RSD was found to be in the tolerable range, i.e., below 5%. The stability of the fabricated electrode was inspected for 10 days. The electrode was implied for the SWASV detection of VEN after three days, six days, and nine days. The %RSD was found to be 4%, as also shown in Figure 11. The stability of the fabricated electrode was acceptable during this period. These outcomes indicate that the developed Co–Pd@Al2O3/GCE had a good repeatability, reproducibility, and stability in the synthesis procedure and voltammetric determination of VEN.

3.3.2. Interference Study

The selectivity of the fabricated GCE electrode was investigated for the determination of VEN, in the presence of potential interfering species that are commonly found in human plasma and pharmaceutical samples. The tolerance limit of these interfering substances was considered as the maximum concentration that gave a %RSD of less than 5% at a 0.5 µM concentration of VEN under optimized conditions. The results obtained showed that the 100-fold excess of citric acid, glucose, ascorbic acid, sucrose, and uric acid had no noteworthy influence on the oxidation peak current of VEN (RSD was found to be less than 5%). The results obtained also displayed excellent recoveries in the range of 95 to 99%, showing that the fabricated GCE electrode had great selectivity for the determination of VEN in biological samples in the presence of the most common interfering species as mentioned in Table 2.
We used the introduced technique to identify VEN in the serum specimens in order to assess its analytical value. The proposed method was employed for the determination of VEN in the test solution, with a known concentration of VEN added to the serum sample, followed by an evaluation of the percentage of recoveries using direct calibration. The recovery percentage of the VEN solution using Co–Pd@Al2O3/GCE in the serum samples under optimized conditions can be seen in Table 3.

4. Conclusions

Co–Pd@Al2O3 was employed for the first time as a modifier of the GCE surface. In an aqueous BRB (pH 7.0) solution, the modified electrode exhibited an excellent venlafaxine oxidation behavior. The quantitative measurement of VEN was achieved using a simple, quick, and sensitive SWASV technique. The EIS and CV results revealed rapid charge transport through Co–Pd@Al2O3/GCE in comparison with the bare GCE. Under optimized conditions, the detection limit of VEN was estimated to be 1.86 pM. The slope of the plot of the peak potential as a function pH revealed the involvement of an equal number of electrons and protons in the oxidation mechanism of VEN. The modified electrode was found to possess the qualities of anti-interference ability, repeatability, reproducibility, and stability. Moreover, it showed a wide linearity range for VEN detection. The performance of the designed sensor in aqueous and serum samples demonstrated the applicability of the adopted approach in the regular analysis of VEN samples. These figures of merit of the Co–Pd@Al2O3/GCE nanosensor for VEN detection point to its promising candidature for applications in clinical diagnostics.

Author Contributions

Conceptualization, S.S. and A.S.; methodology, S.S. and N.F.; formal analysis, S.S., N.F. and J.N.; investigation, S.S., J.N. and M.N.A.; resources, A.S. and I.S.; data curation, S.S., J.N. and I.S.; writing—original draft preparation, S.S.; writing—review and editing, A.S., I.S. and M.N.A.; visualization, S.S. and N.F.; supervision, A.S.; project administration, A.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding. The APC were kindly waived off by the journal.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

Authors gratefully acknowledge the support of Quaid-i-Azam University and the Higher Education Commission of Pakistan.

Conflicts of Interest

Authors declare no conflict of interest.

References

  1. Manjula, N.; Pulikkutty, S.; Chen, T.W.; Chen, S.M.; Liu, X. Hexagon prism-shaped cerium ferrite embedded on GC electrode for electrochemical detection of antibiotic drug ofloxacin in biological sample. Colloids Surf. A Physicochem. Eng. Asp. 2021, 627, 127129. [Google Scholar] [CrossRef]
  2. Aramesh Boroujeni, Z.; Aghbalaghi, Z.A. Electrochemical determination venlafaxine at NiO/GR nanocomposite modified carbon paste electrode. Iran. J. Chem. Chem. Eng. 2021, 40, 1042–1053. [Google Scholar]
  3. Rezvani Jalal, N.; Mehrbod, P.; Shojaei, S.; Labouta, H.I.; Mokarram, P.; Afkhami, A.; Madrakian, T.; Los, M.J.; Schaafsma, D.; Giersig, M.; et al. Magnetic nanomaterials in microfluidic sensors for virus detection: A review. ACS Appl. Nano Mater. 2021, 4, 4307–4328. [Google Scholar] [CrossRef]
  4. Maddah, B.; Cheraghveisi, M.; Najafi, M. Developing a modified electrode based on La3+/Co3O4Nanocubes and its usage to electrochemical detection of venlafaxine. Int. J. Environ. Anal. Chem. 2020, 100, 121–133. [Google Scholar] [CrossRef]
  5. El-Kashef, D.H.; Sharawy, M.H. Venlafaxine mitigates cisplatin-induced nephrotoxicity via down-regulating apoptotic pathway in rats. Chem. Biol. Intract. 2018, 290, 110–118. [Google Scholar] [CrossRef] [PubMed]
  6. Thompson, W.A.; Arnold, V.I.; Vijayan, M.M. Venlafaxine in embryos stimulates neurogenesis and disrupts larval behavior in zebrafish. Environ. Sci. Technol. 2017, 21, 12889–12897. [Google Scholar] [CrossRef] [PubMed]
  7. Parrott, J.L.; Metcalfe, C.D. Assessing the effects of the antidepressant venlafaxine to fathead minnows exposed to environmentally relevant concentrations over a full life cycle. Environ. Pollut. 2017, 229, 403–411. [Google Scholar] [CrossRef]
  8. Zucker, I.; Mamane, H.; Riani, A.; Gozlan, I.; Avisar, D. Formation and degradation of N-oxide venlafaxine during ozonation and biological post-treatment. Sci. Total Environ. 2018, 619, 578–586. [Google Scholar] [CrossRef]
  9. Fard, N.T.; Tadayon, F.; Panahi, H.A.; Moniri, E. The synthesis of functionalized graphene oxide by polyester dendrimer as a pH-sensitive nanocarrier for targeted delivery of venlafaxine hydrochloride: Central composite design optimization. J. Mol. Liq. 2022, 349, 118149. [Google Scholar] [CrossRef]
  10. Norouzi, P.; Larijani, B.; Alizadeh, T.; Pourbasheer, E.; Aghazadeh, M.; Ganjali, M.R. Application of advanced electrochemical methods with nanomaterial-based electrodes as powerful tools for trace analysis of drugs and toxic compounds. Curr. Anal. Chem. 2019, 2, 143–151. [Google Scholar] [CrossRef]
  11. Matoga, M.; Pehourcq, F.; Titier, K.; Dumora, F.; Jarry, C. Rapid high-performance liquid chromatographic measurement of venlafaxine and O-desmethylvenlafaxine in human plasma: Application to management of acute intoxications. J. Chromatogr. B Biomed. Appl. 2001, 760, 213–218. [Google Scholar] [CrossRef]
  12. Mandrioli, R.; Mercolini, L.; Cesta, R.; Fanali, S.; Amore, M.; Raggi, M.A. Analysis of the second generation antidepressant venlafaxine and its main active metabolite O-desmethylvenlafaxine in human plasma by HPLC with spectrofluorimetric detection. J. Chromatogr. B 2007, 856, 88–94. [Google Scholar] [CrossRef]
  13. Bhatt, J.; Jangid, A.; Venkatesh, G.; Subbaiah, G.; Singh, S. Liquid chromatography–tandem mass spectrometry (LC–MS–MS) method for simultaneous determination of venlafaxine and its active metabolite O-desmethyl venlafaxine in human plasma. J. Chromatogr. B 2005, 829, 75–81. [Google Scholar] [CrossRef]
  14. Mastrogianni, O.; Theodoridis, G.; Spagou, K.; Violante, D.; Henriques, T.; Pouliopoulos, A.; Raikos, N. Determination of venlafaxine in post-mortem whole blood by HS-SPME and GC-NPD. Forensic Sci. Int. 2012, 215, 105–109. [Google Scholar] [CrossRef]
  15. Madrakian, T.; Haryani, R.; Ahmadi, M.; Afkhami, A. Spectrofluorometric determination of venlafaxine in biological samples after selective extraction on the superparamagnetic surface molecularly imprinted nanoparticles. Anal. Method. 2015, 7, 428–435. [Google Scholar] [CrossRef]
  16. Guo, C.; Xiao, Y. Negatively charged cyclodextrins: Synthesis and applications in chiral analysis—A review. Carbohydr. Polym. 2021, 256, 117517. [Google Scholar] [CrossRef]
  17. Martins, F.C.; Pimenta, L.C.; De Souza, D. Antidepressants determination using an electroanalytical approach: A review of methods. J. Pharm. Biomed. Anal. 2021, 206, 114365. [Google Scholar] [CrossRef]
  18. Mobin, S.M.; Sanghavi, B.J.; Srivastava, A.K.; Mathur, P.; Lahiri, G.K. Biomimetic sensor for certain phenols employing a copper (II) complex. Anal. Chem. 2010, 82, 5983–5992. [Google Scholar] [CrossRef]
  19. Sanghavi, B.J.; Srivastava, A.K. Simultaneous voltammetric determination of acetaminophen, aspirin and caffeine using an in situ surfactant-modified multiwalled carbon nanotube paste electrode. Electrochim. Acta. 2010, 55, 8638–8648. [Google Scholar] [CrossRef]
  20. Gadhari, N.S.; Sanghavi, B.J.; Karna, S.P.; Srivastava, A.K. Potentiometric stripping analysis of bismuth based on carbon paste electrode modified with cryptand [2.2.1] and multiwalled carbon nanotubes. Electrochim. Acta 2010, 56, 627–635. [Google Scholar] [CrossRef]
  21. Görög, S. The changing face of pharmaceutical analysis. Trends Anal. Chem. 2007, 26, 12–17. [Google Scholar] [CrossRef]
  22. Sanghavi, B.J.; Srivastava, A.K. Adsorptive stripping differential pulse voltammetric determination of venlafaxine and desvenlafaxine employing Nafion–carbon nanotube composite glassy carbon electrode. Electrochim. Acta 2011, 56, 4188–4196. [Google Scholar] [CrossRef]
  23. Aftab, S.; Kurbanoglu, S.; Ozcelikay, G.; Bakirhan, N.K.; Shah, A.; Ozkan, S.A. Carbon quantum dots co-catalyzed with multiwalled carbon nanotubes and silver nanoparticles modified nanosensor for the electrochemical assay of anti-HIV drug Rilpivirine. Sens. Actuators B Chem. 2019, 285, 571–583. [Google Scholar] [CrossRef]
  24. Zen, J.M.; Senthil Kumar, A.; Tsai, D.M. Recent updates of chemically modified electrodes in analytical chemistry. Electroanalysis 2003, 15, 1073–1087. [Google Scholar] [CrossRef]
  25. Xu, J.; Tao, J.; Su, L.; Wang, J.; Jiao, T. A Critical Review of Carbon Quantum Dots: From Synthesis toward Applications in Electrochemical Biosensors for the Determination of a Depression-Related Neurotransmitter. Materials 2021, 14, 3987. [Google Scholar] [CrossRef]
  26. Shah, A.; Akhtar, M.; Aftab, S.; Shah, A.H.; Kraatz, H.B. Gold copper alloy nanoparticles (Au-Cu NPs) modified electrode as an enhanced electrochemical sensing platform for the detection of persistent toxic organic pollutants. Electrochim. Acta 2017, 241, 281–290. [Google Scholar] [CrossRef]
  27. Oyama, M. Recent nanoarchitectures in metal nanoparticle-modified electrodes for electroanalysis. Anal. Sci. 2010, 26, 1–12. [Google Scholar] [CrossRef] [Green Version]
  28. Gao, J.; Fu, L.J.; Zhang, H.P.; Yang, L.C.; Wu, Y.P. Improving electrochemical performance of graphitic carbon in PC-based electrolyte by nano-TiO2 coating. Electrochim. Acta 2008, 53, 2376–2379. [Google Scholar] [CrossRef]
  29. Raj, N.; Crooks, R.M. Detection Efficiency of Ag Nanoparticle Labels for a Heart Failure Marker Using Linear and Square-Wave Anodic Stripping Voltammetry. Biosensors 2022, 4, 203. [Google Scholar] [CrossRef]
  30. Xie, C.; Li, H.; Li, S.; Wu, J.; Zhang, Z. Surface molecular self assembly for organophosphate pesticide imprinting in electropolymerized poly (paminothiophenol) membranes on a gold nanoparticle modified glassy carbon electrode. Anal. Chem. 2010, 82, 241–249. [Google Scholar] [CrossRef]
  31. Wen, X.; Fei, J.; Chen, X.; Yi, L.; Ge, F.; Huang, M. Electrochemical analysis of trifluralin using a nanostructuring electrode with multi-walled carbon nanotubes. Environ. Pollut. 2008, 3, 1015–1020. [Google Scholar] [CrossRef] [PubMed]
  32. Uygun, Z.O.; Dilgin, Y. A novel impedimetric sensor based on molecularly imprinted polypyrrole modified pencil graphite electrode for trace level determination of chlorpyrifos. Sens. Actuators B Chem. 2013, 188, 78–84. [Google Scholar] [CrossRef]
  33. Zhang, L.; Lee, K.; Zhang, J. Effect of synthetic reducing agents on morphology and ORR activity of carbon-supported nano-Pd–Co alloy electrocatalysts. Electrochim. Acta 2007, 52, 7964–7971. [Google Scholar] [CrossRef]
  34. Welch, C.M.; Compton, R.G. The use of nanoparticles in electroanalysis: A review. Anal. Bioanal. Chem. 2006, 384, 601–619. [Google Scholar] [CrossRef]
  35. Firdous, N.; Janjua, N.K. CoPtx/γ-Al2O3 bimetallic nanoalloys as promising catalysts for hydrazine electrooxidation. Heliyon 2019, 5, e01380. [Google Scholar] [CrossRef] [Green Version]
  36. Wang, D.; Li, Y. Bimetallic nanocrystals: Liquid-phase synthesis and catalytic applications. Adv. Mater. 2011, 23, 1044–1060. [Google Scholar] [CrossRef]
  37. White, R.J.; Luque, R.; Budarin, V.L.; Clark, J.H.; Macquarrie, D.J. Supported metal nanoparticles on porous materials. Chem. Soc. Rev. 2009, 38, 481–494. [Google Scholar] [CrossRef]
  38. Trueba, M.; Trasatti, S.P. γ-Alumina as a support for catalysts: A review of fundamental aspects. Eur. J. Inorg. Chem. 2005, 2005, 3393–3403. [Google Scholar] [CrossRef]
  39. Windmiller, J.R.; Bandodkar, A.J.; Parkhomovsky, S.; Wang, J. Stamp transfer electrodes for electrochemical sensing on non-planar and oversized surfaces. Analyst 2012, 137, 1570–1575. [Google Scholar] [CrossRef]
  40. Dong, S.; Kuwana, T. Activation of glassy carbon electrodes by dispersed metal oxide particles: I. ascorbic acid oxidation. J. Electrochem. Soc. 1984, 131, 813. [Google Scholar] [CrossRef]
  41. Chowdhury, S.R.; Ghosh, S.; Bhattachrya, S.K. Enhanced and synergistic catalysis of one-pot synthesized palladium-nickel alloy nanoparticles for anodic oxidation of methanol in alkali. Electrochim. Acta 2017, 250, 124–134. [Google Scholar] [CrossRef]
  42. Li, H.; Xie, C.; Li, S.; Xu, K. Electropolymerized molecular imprinting on gold nanoparticle-carbon nanotube modified electrode for electrochemical detection of triazophos. Colloids Surf. B Biointerfaces 2012, 89, 175–181. [Google Scholar] [CrossRef]
  43. Dequaire, M.; Degrand, C.; Limoges, B. An electrochemical metalloimmunoassay based on a colloidal gold label. Anal. Chem. 2000, 72, 5521–5528. [Google Scholar] [CrossRef]
  44. Tominaga, M.; Taema, Y.; Taniguchi, I. Electrocatalytic glucose oxidation at bimetallic gold–copper nanoparticle-modified carbon electrodes in alkaline solution. J. Electroanal. Chem 2008, 624, 1–8. [Google Scholar] [CrossRef]
  45. Dervisevic, M.; Dervisevic, E.; Çevik, E.; Şenel, M. Novel electrochemical xanthine biosensor based on chitosan–polypyrrole–gold nanoparticles hybrid bio-nanocomposite platform. J. Food Drug Anal. 2017, 25, 510–519. [Google Scholar] [CrossRef] [Green Version]
  46. Hosseini, M.; Jafarian, T.; Rostami, M.; Gobal, F. A low cost and highly active non-noble alloy electrocatalyst for hydrazine oxidation based on nickel ternary alloy at the surface of graphite electrode. J. Electroanal. Chem. 2016, 763, 134–140. [Google Scholar]
  47. Tamašauskaitė-Tamašiūnaitė, L.; Rakauskas, J.; Balčiūnaitė, A.; Zabielaitė, A.; Vaičiūnienė, J.; Selskis, A.; Norkus, E. Gold–nickel/titania nanotubes as electrocatalysts for hydrazine oxidation. J. Power Sources 2014, 272, 362–370. [Google Scholar] [CrossRef]
  48. Jiang, S.; Ma, Y.; Jian, G.; Tao, H.; Wang, X.; Fan, Y.; Chen, Y. Facile construction of Pt–Co/CNx nanotube electrocatalysts and their application to the oxygen reduction reaction. Adv. Mater. 2009, 21, 4953–4956. [Google Scholar] [CrossRef]
  49. Jahani, P.M.; Javar, H.A.; Mahmoudi-Moghaddam, H. A new electrochemical sensor based on Europium-doped NiO nanocomposite for detection of venlafaxine. Measurement 2021, 173, 108616. [Google Scholar] [CrossRef]
  50. Zaimbashi, R.; Mostafavi, A.; Shamspur, T. ZnO nanoflower based electrochemical sensor for the selective determination of venlafaxine. J. Iran. Chem. Soc. 2021, 19, 1329–1337. [Google Scholar] [CrossRef]
  51. Ibrahim, H.; Temerk, Y. Novel sensor for sensitive electrochemical determination of luteolin based on In2O3 nanoparticles modified glassy carbon paste electrode. Sens. Actuators B Chem. 2015, 206, 744–752. [Google Scholar] [CrossRef]
  52. Beitollahi, H.; Jahani, S.; Tajik, S.; Ganjali, M.R.; Faridbod, F.; Alizadeh, T. Voltammetric determination of venlafaxine as an antidepressant drug employing Gd2O3 nanoparticles graphite screen printed electrode. J. Rare Earths 2019, 37, 322–328. [Google Scholar] [CrossRef]
  53. Parida, K.M.; Pradhan, A.C.; Das, J.; Sahu, N. Synthesis and characterization of nano-sized porous gamma-alumina by control precipitation method. Mater. Chem. Phys. 2009, 113, 244–248. [Google Scholar] [CrossRef]
  54. Jiang, Y.; Kang, Q.; Zhang, J.; Dai, H.B.; Wang, P. High-performance nickel–platinum nanocatalyst supported on mesoporous alumina for hydrogen generation from hydrous hydrazine. J. Power Sources 2015, 273, 554–560. [Google Scholar] [CrossRef]
  55. Wang, Z.; Xu, X.; Liu, Z.; Zhang, D.; Yuan, J.; Liu, J. Multifunctional Metal Phosphides as Superior Host Materials for Advanced Lithium-Sulfur Batteries. Eur. J. Chem. 2021, 27, 13494–13512. [Google Scholar] [CrossRef]
  56. Naik, B.; Prasad, V.S.; Ghosh, N.N. Development of a simple aqueous solution based chemical method for synthesis of mesoporous γ-alumina powders with disordered pore structure. J. Porous Mater. 2010, 17, 115–121. [Google Scholar] [CrossRef]
  57. Rudaz, S.; Stella, C.; Balant-Gorgia, A.E.; Fanali, S.; Veuthey, J.L. Simultaneous stereoselective analysis of venlafaxine and O-desmethylvenlafaxine enantiomers in clinical samples by capillary electrophoresis using charged cyclodextrins. J. Pharm. Biomed. Anal. 2000, 23, 107–115. [Google Scholar] [CrossRef]
  58. Vaze, V.D.; Srivastava, A.K. Electrochemical behavior of folic acid at calixarene based chemically modified electrodes and its determination by adsorptive stripping voltammetry. Electrochim. Acta 2007, 53, 1713. [Google Scholar] [CrossRef]
  59. Toan, T.T.T.; Dao, A.Q.; Vasseghian, Y. A state-of-the-art review on the nanomaterial-based sensor for detection of venlafaxine. Chemosphere 2022, 297, 134116. [Google Scholar]
  60. Morais, S.; Ryckaert, C.P.; Delerue-Matos, C. Adsorptive stripping voltammetric determination of venlafaxine in urine with a mercury film microelectrode. Anal. Lett. 2003, 36, 2515–2526. [Google Scholar] [CrossRef] [Green Version]
  61. Khalilzadeh, M.A.; Tajik, S.; Beitollahi, H.; Venditti, R.A. Green synthesis of magnetic nanocomposite with iron oxide deposited on cellulose nanocrystals with copper (Fe3O4@CNC/Cu): Investigation of catalytic activity for the development of a venlafaxine electrochemical sensor. Ind. Eng. Chem. Res. 2020, 59, 4219–4228. [Google Scholar] [CrossRef]
  62. Eslami, E.; Farjami, F. Voltammetric determination of venlafaxine by using multiwalled carbon nanotube-ionic liquid composite electrode. J. Appl. Chem. Res. 2018, 12, 42–52. [Google Scholar]
  63. Ding, L.; Li, L.; You, W.; Gao, Z.N.; Yang, T.L. Electrocatalytic oxidation of venlafaxine at a multiwall carbon nanotubes-ionic liquid gel modified glassy carbon electrode and its electrochemical determination. Croat. Chem. Acta 2015, 88, 81–87. [Google Scholar] [CrossRef]
  64. Senturk, H.; Karadeniz, H.; Erdem, A. Recent Applications of Nanomaterials Based on Electrochemical Drug Analysis. Curr. Org. Chem. 2021, 17, 1215–1228. [Google Scholar] [CrossRef]
Scheme 1. Preparation of the γ-Al2O3 support.
Scheme 1. Preparation of the γ-Al2O3 support.
Chemosensors 10 00400 sch001
Scheme 2. Preparation of the Co–Pd/γ-Al2O3 catalyst using the wet impregnation method.
Scheme 2. Preparation of the Co–Pd/γ-Al2O3 catalyst using the wet impregnation method.
Chemosensors 10 00400 sch002
Figure 1. (a) EDX profile, (b) SEM image, and (c) EDX mapping of the synthesized Co–Pd@Al2O3 bimetallic catalyst.
Figure 1. (a) EDX profile, (b) SEM image, and (c) EDX mapping of the synthesized Co–Pd@Al2O3 bimetallic catalyst.
Chemosensors 10 00400 g001
Figure 2. XRD pattern of the prepared Co–Pd@Al2O3 bimetallic catalyst.
Figure 2. XRD pattern of the prepared Co–Pd@Al2O3 bimetallic catalyst.
Chemosensors 10 00400 g002
Figure 3. (a) N2 adsorption–desorption isotherm and (b) pore size distribution plot of the Co–Pd@Al2O3 bimetallic catalyst.
Figure 3. (a) N2 adsorption–desorption isotherm and (b) pore size distribution plot of the Co–Pd@Al2O3 bimetallic catalyst.
Chemosensors 10 00400 g003
Figure 4. TPR profile of the prepared catalysts.
Figure 4. TPR profile of the prepared catalysts.
Chemosensors 10 00400 g004
Figure 5. Nyquist plots for bare GCE and Co–Pd@Al2O3/GCE in the 5 mM solution of the redox probe K3[Fe(CN)6].
Figure 5. Nyquist plots for bare GCE and Co–Pd@Al2O3/GCE in the 5 mM solution of the redox probe K3[Fe(CN)6].
Chemosensors 10 00400 g005
Figure 6. Cyclic voltammograms at bare GCE and Co–Pd@Al2O3/GCE in a 5 mM solution of K3[Fe(CN)6] in 0.1 M KCl.
Figure 6. Cyclic voltammograms at bare GCE and Co–Pd@Al2O3/GCE in a 5 mM solution of K3[Fe(CN)6] in 0.1 M KCl.
Chemosensors 10 00400 g006
Figure 7. (a) Blank voltammogram obtained at Co–Pd@Al2O3/GCE is for the buffered medium (BRB solution of pH of 7), (b) voltammogram of the 2 µM solution of VEN in BRB of pH 7 at bare GCE, (c) voltammogram of the 2 µM solution of VEN in BRB of pH 7 at Co–Pd@Al2O3/GCE. Conditions: scan rate of 50 mV/s; deposition potential of −0.5 V; deposition time of 60 s.
Figure 7. (a) Blank voltammogram obtained at Co–Pd@Al2O3/GCE is for the buffered medium (BRB solution of pH of 7), (b) voltammogram of the 2 µM solution of VEN in BRB of pH 7 at bare GCE, (c) voltammogram of the 2 µM solution of VEN in BRB of pH 7 at Co–Pd@Al2O3/GCE. Conditions: scan rate of 50 mV/s; deposition potential of −0.5 V; deposition time of 60 s.
Chemosensors 10 00400 g007
Figure 8. (A) Square wave voltammograms of the 2 µM solution of VEN in media of different pH obtained at a scan rate of 50 mV/s, accumulation potential of −0.5 V, and accumulation time of 60 s. (B) Plots of the peak potentials and peak currents versus pH.
Figure 8. (A) Square wave voltammograms of the 2 µM solution of VEN in media of different pH obtained at a scan rate of 50 mV/s, accumulation potential of −0.5 V, and accumulation time of 60 s. (B) Plots of the peak potentials and peak currents versus pH.
Chemosensors 10 00400 g008
Scheme 3. Proposed mechanism of the oxidation of VEN at the Co–Pd@Al2O3/GCE.
Scheme 3. Proposed mechanism of the oxidation of VEN at the Co–Pd@Al2O3/GCE.
Chemosensors 10 00400 sch003
Figure 9. (A) The effect of the deposition potential on the oxidation peak current of VEN, keeping the deposition time constant (60 s), and (B) the effect of the accumulation time on the oxidation peak current maintaining its accumulation potential (−0.5 V) for Co–Pd@Al2O3/GCE in the 2 μM solution of INZ by applying the SWASV technique.
Figure 9. (A) The effect of the deposition potential on the oxidation peak current of VEN, keeping the deposition time constant (60 s), and (B) the effect of the accumulation time on the oxidation peak current maintaining its accumulation potential (−0.5 V) for Co–Pd@Al2O3/GCE in the 2 μM solution of INZ by applying the SWASV technique.
Chemosensors 10 00400 g009
Figure 10. SWAS voltammograms of VEN in the concentration range of 0.5 µM to 1.95 nM at Co–Pd@Al2O3/GCE in a BRB solution with a pH of 7 under optimized conditions (scan rate of 50 mV/s, accumulation potential of −0.5 V, and accumulation time of 60 s) and (graph in the inset) the plot of Ipa versus the concentration of VEN.
Figure 10. SWAS voltammograms of VEN in the concentration range of 0.5 µM to 1.95 nM at Co–Pd@Al2O3/GCE in a BRB solution with a pH of 7 under optimized conditions (scan rate of 50 mV/s, accumulation potential of −0.5 V, and accumulation time of 60 s) and (graph in the inset) the plot of Ipa versus the concentration of VEN.
Chemosensors 10 00400 g010
Figure 11. Current response in the context of (A) repeatability, (B) fabrication reproducibility, and (C) stability of the proposed electrode.
Figure 11. Current response in the context of (A) repeatability, (B) fabrication reproducibility, and (C) stability of the proposed electrode.
Chemosensors 10 00400 g011
Table 1. The comparison of the reported detection limits of VEN with this work.
Table 1. The comparison of the reported detection limits of VEN with this work.
ElectrodesLinearity Range
(µM)
Limit of Detection
(µM)
References
La3+/Co3O4 nanocubes/SPE1–5000.5[4]
Eu3+ doped NiO/CPE0.04–3000.01[49]
NAF-CNT-GCE0.038–62.20.012[59]
Mercury film microelectrode1.27–24.30.69[60]
Fe3O4@CNC/Cu/GSPE0.05–6000.01[61]
MWCNT/CILE/CPE10–5000.47[63]
MWCNT-RTIL/GCE2–20001.69[64]
Gd2O3/SPE5–9000.21[52]
Co–Pd@Al2O3/GCE1.95 nM–0.5 µM1.86 pMPresent work
Table 2. Interference study in the presence of various interfering species commonly found in the pharmaceutical preparation of VEN.
Table 2. Interference study in the presence of various interfering species commonly found in the pharmaceutical preparation of VEN.
InterferentsVEN:InterferentsRecovery% ± RSD a
Citric Acid1:50
1:100
98.7% ± 0.45
99.0% ± 0.57
Glucose1:50
1:100
98.1% ± 0.39
98.3% ± 0.35
Ascorbic Acid1:50
1:100
97.1% ± 0.48
98.0% ± 0.43
Sucrose1:50
1:100
97.9% ± 0.46
98.1% ± 0.43
Uric Acid1:50
1:100
98.2% ± 0.49
98.7% ± 0.45
a: average of three readings.
Table 3. The recovery percentage of the VEN solution using Co–Pd@Al2O3/GCE in serum samples.
Table 3. The recovery percentage of the VEN solution using Co–Pd@Al2O3/GCE in serum samples.
SampleBefore AdditionAdded
(µM)
Found
(µM)
Recovery (%) ± RSD
Sample 1-0.500.49098.0 ± 0.48
Sample 2-0.250.24698.4 ± 0.49
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Sultan, S.; Shah, A.; Firdous, N.; Nisar, J.; Ashiq, M.N.; Shah, I. A Novel Electrochemical Sensing Platform for the Detection of the Antidepressant Drug, Venlafaxine, in Water and Biological Specimens. Chemosensors 2022, 10, 400. https://doi.org/10.3390/chemosensors10100400

AMA Style

Sultan S, Shah A, Firdous N, Nisar J, Ashiq MN, Shah I. A Novel Electrochemical Sensing Platform for the Detection of the Antidepressant Drug, Venlafaxine, in Water and Biological Specimens. Chemosensors. 2022; 10(10):400. https://doi.org/10.3390/chemosensors10100400

Chicago/Turabian Style

Sultan, Sundas, Afzal Shah, Naveeda Firdous, Jan Nisar, Muhammad Naeem Ashiq, and Iltaf Shah. 2022. "A Novel Electrochemical Sensing Platform for the Detection of the Antidepressant Drug, Venlafaxine, in Water and Biological Specimens" Chemosensors 10, no. 10: 400. https://doi.org/10.3390/chemosensors10100400

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop