Home An effective, novel, and cheap carbon paste electrode for naproxen estimation
Article Open Access

An effective, novel, and cheap carbon paste electrode for naproxen estimation

  • Mohamed Abd-Elsabour , Mortaga M. Abou-Krisha , Abdulrahman G. Alhamzani and Tarek A. Yousef EMAIL logo
Published/Copyright: August 19, 2022
Become an author with De Gruyter Brill

Abstract

Herein, a carbon paste electrode (CPE) modified with poly(reduced-o-nitrobenzoic acid [r-o-NBA]) supported in graphene quantum dots (GQDs) was fabricated for the first time. The fabricated electrode’s surface morphology and composition were characterised by scanning electron microscope and transmission electron microscope. The poly(r-o-NBA)/GQDs/CPE showed high electrocatalytic activity towards the oxidation of naproxen (NPX) using cyclic and differential pulse voltammetric methods. The effect of scan rate on the oxidation peak of NPX suggests that the electrode process was typically diffusion-controlled. In addition, the effect of pH reflects the participation of protons in the oxidation process of NPX. The peak current is linearly proportional to the concentration of NPX ranging from 1.0 to 100.0 µM, with the correlation coefficient (R 2), sensitivity, limit of detection (3σ), and limit of quantification (10σ) being 0.9995, 0.419 µA·µM−1·cm−2, 0.672, and 2.241 µM, respectively. Using chronoamperometry, the diffusion coefficient of NPX at the poly(r-o-NBA)/GQDs/CPE was estimated to be 5.36 × 10−6 cm2·s−1. The proposed electrode has good reproducibility, stability, and high selectivity for NPX oxidation. The obtained recovery range (96.7–102.0%) means that the proposed sensor performed satisfactorily when applied for the detection of NPX in its pharmaceutical formulations.

1 Introduction

Naproxen (NPX), also known as 2-(6-methoxynaphthalen-2-yl)propanoic acid, is a non-steroidal anti-inflammatory drug. The importance of NPX as a non-steroidal anti-inflammatory comes from its analgesic and antipyretic properties [1]. NPX is commonly used in the treatment of severe pain, fever, inflammation caused by rheumatoid arthritis, ankylosing spondylitis, primary dysmenorrhea, and degenerative joint disease [2,3]. NPX acts in the inhibition of the cyclooxygenase enzyme, which prevents the biosynthesis of certain prostaglandins [4]. Therefore, accurate quantification of NPX is essential and has attracted much attention.

In recent years, the most frequently used conventional methods of NPX detection have been spectrophotometry [5,6], spectrofluorimetry [7,8], capillary electrophoresis [9], and chromatography techniques [10,11]. These methods are undoubtedly precise, but they require protracted analysis, time-consuming detection processes, and professional operators [12,13]. However, it is important to remember that chromatographic methods can be used to estimate more than one analyte at a time.

The electrochemical methods, on the other hand, are among the various sensors used in recent years for NPX detection due to their high sensitivity, portability, selectivity towards electroactive species, simple preparation methods, and quick response [14,15]. Unfortunately, the electrochemical determination of NPX on a bare electrode suffers from a poor response and high overpotential. Therefore, a variety of modified electrodes can provide some useful advantages in electrochemical responses.

Graphene quantum dots (GQDs) consist of thin sheets of graphene that can provide a large surface area, high electrocatalytic activity, speedy electron transfer rate, and high sensitivity of the electrode [16,17]. The new zero-dimensional compound (GQDs) consists of single or multiple layers of graphene with a thickness and a diameter of less than 30 and 20 nm, respectively [17]. The absence of electrode fouling is an important factor for the modified electrodes [18]. Therefore, their low cytotoxicity, high surface area, and excellent solubility make GQDs excellent candidates for the modification of different substrates (such as glass, carbon, and graphite). Furthermore, a polymeric thin film can increase the modified electrodes’ stability, reproducibility, and active sites [19]. A polymer film of reduced-o-nitrobenzoic acid (r-o-NBA) can be easily formed on the surface of carbon paste electrode (CPE) through electropolymerisation. To the best of our knowledge, the application of a polymer film of r-o-NBA on GQDs/CPE to detect NPX as reported here is the first of its kind.

In this study, our interest was to further improve the properties of the CPE with poly(r-o-NBA) supported in GQDs to facilitate the detection of NPX. It was investigated whether the presence of a polymer film of r-o-NBA along with GQDs makes CPE a sensitive electrode for NPX determination in real samples.

2 Experimental section

2.1 Chemical reagents and solutions

NPX (98.0%), hydrochloric acid (HCl, 32.0%), and sulphuric acid (H2SO4, 98.0%) were purchased from Sigma-Aldrich. Citric acid (CA), sodium hydroxide, monosodium phosphate, and disodium phosphate were procured from El-Nasr Pharmaceutical Chemicals (Egypt). In our experiments, all the chemical reagents were of analytical grade unless otherwise stated and were used directly without any further purification. All solutions were freshly prepared using ultrapure water (18.2 MΩ·cm−1) at room temperature. A buffer solution of 0.1 M NaH2PO4–Na2HPO4 (phosphate-buffered saline [PBS], pH 7.0) was used as the supporting electrolyte. The NPX stock solution was prepared daily by dissolving an exact amount of NPX in PBS, with a concentration of 1.0 mM of NaOH and HCl solutions being used to adjust for the desired pH.

2.2 Apparatus and cell

The electrochemical experiments were performed using an EG&G Princeton Applied Research Potentiostat/Galvanostat model 263A (USA) in a three-electrode microcell (model K0264), in which a highly pure platinum wire (model K0266) and Ag/AgCl saturated with KCl (model K0265) were utilised as auxiliary and reference electrodes, respectively. At the same time, a bare CPE and the poly(r-o-NBA)/GQDS/CPE were used as the working electrode. A CyberScan pH 500 Meter (EuTech, India) was employed to adjust pH values. The synthesised GQDs were characterised by a Fourier-transform infrared spectrometer (FT-IR; JASCO [FT/IR-4100]) and powder X-ray diffraction (XRD; X’Pert3 Powder; Malvern Panalytical B.V., The Netherlands). The surface morphology of the working electrodes was probed with a scanning electron microscope (SEM; QUANTA FEG250) and a transmission electron microscope (TEM; JEOL 2100 HRTEM 200V, Tokyo, Japan). Ultra-performance liquid chromatographic analysis was utilised using LC-20ADXR liquid chromatography (Shimadzu, Japan).

2.3 Synthesis and characterisation of GQDs

The GQDs were synthesised by directly pyrolysing CA, as described by Dong et al. [20]. In brief, a balloon flask containing 2.0 g of CA was slowly heated to 200°C. The colour of the molten CA changed from pale yellow to orange in 30 min, implying the formation of GQDs. The orange liquid obtained was added drop by drop into 100 mL of 10 mg·mL−1 NaOH solution with continuous stirring. Finally, the resulting solution was neutralised (pH 7.0) with 1 M HCl. The prepared GQDs were characterised by FT-IR using a PerkinElmer Spectrum-100. As shown in Figure 1a, the FT-IR of the synthesized GQDs exhibits absorption of functional groups characteristic of the hydrophilic regions. The absorption bands at near ∼3,429, 2,854, 1,629, and 1,275 cm−1 are attributed to stretching vibration of OH, CH, CO, and COC, respectively. This means the surface of GQDs contains hydroxyl, alkyl, carboxyl, and carbonyl groups [20]. In addition, the XRD analysis of GQDs is shown in Figure 1b to investigate the crystalline nature of GQDs. As can be seen, there are two diffraction peaks centred at around 9.5° and 25.1° (2θ), which correspond to amorphous carbon. The appearance peak with a high degree (as a result of the carbonisation of CA) was corresponding to an interlayer d-spacing of 0.36 nm. The UV-Vis absorption spectra for the synthesized GQDs in the wavelength range of 200–500 nm are shown in Figure 1c. The GQDs display two absorption peaks at wavelengths of 210 and 262 nm, as can be seen. It can be observed that the GQDs exhibit two absorption peaks near ∼210 and 262 nm. The first one may be attributed to the absorption of graphene derivatives. While the other one originates from a π–π* transition of aromatic C═C bonds.

Figure 1 
                  FT-IR spectrum (a), XRD pattern (b), and UV-Vis absorption (c) of GQDs.
Figure 1

FT-IR spectrum (a), XRD pattern (b), and UV-Vis absorption (c) of GQDs.

2.4 Fabrication of the bare CPE and the poly(r-o-NBA)/GQDs/CPE

Fabrication of CPE was performed by grinding graphite powder and paraffin wax, which was heated till melting, in the ratio of 70:30 (w/w) in a mortar and pestle for 25 min. After homogenisation, the paste was carefully packed into the cavity of the Teflon tube and smoothed on a soft paper. Furthermore, a copper wire inserted into the tube served to provide electrical contact. A CPE was modified with GQDs by hand mixing 60% graphite powder and 10% GQDs with 30% paraffin wax, while the polymer film of r-o-NBA could be formed by immersing the GQDs/CPE in 0.1 M PBS (pH = 7.0) containing 5.0 mM monomer of o-NBA. Then, the potential was cycled between 0.8 and –1.5 V (vs Ag/AgCl) for 8.0 cycles at a scan rate of 100 mV·s−1. The poly(r-o-NBA)/GQDs/CPE was thoroughly rinsed with ethanol and deionised water to remove unreacted monomer before drying at room temperature.

2.5 Pharmaceutical sample preparation

Five sodium NPX tablets (containing various species) were accurately weighed and ground to a fine powder to prepare a 0.1 M NPX stock solution in a 25 mL volumetric flask containing PBS (pH = 7.0). After analysis by the proposed sensor, the percent recovered was estimated, relating the concentration of the analyte found to the concentration expected.

3 Results and discussion

3.1 Electropolymerisation of o-NBA on the surface of GQDs/CPE

The electropolymerisation of o-NBA was performed on the surface of GQDs/CPE by cyclic voltammetry (CV). In an aqueous medium, o-NBA exists as o-nitrobenzoate with a pK a value of 3.41 [21]. Figure 2 depicts the CVs of 5.0 mM o-NBA in 0.1 M PBS (pH = 7.0) at a scan rate of 100 mV·s−1. As can be seen in the first scan, three reductive and one oxidative peak appeared at potentials of around –0.32, –0.58, –0.86, and 0.04, respectively. This means the reduction of nitrobenzoate to hydroxylamine benzoate is represented by P C1. At the same time, further reduction of the formed hydroxylamine benzoate to an aminobenzoate is represented by P C2 [22]. Also, the cathodic peak (P C) was paired with the anodic peak (P A) and corresponded to the redox of nitrobenzoate [23]. Furthermore, all the peak currents increased gradually with the enhancement of cyclic sweep, indicating the polymer film’s continuous growth.

Figure 2 
                  CVs of 5.0 mM o-NBA in 0.1 M PBS (pH = 7.0) at the GQDs/CPE with a scan rate of 100 mV·s−1.
Figure 2

CVs of 5.0 mM o-NBA in 0.1 M PBS (pH = 7.0) at the GQDs/CPE with a scan rate of 100 mV·s−1.

3.2 Characterisation of the bare CPE and the poly(r-o-NBA)/GQDs/CPE

SEM is one of the important techniques used to study the surface morphology of thin films. Therefore, Figure 3 shows the SEM images of a bare CPE (Figure 3a) and the poly(r-o-NBA)/GQDs/CPE (Figure 3b). It is clear from Figure 3a that the surface of the bare CPE reveals a randomly arranged flakes-like structure with ridges and grooves. Moreover, a very thin film of paraffin wax can be seen covering the graphite particles. However, the SEM image of the poly(r-o-NBA)/GQDs/CPE (Figure 3b) revealed a completely different structure morphology, which is disclosed to be uniform. A homogeneous dispersion of the composites with scaly shape on the surface of the CPE can be observed. In addition, the TEM image was used to confirm the decoration of the poly(r-o-NBA) at the prepared GQDs/CPE (Figure 3c). As is observed, GQDs exhibit spherical shape, with a diameter ranging from 4 to 15 nm and are covered with a polymer film of r-o-NBA [24,25].

Figure 3 
                  (a) SEM image of the bare CPE, (b) SEM image of the poly(r-o-NBA)/GQDs/CPE, and (c) TEM image of the poly(r-o-NBA)/GQDs/CPE.
Figure 3

(a) SEM image of the bare CPE, (b) SEM image of the poly(r-o-NBA)/GQDs/CPE, and (c) TEM image of the poly(r-o-NBA)/GQDs/CPE.

3.3 Electrochemical response of NPX at various electrodes

The electrochemical response of NPX at the bare CPE, the GQDs/CPE, and the poly(r-o-NBA)/GQDs/CPE was investigated by CV in a potential window ranging from –0.5 to 1.5 V (vs Ag/AgCl). Figure 4 depicts CVs of 0.5 mM NPX in 0.1 M PBS with pH 7.0 at a scan rate of 100 mV·s−1 for the bare CPE (Figure 4a), the GQDs/CPE (Figure 4b), and the poly(r-o-NBA)/GQDs/CPE (Figure 4c). As can be observed, there is no response for the bare CPE towards NPX. However, the presence of GQDs caused a well-defined oxidative peak that appeared at 0.88 V, with no apparent peak found in the reverse scan. This potential value for NPX oxidation may be as a result of the formation of a cation that is stabilised through the resonance structures of the methoxy-naphthyl ring. This behaviour indicates the totally irreversible electrode process for the oxidation reaction of NPX. Under identical conditions, the peak current signal of NPX was significantly enhanced on the poly(r-o-NBA)/GQDs/CPE. This might be attributed to the more active sites caused by the composite. In addition, the polymer film improved the conductivity and facilitated electron transfer between NPX and the electrode surface.

Figure 4 
                  CVs of 0.5 mM NPX in 0.1 M PBS (pH = 7.0) using a bare CPE (a), the GQDs/CPE (b), and the poly(r-o-NBA)/GQDs/CPE (c) at a scan rate of 100 mV·s−1.
Figure 4

CVs of 0.5 mM NPX in 0.1 M PBS (pH = 7.0) using a bare CPE (a), the GQDs/CPE (b), and the poly(r-o-NBA)/GQDs/CPE (c) at a scan rate of 100 mV·s−1.

3.4 Effect of scan rate

The effect of scan rate helps understand the nature of the electrocatalytic process. Figure 5 depicts the voltammetric response of 0.5 mM NPX in 0.1 M PBS (pH 7.0) using the poly(r-o-NBA)/GQDs/CPE at various scan rates. It is revealed that the anodic peak current of NPX goes on increasing over the scan rate of 20–200 mV·s−1. In addition, the oxidation peak potential shifted to more positive values as the scan rate was increased, which verified the irreversible oxidation of NPX. The inset of Figure 5 displays the linear dependence of the anodic peak current on the square root of scan rate. The regression line equation can be represented as:

(1) I p ( µA ) = 12.991 ν 1 / 2 ( mV s 1 ) 1 / 2 + 20.62

with R 2 = 0.9995, suggesting that the electrode process for NPX was typically diffusion-controlled [26,27].

Figure 5 
                  CVs of 0.5 mM NPX in 0.1 M PBS (pH = 7.0) on poly(r-o-NBA)/GQDs/CPE at different scan rates from a to h (20, 30, 50, 60, 80, 100, 150, and 200 mV·s−1). Inset: the plot of I
                     P versus ν1/2.
Figure 5

CVs of 0.5 mM NPX in 0.1 M PBS (pH = 7.0) on poly(r-o-NBA)/GQDs/CPE at different scan rates from a to h (20, 30, 50, 60, 80, 100, 150, and 200 mV·s−1). Inset: the plot of I P versus ν1/2.

3.5 Effect of pH

It is known that the pH of the buffer solution has an important role in the analyte’s current response. Therefore, the effect of pH on the peak current and peak potential of NPX was investigated using 0.1 M PBS in the pH range (3.0–9.0), as shown in Figure 6a. Results revealed that the peak potential of NPX decreased linearly along with an increase in the pH value. The regression line equation E p (V) = 0.189–0.013 pH with R 2 = 0.9982 (far from the theoretical slope, 59 mV per pH unit) indicates that proton was not involved in NPX oxidation [28,29]. In addition, with the increase in pH values, the anodic peak current of NPX goes on increasing up to 7.0, then decreases. Thus, pH of 7.0 was used as an optimal value for the electrochemical oxidation of NPX at the poly(r-o-NBA)/GQDs/CPE. Tafel curve (Figure 6b) was employed to determine the number of electrons involved in the electrochemical oxidation of NPX. Thus, the Tafel slope of 0.323 V was obtained, which means that the electrochemical oxidation process of NPX involves a one-electron transfer. According to the aforementioned results, a reaction oxidation mechanism of NPX was proposed as shown in Scheme 1, which is in agreement with previously reported data [28,29].

Figure 6 
                  (a) Plot of I
                     p and E
                     p versus pH = 3.0–9.0 for 0.5 mM NPX in 0.1 M PBS at the poly(r-o-NBA)/GQDs/CPE at a scan rate of 100 mV·s−1; (b) Tafel curve.
Figure 6

(a) Plot of I p and E p versus pH = 3.0–9.0 for 0.5 mM NPX in 0.1 M PBS at the poly(r-o-NBA)/GQDs/CPE at a scan rate of 100 mV·s−1; (b) Tafel curve.

Scheme 1 
                  Proposed mechanism of the electrochemical oxidation of NPX.
Scheme 1

Proposed mechanism of the electrochemical oxidation of NPX.

3.6 Effect of NPX concentration

Differential pulse voltammetry (DPV) is one of the voltammetric techniques with high sensitivity and better resolution [30,31]. Accordingly, DPV was used to measure NPX at different concentrations. In this method, the poly(r-o-NBA)/GQDs/CPE was immersed in 0.1 M PBS (pH = 7.0) containing various concentrations of NPX as shown in Figure 7. As a result, the rise in NPX concentration from 1.0 to 100 µM led to an obvious increase in the anodic peak current. Furthermore, the inset of Figure 7 depicts the calibration curve obtained by plotting the anodic peak current versus NPX concentration. A good linear equation was calibrated as:

(2) I p ( µA ) = 0.419 C ( µM ) + 0.325

with the correlation coefficient (R 2), sensitivity, limit of detection (LOD; 3σ), and limit of quantification (LOQ; 10σ) being 0.9995, 0.419 µA·µM−1·cm−2, 0.672, and 2.241 µM, respectively. As shown in Table 1, the poly(r-o-NBA)/GQDs/CPE has a lower detection limit for NPX compared to the other reported electrodes.

Figure 7 
                  DPVs at the poly(r-o-NBA)/GQDs/CPE for different concentrations of NPX from a to n (1.0, 5.0, 17.5, 22.5, 30.0, 37.5, 45.0, 50.0, 60.0, 67.5, 75.0, 85.0, 92.5, and 100 µM in 0.1 M PBS [pH = 7.0]). Pulse amplitude, 25 mV; pulse width, 50 ms; and scan rate, 20 mV·s−1. Inset: the plot of the peak current versus NPX concentration.
Figure 7

DPVs at the poly(r-o-NBA)/GQDs/CPE for different concentrations of NPX from a to n (1.0, 5.0, 17.5, 22.5, 30.0, 37.5, 45.0, 50.0, 60.0, 67.5, 75.0, 85.0, 92.5, and 100 µM in 0.1 M PBS [pH = 7.0]). Pulse amplitude, 25 mV; pulse width, 50 ms; and scan rate, 20 mV·s−1. Inset: the plot of the peak current versus NPX concentration.

Table 1

Analytical parameters of several reported electrodes for the determination of NPX

Method Electrode Linear range (µM) Sensitivity (µA·µM−1) LOD (µM) Reference
Square-wave voltammetry ZnO/multi-walled carbon nanotubes (MWCNT)/CPE 1.0–200 0.256 0.23 [26]
DPV Graphite electrode 1.96–18.18 4.19 0.68 [27]
Linear sweep voltammetry Poly(l-serine)-glassy carbon electrode (GCE) 4.3–65.0 0.28 0.69 [28]
Amperometry (AMP) Shorter diameter-MWCNT/GCE 10.0–1,000 0.050 1.90 [29]
DPVA Grapene oxide/GCE 10.0–1,000 0.60 1.94 [30]
AMP MWCNT–GCE 10.0–100 0.394 0.60 [31]
DPV MWCNTs-Gr-Il/GCE 1.0–100 0.125 [32]
DPV Pt electrode 4.0–100 0.114 1.0 [33]
DPV Poly(r-o-NBA)/GQDs/CPE 1.0–100 0.419 0.672 This work

3.7 Chronoamperometry studies

Chronoamperometry, which measures current versus time in response to a potential pulse, was employed to study the electrochemical oxidation of NPX at the poly(r-o-NBA)/GQDs/CPE. Figure 8 displays the chronoamperograms of 0.1 M of PBS (pH = 7.0) containing different concentrations of NPX (0.5−3.0 mM). Under diffusion control, a linear dependence of the current on t −1/2 was obtained for various concentrations of NPX (Figure 8a). The resulting slopes of these relationships were then plotted versus the NPX concentration, as shown in Figure 8b. Using the slopes and the Cottrell equation [34]:

(3) I = n F A D 1 / 2 C b π 1 / 2 t 1 / 2

where the bulk concentration and the diffusion coefficient are represented by C b (mol·cm−3) and D (cm2·s−1), respectively, and the other symbols have their usual meaning. The D value of NPX at the poly(r-o-NBA)/GQDs/CPE was estimated to be 5.36 × 10−6 cm2·s−1.

Figure 8 
                  Chronoamperograms obtained at the poly(r-o-NBA)/GQDs/CPE in 0.1 M of PBS (pH = 7.0) for different concentrations of NPX from a to c (0.5, 2.0, and 3.0 mM). Insets: (a) plot of I versus t
                     −1/2 and (b) plot of the slope of straight lines versus NPX concentration.
Figure 8

Chronoamperograms obtained at the poly(r-o-NBA)/GQDs/CPE in 0.1 M of PBS (pH = 7.0) for different concentrations of NPX from a to c (0.5, 2.0, and 3.0 mM). Insets: (a) plot of I versus t −1/2 and (b) plot of the slope of straight lines versus NPX concentration.

3.8 Reproducibility and stability

The reproducibility of the poly(r-o-NBA)/GQDs/CPE towards the NPX detection was investigated. The reproducibility of the fabricated electrode was evaluated daily by successive measurements of seven solutions of 0.1 M PBS (pH = 7.0) containing 0.5 mM of NPX as shown in Figure 9, leading to a relative standard deviation of 2.73%. Moreover, the long-term stability of the modified electrode was evaluated. Over 5 weeks, the anodic peak signal decreased by 2.3% when the poly(r-o-NBA)/GQDs/CPE was kept in 0.1 M PBS at room temperature. The absence of any major change in the current response of NPX indicates that the proposed electrode has good reproducibility and high stability.

Figure 9 
                  Repetitive measurements of the peak currents for seven solutions of 0.1 M PBS (pH = 7.0) containing 0.5 mM of NPX at the poly(r-o-NBA)/GQDs/CPE.
Figure 9

Repetitive measurements of the peak currents for seven solutions of 0.1 M PBS (pH = 7.0) containing 0.5 mM of NPX at the poly(r-o-NBA)/GQDs/CPE.

3.9 Interference effect

To appraise the selectivity of the poly(r-o-NBA)/GQDs/CPE, the influence of several interferences on the detection of NPX was investigated. Various interference species were individually added to 0.1 M PBS (pH = 7.0) containing 0.5 mM of NPX. The results indicated that the common ions such as Na+, K+, Ca2+, Mg2+, NO3 , SO4 2−, and Cl did not show interference with NPX detection. CA, tartaric acid, glucose, lactose, alanine, and cysteine have no obvious interference in NPX determination. Not only this, but also no notable response to ibuprofen, diclofenac, and meloxicam was observed, as summarised in Table 2.

Table 2

Interference of some foreign substances for 0.5 mM of NPX

Interferents Folds
Na+, K+, Ca2+, Mg2+, NO3 , SO4 , and Cl 1,000
CA and tartaric acid 500
Glucose and lactose 500
Alanine and cysteine 100
Ibuprofen, diclofenac, and meloxicam 50

3.10 Sample analysis

According to the obtained results, the poly(r-o-NBA)/GQDs/CPE could be applied to the quantitative analysis of pharmaceutical forms (tablets, 250 and 500 mg) containing NPX. The standard addition method was used to obtain the various concentrations of NPX, and the results are given in Table 3. Satisfactory recoveries ranging from 96.7% to 102.0% were obtained, which indicates the accuracy of the proposed method. Moreover, two pharmaceutical samples of NPX were analysed using the proposed DPV and high-performance liquid chromatography (HPLC) methods, and the results are summarised in Table 4. The data confirmed that the proposed method is in strong agreement with the HPLC method. Therefore, the poly(r-o-NBA)/GQDs/CPE based on the proposed method can be successfully applied for NPX detection in pharmaceutical tablets.

Table 3

Detection of NPX in pharmaceutical tablets by standard addition method

NPX added (µM) NPX found (µM) Recovery (%) RSDa (%)
1.04 1.2
5.0 6.09 100.8 1.5
10.0 11.26 102.0 0.9
15.0 15.87 98.9 1.3
20.0 20.35 96.7 0.8

aRelative standard deviation determined for n = 3.

Table 4

Detection of NPX in pharmaceutical tablets by the proposed DPV and HPLC methods

Sample Label value (mg) DPV method (mg) HPLC method (mg)
1 250 256 252
2 500 493 504

4 Conclusion

A simple and cost-effective electrochemical modification was employed to further improve the properties of the CPE. Herein, a CPE was modified with synthesised GQDs and a polymer film of r-o-NBA, and its surface was characterised with SEM and TEM analyses. The fabricated poly(r-o-NBA)/GQDs/CPE showed favourable electrocatalytic behaviour towards the NPX oxidation. The peak current of NPX increased with an increase in the scan rate in the range of 20–200 mV·s−1. Also, a linear dependence of the anodic peak current versus NPX concentration was obtained from 1.0 to 100 µM using the DPV technique. The values of correlation coefficient (R 2), sensitivity, LOD (3σ), and LOQ (10σ) were calculated to be 0.9995, 0.419 µA·µM−1·cm−2, 0.672, and 2.241 µM, respectively. According to the chronoamperometry study, the diffusion coefficient of NPX at the poly(r-o-NBA)/GQDs/CPE was calculated to be 5.36 × 10−6 cm2·s−1. Good reproducibility, stability, and high selectivity were achieved for the fabricated electrode in NPX detection in the real sample.

Acknowledgments

The authors extend their appreciation to the Deanship of Scientific Research at Imam Mohammad Ibn Saud Islamic University for funding this work through Research Group No. RG-21-09-72.

  1. Author contributions: Mohamed Abd-Elsabour: experimental work, writing – original draft, writing – review and editing; Mortaga M. Abou-Krisha: writing – review and editing, formal analysis; Abdulrahman G Alhamzani: visualisation, writing; Tarek A. Yousef: writing – review and editing, formal analysis, project administration.

  2. Conflict of interest: The authors state no conflict of interest.

  3. Data availability statement: All data are included in the manuscript.

References

[1] Norouzi P, Dousty F, Ganjali MR, Daneshgar P. Dysprosium nanowire modified carbon paste electrode for the simultaneous determination of naproxen and paracetamol: application in pharmaceutical formulation and biological fluid. Int J Electrochem Sci. 2009;4:1373–86.10.1016/S1452-3981(23)15230-8Search in Google Scholar

[2] Todd PA, Clissold SP. Naproxen – a reappraisal of its pharmacology, and therapeutic use in rheumatic diseases and pain states. Drugs. 1990;40:91–137.10.2165/00003495-199040010-00006Search in Google Scholar

[3] Al-Qahtani AD, Hameed A, Aljuhani E, Shah R, Alharbi A, Asghar BH, et al. Iron oxide nanopowder based electrochemical sensor for sensitive voltammetric quantification of midodrine. Arab J Chem. 2021;14:103446.10.1016/j.arabjc.2021.103446Search in Google Scholar

[4] Li JJ, Norton MB, Reinhard EJ, Anderson GD, Gregory SA, Isakson PC, et al. Novel terphenyls as selective cyclooxygenase-2 inhibitors and orally active anti-inflammatory agents. J Med Chem. 1996;39:1846–56.10.1021/jm950878eSearch in Google Scholar

[5] Abdel-Hay MH, Korany MA, Bedair MM, Gazy AA. Colorimetric determination of seven nonsteroidal anti-inflammatory drugs using 2-Nitrophenylhydrazine hydrochloride. Anal Lett. 1990;23:281–94.10.1080/00032719008052444Search in Google Scholar

[6] Panderi I, Parissipoulou M. Second-derivative spectrophotometric determination of naproxen in the presence of its metabolite in human plasma. Analyst. 1994;119:697–701.10.1039/an9941900697Search in Google Scholar

[7] Damiani P, Bearzotti M, Cabezon MA. Spectrofluorometric determination of naproxen in tablets. J Pharm Biomed Anal. 2002;29:229–38.10.1016/S0731-7085(02)00063-8Search in Google Scholar

[8] Patricia CD, Mariano DB, Alejandro CO. Direct and simultaneous spectrofluorometric determination of naproxen and salicylate in human serum assisted by chemometric analysis. Anal Chim Acta. 2002;471:87–96.10.1016/S0003-2670(02)00925-XSearch in Google Scholar

[9] Sádecká J, Čakrt M, Hercegová A, Polonský J, Skacáni I. Determination of ibuprofen and naproxen in tablets. J Pharm Biomed Anal. 2001;25:881–91.10.1016/S0731-7085(01)00374-0Search in Google Scholar

[10] Elsinghorst PW, Kinzig M, Rodamer M, Holzgrabe U, Sorgel F. An LC–MS/MS procedure for the quantification of naproxen in human plasma: Development, validation, comparison with other methods, and application to a pharmacokinetic study. J Chromatogr B. 2011;879:1686–96.10.1016/j.jchromb.2011.04.012Search in Google Scholar PubMed

[11] Ekpe A, Tong J, Rodriguez L. High-performance liquid chromatographic method development and validation for the simultaneous quantitation of naproxen sodium and pseudoephedrine hydrochloride impurities. J Chromatogr Sci. 2001;39:81–6.10.1093/chromsci/39.3.81Search in Google Scholar PubMed

[12] Al-nami SY, Azher OA, Aljuhani E, Shah R, Al-Qahtani SD, Khalifa ME, et al. Voltammetric behavior of acidic catecholamine metabolites in presence of cationic surfactants. J Electrochem Soc. 2021;168:106507.10.1149/1945-7111/ac2dceSearch in Google Scholar

[13] Eldin GMG, Khalifa ME, Munshi AM, Aldawsari AM, El-Metwaly NM. Determining nadifloxacin in pharmaceutical formulations using novel differential pulse voltammetric approach. Microchem J. 2021;163:105942.10.1016/j.microc.2021.105942Search in Google Scholar

[14] Hovancová J, Šišoláková I, Oriňaková R, Oriňak A. Nanomaterial-based electrochemical sensors for detection of glucose and insulin. J Solid State Electrochem. 2017;21:2147–66.10.1007/s10008-017-3544-0Search in Google Scholar

[15] Yang S, Yang R, Li G, Qu L, Li J, Yu L. Nafion/multi-wall carbon nanotubes composite film coated glassy carbon electrode for sensitive determination of caffeine. J Electroanal Chem. 2010;639:77–82.10.1016/j.jelechem.2009.11.025Search in Google Scholar

[16] Hasanzadeh M, Hashemzadeh N, Shadjou N, Eivazi-Ziaei J, Khoubnasabjafari M, Jouyban A. Sensing of doxorubicin hydrochloride using graphene quantum dot modified glassy carbon electrode. J Mol Liq. 2016;221:354–7.10.1016/j.molliq.2016.05.082Search in Google Scholar

[17] Ghiasi T, Ahmadi S, Ahmadi E, Bavil Olyai MRT, Khodadadi Z. Novel electrochemical sensor based on modified glassy carbon electrode with graphene quantum dots, chitosan and nickel molybdate nanocomposites for diazinon and optimal design by the Taguchi method. Microchem J. 2021;160:105628–39.10.1016/j.microc.2020.105628Search in Google Scholar

[18] Stefano JS, Lima AP, Montes RHO, Richter EM, Muñoz RAA. Fast determination of naproxen in pharmaceutical formulations by batch injection analysis with pulsed amperometric detection. J Braz Chem Soc. 2012;23(10):1834–8.10.1590/S0103-50532012005000055Search in Google Scholar

[19] Rageh H, Abdel-sabour M. Pharmaceutical electrochemistry: The electrochemical behaviour of paracetamol at ZnO nanoparticales/1, 2-napthaquinone-4-sulphonic acid glassy carbon modified electrode. Anal Bioanal Electrochem. 2017;9:351–64.Search in Google Scholar

[20] Dong Y, Shao J, Chen C, Li H, Wang R, Chi Y, et al. Blue luminescent graphene quantum dots and graphene oxide prepared by tuning the carbonization degree of citric acid. Carbon. 2012;50:4738–43.10.1016/j.carbon.2012.06.002Search in Google Scholar

[21] Perrin DD. Dissociation Constants of Organic Bases in Aqueous Solution. London: Butterworths; 1965.Search in Google Scholar

[22] Shi C, Zbang W, Birke RL, Cosser DK, Lombardi JR. Time-resolved SERS, cyclic voltammetry, and digital simulation of the electroreduction of p-nitrobenzoic acid. J Phys Chem. 1991;95:6276–85.10.1021/j100169a040Search in Google Scholar

[23] Kokkinidis G, Kelaidopoulou A. Electrochemical behaviour of nitroindoles: oxidative electropolymerization and reduction of the nitro group of polymerized and non-polymerized 4-nitro and 5-nitroindole. J Electroanal Chem. 1996;414:197–208.10.1016/S0022-0728(96)04682-7Search in Google Scholar

[24] Xu S, Zhu Q, Xu S, Yuan M, Lin X, Lin W, et al. The phase behavior of n-ethylpyridinium tetrafluoroborate and sodium-based salts ATPS and its application in 2-chlorophenol extraction. Chin J Chem Eng. 2021;33:76–82.10.1016/j.cjche.2020.07.024Search in Google Scholar

[25] Guo J, Xu S, Qin Y, Li Y, Lin X, He C, et al. The temperature influence on the phase behavior of ionic liquid based aqueous two-phase systems and its extraction efficiency of 2-chlorophenol. Fluid Ph Equilibria. 2020;506:112394.10.1016/j.fluid.2019.112394Search in Google Scholar

[26] Tashkhourian J, Hemmateenejad B, Beigizadeh H, Hosseini-Sarvari M, Razmi Z. ZnO nanoparticles and multiwalled carbon nanotubes modified carbon paste electrode for determination of naproxen using electrochemical techniques. J Electroanal Chem. 2014;714–715:103–8.10.1016/j.jelechem.2013.12.026Search in Google Scholar

[27] Aguilar-Lira GY, lvarez Romero GA, Rojas-Hernμndez A, Pmez-Hernμndez ME, Rodríguez-vila JA, Romero-Romo MA. Voltammetric analysis of naproxen in graphite electrodes and its determination in pharmaceutical samples. Electroanalysis. 2014;26:1573–81.10.1002/elan.201400119Search in Google Scholar

[28] Hung C-M, Huang CP, Chen S-K, Chen C-W, Dong C-D. Electrochemical analysis of naproxen in water using poly(l-serine)-modified glassy carbon electrode. Chemosphere. 2020;254:126686.10.1016/j.chemosphere.2020.126686Search in Google Scholar

[29] Montes RHO, Lima AP, Cunha RR, Guedes TJ, dos Santos WTP, Nossol E, et al. Size effects of multi-walled carbon nanotubes on the electrochemical oxidation of propionic acid derivative drugs: Ibuprofen and naproxen. J Electroanal Chem. 2016;775:342–9.10.1016/j.jelechem.2016.06.026Search in Google Scholar

[30] Qian L, Thiruppathi AR, Elmahdy R, Zalm J, Chen A. Graphene-oxide-based electrochemical sensors for the sensitive detection of pharmaceutical drug naproxen. Sensors. 2020;20:1252.10.3390/s20051252Search in Google Scholar

[31] Montes RHO, Stefano JS, Richter EM, Munoz RAA. Exploring multiwalled carbon nanotubes for naproxen detection. Electroanalysis. 2014;26:1449–53.10.1002/elan.201400113Search in Google Scholar

[32] Sarhangzadeh K. Application of multi wall carbon nanotube–graphene hybrid for voltammetric determination of naproxen. J Iran Chem Soc. 2015;12:2133–40.10.1007/s13738-015-0690-0Search in Google Scholar

[33] Adhoum N, Toumi M, Boujlel K. Determination of naproxen in pharmaceuticals by differential pulse voltammetry at a platinum electrode. Anal Chim Acta. 2003;495:69–75.10.1016/S0003-2670(03)00922-XSearch in Google Scholar

[34] Bard AJ, Faulkner LR. Electrochemical methods fundamentals and applications. 2nd edn. New York: Wiley; 2001.Search in Google Scholar

Received: 2021-10-11
Revised: 2022-05-24
Accepted: 2022-06-05
Published Online: 2022-08-19

© 2022 Mohamed Abd-Elsabour et al., published by De Gruyter

This work is licensed under the Creative Commons Attribution 4.0 International License.

Articles in the same Issue

  1. Research Articles
  2. Abnormal retention of s-triazine herbicides on porous graphitic carbon
  3. One-factor-at-a-time method combined with ICP-MS for determining 11 elements in soy sauce and their migration from the containing glass bottles
  4. Analysis of initiator content of prepreg by near-infrared spectroscopy
  5. Simultaneous MEKC-DAD and smart spectrophotometric assays of thiocolchicoside and etoricoxib in challenging concentration ratio mixtures
  6. Alternative analytical methods for ibrutinib quantification in pharmaceutical formulation: A statistical comparison
  7. Chemometric determination of common cold infection drugs in human urine
  8. An effective, novel, and cheap carbon paste electrode for naproxen estimation
  9. Fabrication of ultra-sensitive carbon paste electrode with nanocomposite CdS modification for electroanalysis of rafoxanide in dosage form and biological fluids
  10. Purification and characterisation of phytochemicals extracted from Rhizophora mucronata: Their efficacy against Pseudomonas aeruginosa infection in Catla catla
  11. Review Articles
  12. Recent applications of quantitative analytical FTIR spectroscopy in pharmaceutical, biomedical, and clinical fields: A brief review
  13. Review of characteristics and analytical methods for determination of indomethacin
  14. A review of the application of comprehensive two-dimensional gas chromatography MS-based techniques for the analysis of persistent organic pollutants and ultra-trace level of organic pollutants in environmental samples
  15. Enrichment and analysis of glycated proteins
  16. Round robin tests of secondary raw materials: A systematic review of performance parameters
  17. Paper-based microfluidic devices: Fabrication, detection, and significant applications in various fields
  18. Applications of headspace solid-phase microextraction in human biological matrix analysis
  19. Chemometrics and infrared spectroscopy – A winning team for the analysis of illicit drug products
  20. Canagliflozin: A review with specific focus on analytical methods in biological matrices and pharmaceuticals
  21. RNA-based isothermal amplification technology and its clinical application in pathogen infection
  22. Detection of diarrheal shellfish toxins
  23. Special Issue: Nanomaterials with mimetic enzymatic properties for label-free sensors and biosensors (Guest Editors: Gang Wei and Zhiqiang Su)
  24. Preparation of cuprous oxide-supported silver-modified reduced graphene oxide nanocomposites for non-enzymatic electrochemical sensor
  25. Recent advances in matrix metalloproteinases-responsive nanoprobes for cancer diagnosis and therapy
  26. A simple salicylaldehyde-bearing pyrazine as a turn-on fluorescent chemosensor for Al3+ and Zn2+ recognition and its applications
  27. Preparation of silver nanosheet-assembled film as a surface-enhanced Raman scattering substrate
  28. Synthesis of group I–III–VI semiconductor quantum dots and its application in food safety testing
Downloaded on 27.9.2025 from https://www.degruyterbrill.com/document/doi/10.1515/revac-2022-0041/html
Scroll to top button