Abstract
Selective and sensitive sarcosine detection is crucial due to its recent endorsement as a prostate cancer (PCa) biomarker in clinical diagnosis. The reduced graphene oxide-cobalt nickel tungsten boron oxides (CoNiWBO/rGO) nanocomposite is developed as a non-enzymatic electrochemical sensor for sarcosine detection in PCa patients’ serum. CoNiWBO/rGO is synthesized by the chemical reduction method via a one-pot reduction method followed by calcination at 500 °C under a nitrogen environment for 2 h and characterized by UV-Vis, XRD, TGA, and SEM. CoNiWBO/rGO is then deposited on a glassy carbon electrode, and sarcosine sensing parameters are optimized, including concentration and pH. This non-enzymatic sensor is employed to directly determine sarcosine in serum samples. Differential pulse voltammetry (DPV) and linear sweep voltammetry (LSV) are employed to monitor the electrochemical behavior where sarcosine binding leads to oxidation. Chronoamperometric studies show the stability of the developed sensor. The results demonstrate a wide linear range from 0.1 to 50 µM and low limits of detection, i.e., 0.04 µM and 0.07 µM using DPV and LSV respectivel. Moreover, the calculated recovery of sarcosine in human serum of prostate cancer patients is 78–96%. The developed electrochemical sensor for sarcosine detection can have potential applications in clinical diagnosis.
Similar content being viewed by others
Introduction
Prostate cancer (PCa) has emerged as the second most prevalent malignancy in the male genito-urinary system1, characterized by the uncontrolled proliferation of prostate cells2. Prostate-specific antigen (PSA) detection and quantification are essential biomarkers for early PCa screening and diagnosis3,4. However, PSA detection is hampered by its limited sensitivity and specificity, which results in false-positive outcomes and needless biopsy of individuals5,6. Therefore, it is thus essential to explore new biomarkers and develop efficient screening methods for PCa diagnosis.
Sarcosine (C3H7NO2), also called N-methyl glycine, is a natural non-essential amino acid in muscles and various body tissues7. It naturally occurs as an intermediate in the choline-to-glycine metabolic pathway and is produced from the synthesis and breakdown of glycine8. Sarcosine metabolizes into glycine through sarcosine oxidase9. Sarcosine levels vary from 1.4 to 0.6 µM in healthy blood and 2–10 µM in PCa patients10. Sarcosine elevated levels in blood serve as a potential biomarker for PCa detection. Its concentration increases in invasive prostate cancer cell lines compared to benign prostate epithelial cells. The sarcosine concentration in biological samples may surpass several micromoles depending on PCa aggressiveness11. Monitoring sarcosine levels can thus provide insights into PCa progression and severity.
The methods employed for sarcosine determination are GC-MS12, HPLC with fluorescence detection13, capillary electrophoresis14, differential mobility analysis15, chemiluminescence16, fluorimetry17, surface-enhanced Raman scattering (SERS) spectroscopy18, fluorescence spectroscopy17, enzymatic colorimetry, fluorometric methods19,20, enzyme-based21, immunoassay22, and electrochemical methods23. However, these methods have limitations such as time-taking, high cost, need of skilled operator, complex sample pretreatment, sensitivity to interferences, and reduced sensitivity. Electrochemical techniques have advantages of simplicity, swift response, sensitivity, linear range, cost-effectiveness, and miniaturization24. These sensors can identify biomarkers linked to disease diagnosis via in vitro or in vivo systems25,26.
The electrochemical sensors for sarcosine predominantly use sarcosine oxidase (SOx) due to its high specificity and catalytic activity. SOx facilitates sarcosine catalysis and transforms into glycine, formaldehyde, and hydrogen peroxide8,10. Few sarcosine biosensors are based on SarOx/CHIT/CuNPs/c-MWCNT/Au8, Sox/CHIT/GNRs/AuE27, and SOx/graphene/chitosan/AgNPs/GCE28. SOx immobilization onto the working electrode is essential in these biosensors. However, directly immobilizing enzymes on support causes denaturation and reduces enzyme activity and stability. Thus, an affordable, easily operatable, and selective non-enzymatic electrochemical sensor is needed for sarcosine detection.
This study proposes a reduced graphene oxide-cobalt nickel tungsten boron oxides (CoNiWBO/rGO) nanocomposite as a non-enzymatic electrochemical sensor for sarcosine. Cobal-based composites have been widely used in electrochemical applications, including sensors and supercapacitors29,30. The electrochemical parameters of sarcosine are evaluated using CoNiWBO/rGO through LSV, DPV, and EIS. Results indicate that the CoNiWBO/rGO nanocomposite functions as a non-enzymatic electrochemical sensor with remarkable attributes, i.e., higher conductivity, selectivity, stability, rapid analysis, and cost-efficacy in sarcosine sensing. The employment of novel sensing nanocomposite for sarcosine determination presents a leap forward in analyzing the fabricated compound´s potential.
Experimental
Chemicals and reagents
Nickel nitrate hexahydrate (Ni(NO3)2·6H2O, 99%), cobalt nitrate hexahydrate (Co(NO3)2·6H2O, 99%), sodium tungstate dihydrate (Na2WO4.2H2O, ≥ 97%), graphene oxide (GO), absolute ethanol, and sodium borohydrate (NaBH4) were purchased from Sigma Aldrich, USA. Dihydrogen potassium phosphate (KH2PO4, 98.5%) and monohydrogen potassium phosphate (K2HPO4, 98%) were used for preparing 0.1 M phosphate buffer saline (PBS). All chemicals used in the experiments were of analytical grade and employed without further purification.
Synthesis of CoNiWBO/rGO nanocomposite
CoNiWBO/rGO nanocomposite was synthesized using the chemical reduction method31. Briefly, 12 mM Co(NO3)2.6H2O, 6 mM Na2WO4.2H2O, 6 mM Ni(NO3)2.6H2O, and 40 mg GO were mixed in 60 mL absolute ethanol and stirred for 24 h. Subsequently, 20 mL NaBH4 (10 wt%) solution was added to the mixture and stirred for one hour. The black product (CoNiWBO/rGO) was collected, rinsed with deionized water and absolute ethanol, and air-dried for 24 h at 25 °C. CoNiWBO/rGO was calcinated at 500 °C under nitrogen for 2 h.
Electrode fabrication
The glassy carbon electrode (GCE) used in the study was sonicated in water and methanol for 15 min. Ten mg CoNiWBO/rGO was mixed with 5 mL deionized water to make a slurry. 10 µL of this mixture was applied to the working GCE and dried. The developed electrode was used for the experiments at the electrochemical workstation (CORRTEST-CS120 model) in the three-electrode system of Ag/AgCl as the reference electrode, glassy carbon (GC) as the working electrode, and Pt wire as the counter electrode. Electrochemical measurements were made at room temperature, including CV, DPV, LSV, EIS, and chronoamperometry (CA).
Optimizations for sarcosine detection
0.1 M PBS was used as an electrolyte solution. The sarcosine sensor CoNiWBO/rGO sensitivity was assessed by recording the oxidation current as a function of applied potential from − 0.1 to 0.94 V in 30 mL solutions of increasing sarcosine concentrations (0.1–50 µM). An electrochemical response of sarcosine was examined at pH of 7.8, 7.6, 7.4, 7.2, 7.0, and 6.8 in 0.1 M PBS.
Serum sample collection and processing
Blood samples of PCa patients were obtained from MINAR Cancer Hospital, Multan, Pakistan, with prior approval by the MINAR Ethical Committee. All procedures complied with the guidelines of the Ethical Committee of MINAR Cancer Hospital, Multan, Pakistan. The blood samples were collected from the volunteers with informed consent. Samples were centrifuged for 10 min at 4000 rpm to remove solid particulate matter and stored at − 20 °C before sarcosine detection. Potentiostat was used to determine sarcosine in serum samples of PCa patients.
Results and discussion
Characterization of CoNiWBO/rGO
X-ray diffraction (XRD), UV-Vis, thermogravimetric analysis (TGA), and scanning electron microscopy (SEM) are used to examine the chemical composition and morphology. XRD analysis of CoNiWBO/rGO is conducted on Bruker D8 Advance Powder Diffractometer in the diffraction range of 5°–80°. GO diffraction peak is observed at 10°32. The inter-atomic or d-spacing is calculated using Bragg’s law33.
The mean inter-atomic spacing of GO is 6 Å. XRD pattern of CoNiWBO/rGO does not reveal distinct peaks and thus indicates its predominantly amorphous nature (Fig. 1A)31. The absence of GO peaks confirms the complete reduction of GO to rGO by NaBH434.
The UV spectrum of CoNiWBO/rGO has an absorbance of 243 nm (Fig. 1B), which indicates the presence of metals in the prepared nanocomposite. The presence of cobalt and nickel in CoNiWBO/rGO is confirmed because of their absorbance in the UV range of 180 to 450 nm35. TGA reveals 6.5% weight loss at 100 °C corresponding to the elimination of water molecules in the nanocomposite (Fig. 1C). There is a minor weight loss of 7.2% at 153 °C. However, CoNiWBO/rGO exhibits 13.4% weight loss from 160 to 550 °C, indicating nanocomposite degradation at elevated temperatures.
Fourier transform infrared spectroscopy (FTIR) examined (Fig. 1D) and measured the composition of GO, CoNiWBO, and CoNiWBO/rGO in the mid-infrared range (4000–400 cm− 1). The peaks located at 3400 cm− 1 are attributed to the O–H bond, and the bands at 2925 cm− 1 and 1715 cm− 1 show the stretching of the C–H bond and C=O36, while the bands at the low wavelength region (400–1100 cm− 1) show stretching vibrations of the Ni–O, W–O, and Co–O bonds31. These results indicate that the absorption bands of the metal-oxygen bond are shifted and broadened and show that Ni, Co, and W elements were oxidized in the nanocomposites.
SEM analysis provides the nanocomposite morphology (Fig. 2A), where it has well-dispersed flower-like characteristics to indicate the reduction in CoNiWBO/rGO nanoflakes due to rGO. CoNiWBO/rGO displays a porous structure formed by the aggregation of nanosheets to attain significant pore volume and surface area. ImageJ software, specifically NIH version 1.52 g, examines the fiber diameters in CoNiWBO/rGO. An average particle size of 29 nm is found for the synthesized nanocomposite (Fig. 2B). The elemental composition of CoNiWBO/rGO is determined by energy dispersive X-ray (EDS) spectroscopy, and obtained results suggest that all representative elements (C, O, Co, Ni, W, and B) are present at an appropriate ratio (Fig. 2C). Atomic force microscopy (AFM) analysis of the synthesized composite suggests that particles exhibit a rougher surface (Fig. 2D).
Electrochemical characterization of CoNiWBO/rGO
Electrochemical measurements characterize the synthesized nanocomposite through CV, stability, scan rate, chronoamperometry, and EIS studies in 40 mM K4[Fe(CN)6].3H2O and 0.1 M KCl solutions. No redox behavior has been observed at bare GCE and reduced graphene oxide modified electrode (rGO)/GCE (Fig. 3A). Whereas CoNiWBO shows redox potential, but prominent anodic (oxidation peak at − 0.08 V with a current of 1.34 mA) and cathodic (reduction peak at − 0.2 V with a current of − 1.07 A) potential is given by CoNiWBO/rGO. The reduction peak of CoNiWBO/rGO in electrolyte solution depicts the cathodic process, and the oxidation peak depicts an anodic process. The current densities of CoNiWBO/rGO are greater when compared with other electrodes, which indicates that the synthesized nanocomposite has an active site for the redox process and shows rapid electron transfer kinetics. CV curves exhibit distinct redox peaks. B2O3 does not show electrochemical activity in the composite. However, it has a buffering role in accommodating volume changes in the charge and discharge process. The following equations can represent the faradaic reactions of surface oxidation species of the nanocomposite37,38,39,40:
CV analyses of CoNiWBO/rGO and bare GCE illustrate different electrochemical behaviors. Bare GCE acts as a baseline in this process. CoNiWBO/rGO depicts oxidation and reduction peaks with increased current densities, suggesting better electron transfer kinetics and increased reactivity. Comparison with the literature validates that the electrochemical behavior of CoNiWBO/rGO is superior31,41.
CoNiWBO/rGO performance is assessed through 100 electrochemical cyclic stability tests. Each cycle involves repetitive redox reactions at a scan rate of 30 mVs− 1 and potential window − 0.39 V to 0.11 V. CV analysis displays consistent redox peaks to affirm the nanocomposite stability (Fig. 3B). The material can thus sustain repeated redox reactions without notable decline in electrochemical performance.
Varying scan rates affect the redox behavior of CoNiWBO/rGO. CV responses are recorded with the scan rates from 30 to 100 mVs− 1 in a potential window of − 0.39 V to 0.11 V. The redox process involving CoNiWBO/rGO exhibits rapid electron transfer with increased oxidation and reduction currents at higher scan rates (Fig. 3C). The peak current reaches a maximum at a scan rate of 100 mVs− 1. A calibration plot shows a linear correlation between scan rate and peak current (Fig. 3D). CoNiWBO/rGO current increases linearly with the scan rate, and the oxidation peaks shift to higher potential while reduction peaks to lower potential, resulting in significant potential separation31.
Chronoamperometry (CA)
CA experiments are conducted on CoNiWBO/rGO to explore the nature of steady-state current. CA curve in Fig. 4A provides information about the electrode’s time-dependent behavior. Despite a minor initial decrease, the curve highlights the stability of CoNiWBO/rGO-modified GCE. The voltammogram indicates a stable current for 12 h to reflect the sustained electrochemical activity of the electrode.
Electrochemical surface area (ECSA) of CoNiWBO/rGO-GCE
ECSA measurements are carried out at scan rates of 30 to 100 mVs− 142. A graph indicates the linear relationship between current density and scan rates to determine ECSA (Fig. 4B). The line graph of ECSA has a slope of 0.0002 and R2 of 0.99 (Fig. 4C). The calculated ECSA for CoNiWBO/rGO-GCE is 0.1 cm2, which is higher than bare GCE, i.e., 0.073 cm2. This suggests a larger surface area for redox reactions and the overall electrochemical activity provided by CoNiWBO/rGO. Previous study indicates that the electrochemical performance is higher for CoNiWBO/rGO electrodes, as inferred from the CV curve and corresponding area43.
Roughness factor (Rf)
Rf is defined as the ratio of the surface area of modified GCE (A2) to that of unmodified GCE (A1). It provides information on the changes in surface area resulting from electrode modification with CoNiWBO/rGO regarding oxidation phases and overall surface area dimensions.
ECSAs of modified GCE (A2) and bare GCE (A1) are compared to determine Rf for CoNiWBO/rGO-GCE-modified electrodes. ECSA is 0.073 cm2 for bare GCE and 0.1 cm2 for CoNiWBO/rGO-GCE. Rf is calculated as 1.43 for CoNiWBO/rGO-GCE.
Electrochemical detection of sarcosine
The sarcosine molecule undergoes oxidation in the electrochemical sensing process at appropriate electrode potentials. Figure 5 illustrates the sarcosine response in electrochemical oxidation under alkaline conditions23.
Incorporating rGO in CoNiWBO enhances the electrochemical detection of sarcosine via linear sweep voltammetry (LSV) and differential pulse voltammetry (DPV). The nanocomposite provides good electrical conductivity and higher working electrode surface area when employed on the electrode. There is swift electron transfer between the target analyte and electrode for improved response and lower detection limits. The sensor sensitivity for sarcosine is assessed by measuring current versus applied potential in various sarcosine concentrations.
Linear sweep voltammetry (LSV)
LSV measurements are conducted with and without sarcosine in 0.1 M PBS. No peak exists in the potential from − 0.1 to 0.94 V for a blank solution without sarcosine. CoNiWBO/rGO serves as the sensor. An oxidation peak with a current of 13.77 µA at 0.6 V is observed by introducing 0.1 µM sarcosine to the blank solution (Fig. 6A). Peak current is linearly related to sarcosine concentration having a high correlation coefficient, i.e., R2 = 0.99 (Fig. 6B)44.
The pH’s impact on the electrochemical behavior of CoNiWBO/rGO is tested through LSV in 0.1 M PBS. The oxidation peak current shifts from pH 6.8 to 7.8 with the apex at pH 7.4 (Fig. 6C). Figure 6D shows a line graph at different pH.
Differential pulse voltammetry (DPV)
Differences in Faradaic current and rate of charging decay are measured by differential pulse voltammetry based on applying pulsed potentials within a linear potential scanning. The baseline for this study is developed by recording DPV under optimized conditions as 0.1 M PBS solution at room temperature without sarcosine. No oxidation peak is observed using CoNiWBO/rGO as an electrochemical sensor in the potential from − 0.1 V to 0.94 V at a scan rate of 30 mVs− 1. Distinct oxidation peaks are observed at 0.6 V upon adding 0.1 µM sarcosine (Fig. 7A). DPV curves are consistent with the sarcosine concentrations of 0.1 µM to 50 µM. The oxidation peak current and sarcosine concentration have a linear trend (Fig. 7B)45.
The pH’s impact on the electrochemical behavior of 50 µM sarcosine is determined (Fig. 7C). The lowest oxidation peak with a current of 5.5 µA is observed at pH 6.8, while the highest is at pH 7.4. Figure 7D shows a line graph at different pH.
Differential pulse voltammetry (DPV) response of CoNiWBO/rGO for sarcosine sensing in 0.1 M PBS with different parameters, (A) sarcosine concentrations of 0.1, 10, 20, 30, 40, 50 µM, (B) line graph between sarcosine concentration and current, (C) pH of 7.8, 7.6, 7.4, 7.2, 7.0, 6.8, and (D) line graph between pH and current.
Electrochemical impedance studies on CoNiWBO/rGO
Bare GCE has the highest semicircle diameter and thus elevated ohmic resistance (Fig. 8A). This implies limited electrochemical conductivity, causing impedance to charge flow. CoNiWBO/rGO exhibits enhanced electrochemical conductivity compared to GCE. The smaller semicircle in the Nyquist plot signifies lower ohmic resistance. CoNiWBO/rGO enhances charge transfer kinetics at the electrode and electrolyte interface and promotes electron transport in the electrochemical process.
The impact of varying sarcosine concentrations (0.1 µM to 50 µM) on the impedance of the CoNiWBO/rGO electrode system is explored (Fig. 8B). The lowest impedance indicates optimal charge transfer and minimal ohmic resistance occurring at sarcosine concentration of 0.1 µM and the highest at 50 µM. The pH of PBS in sarcosine sensing affects CoNiWBO/rGO electrode system impedance (Fig. 8C). Electrode showing the smallest semicircle indicates minimum charge transfer resistance for sarcosine sensing at pH 7.4.
Heterogenous rate constant (Ko)
Ko is a redox reaction’s equilibrium constant which influences reaction kinetics. A higher Ko indicates rapid and sustainable equilibrium attainment compared to that with lower Ko. Ko is calculated from the following equation46.
where Ko is the heterogeneous rate constant, F the Faraday constant (96485 Cmol− 1), T the thermodynamic temperature (298.15 K), Rct the electron transfer resistance (6842 Ω for bare GCE, 4087 Ω for CoNiWBO/rGO-GCE, and 3245 Ω for sarcosine), R the general gas constant (8.314 Jmol− 1K− 1), A the electrode surface area, and C the KCl concentration (0.1 M). Ko is 53 × 10− 6 cms− 1, 65 × 10− 6 cms− 1, and 82 × 0− 6 cms− 1 for bare GCE, CoNiWBO/rGO-GCE, and sarcosine detection by CoNiWBO/rGO.
Exponential factor
The exponential factor is calculated by the following equation:
where R is the charge transfer resistance, pi the circle’s perimeter ~ 3.14, f the frequency i.e., 0.01 Hz, and τ the time constant. τ = R × C where C is the capacitance. The exponential factor for CoNiWBO/rGO is 2.4 × 106, for bare GCE is 1.2 × 108, and for sarcosine detection is 1.3 × 108 at optimum pH and concentration (Rct for optimum pH and concentration are 5304 Ω and 7574 Ω).
Repeatability, selectivity, and stability of the sensor
The repeatability is examined by running seven measurements under the same conditions with 50 µM sarcosine in 0.1 M PBS. DPV (Fig. 9A) and LSV (Fig. 9B) experiments support the repeatability and stability of CoNiWBO/rGO as electrochemical sensor for sarcosine detection. The selectivity of CoNiWBO/rGO-GCE electrode is assessed in the presence of 50 µM interfering substances (uric acid, glucose, and creatinine) in 0.1 M PBS of pH 7.4. The electrode performance is unaffected by the presence of these substances (Fig. 9C and D).
(A) DPV and (B) LSV repeatability and reproducibility results of CoNiWBO/rGO-GCE from successive seven readings of 50 µM sarcosine detection in 0.1 M PBS. The electrochemical response of CoNiWBO/rGO-GCE in the presence of equal concentrations (50 µM) of interfering species like uric acid, glucose, and creatinine in 0.1 M PBS via (C) DPV and (D) LSV.
Limit of detection (LOD)
LOD is the minimum detectable quantity of analyte. It depends on reaction parameters, analyte concentration, and the detection technique. LOD is determined by the following Eq.
where LOD is the detection limit, ‘s’ the standard deviation, i.e., 0.004 and 0.01 for DPV and LSV, respectively, and ‘m’ the slope, i.e., 0.26 and 0.38 for DPV and LSV. The calculated LODs for sarcosine detection are 0.04 µM and 0.07 µM by DPV and LSV, respectively.
Sarcosine determination in prostate cancer patients
The objective of this study is to assess CoNiWBO/rGO for quantifying sarcosine in human serums of PCa patients and healthy persons. The serum samples are diluted fifty times using 0.1 M PBS of pH 7.4. In this study, 50 µM sarcosine is added into serum samples, and its concentration is determined by the linear relationship depicted in Figs. 6B and 7B. Results are presented in Table 2. Higher intensities are observed in PCa patients compared to control group (Fig. 10A and B). Sarcosine concentrations in PCa patients are ~ 3 times greater than healthy donors (Fig. 10). Table 1 displays recoveries to affirm the efficacy of modified sensor for sarcosine quantification in human serum.
A detailed comparison with previously published literature is given in Table 2. Comparison suggests that the designed sensors exhibits enhanced electrochemical sensing performance than several previously reported sensors.
Conclusion
An economical non-enzymatic electrochemical sensor CoNiWBO/rGO was developed to determine serum sarcosine levels, in PCa affected individuals. The working electrode modified with CoNiWBO/rGO enhanced surface area, conductivity, and accuracy for detecting micro-molar concentrations of sarcosine. CoNiWBO/rGO was prepared by co-precipitation method, and morphological and structural analysis confirms the formation of 3D flower shaped CoNiWBO/rGO. The voltametric studies, i.e., LSV and DPV demonstrates that the proposed sensor exhibits higher selectivity and sensitivity for sarcosine detection, while chronoamperometric studies shows stability of the developed sensor. The results demonstrate wide linear range from 0.1 to 50 µM and low limits of detection, i.e., 0.04 µM and 0.07 µM for DPV and LSV, respectively. The optimized sensor for sarcosine can be appropriate to screen PCa in labs via this biomarker, as it is superior to prostate specific antigen (PSA). This sensor in future can be employed as point of care devices, robotics, and automations as it depicts higher reproducibility and recovery.
Data availability
All data generated or analysed during this study are included in this published article.
References
Crocetto, F. et al. Liquid biopsy in prostate cancer management—Current challenges and future perspectives. Cancers 14(13), 3272 (2022).
Islami, F. et al. American Cancer Society’s report on the status of cancer disparities in the United States, 2021. CA Cancer J. Clin. 72(2), 112–143 (2022).
Garg, S. et al. Point-of-care prostate specific antigen testing: Examining translational progress toward clinical implementation. ACS Sens. 8(10), 3643–3658 (2023).
Boehm, B. E. et al. Biomarkers of aggressive prostate cancer at diagnosis. Int. J. Mol. Sci. 24(3), 2185 (2023).
Liao, C. et al. Nurturing the marriages of urinary liquid biopsies and nano-diagnostics for precision urinalysis of prostate cancer. Smart Med. 2(1), e20220020 (2023).
Berenguer, C. V. et al. Underlying features of prostate cancer—Statistics, risk factors, and emerging methods for its diagnosis. Curr. Oncol. 30(2), 2300–2321 (2023).
Schcolnik-Cabrera, A. & Juárez-López, D. Dual contribution of the mTOR pathway and of the metabolism of amino acids in prostate cancer. Cell. Oncol. 45(5), 831–859 (2022).
Narwal, V. et al. Fabrication of an amperometric sarcosine biosensor based on sarcosine oxidase/chitosan/CuNPs/c-MWCNT/Au electrode for detection of prostate cancer. Enzyme Microb. Technol. 113, 44–51 (2018).
Lahham, M. et al. The family of sarcosine oxidases: Same reaction, different products. Arch. Biochem. Biophys. 704, 108868 (2021).
Kumar, P. et al. Construction and application of amperometric sarcosine biosensor based on SOxNPs/AuE for determination of prostate cancer. Biosens. Bioelectron. 122, 140–146 (2018).
Krishnan, S. et al. Circulating metabolite biomarkers: A game changer in the human prostate cancer diagnosis. J. Cancer Res. Clin. Oncol. 149(3), 951–967 (2023).
Markin, P. A. et al. Plasma sarcosine measured by gas chromatography-mass spectrometry distinguishes prostatic intraepithelial neoplasia and prostate cancer from benign prostate hyperplasia. Lab. Med. 51(6), 566–573 (2020).
Chung, T.-C. et al. High-performance liquid chromatographic analysis of sarcosine as a fluorescent levofloxacin derivative. J. Chromatogr. Sci. 53(8), 1310–1315 (2015).
Ramezani, Z., Safdarian, M. & Ghadiri, A. A. Metal-coded hydrogel magnetic molecularly imprinted polymer for preconcentration and cleanup of sarcosine: Determination in urine; Coupled to on-column capillary electrophoresis. Talanta 230, 122309 (2021).
Berthias, F. et al. Identification and quantification of amino acids and related compounds based on differential mobility spectrometry. Analyst 145(14), 4889–4900 (2020).
Valenti, G. et al. An electrochemiluminescence-supramolecular approach to sarcosine detection for early diagnosis of prostate cancer. Faraday Discuss. 185, 299–309 (2015).
Wang, M. et al. Fluorescence sensing platform for sarcosine analysis based on nitrogen-doping copper nanosheets and gold nanoclusters. Anal. Chim. Acta 1223, 340188 (2022).
Haroon, M. et al. Surface-enhanced Raman scattering (SERS) spectroscopy for prostate cancer diagnosis: A review. Photodiagn. Photodyn. Ther. 37, 102690 (2022).
Masumoto, M. et al.Colorimetric paper-based sarcosine assay with improved sensitivity. Anal. Bioanal. Chem. 1–11 (2022).
Qin, W. et al. Facile synthesis and ozonation of carbon dots using mango pulps for dual-mode colorimetry and ratiometric fluorescence detection of sarcosine. Microchem. J. 195, 109468 (2023).
Hu, Q. et al. Determination of sarcosine based on magnetic cross-linked enzyme aggregates for diagnosis of prostate cancer. Biochem. Eng. J. 172, 108039 (2021).
Heger, Z. et al. Paramagnetic nanoparticles as a platform for FRET-based sarcosine picomolar detection. Sci. Rep. 5(1), 8868 (2015).
Bahrami, H., Mousavi, M. & Maghsoudi, S. Sensitive voltammetric method for rapid determination of sarcosine as a new biomarker for prostate cancer using a TiO2 nanoparticle/ionic liquid modified carbon paste electrode. Russ. J. Electrochem. 57, 149–158 (2021).
Saleh, H. M. & Hassan, A. I. Synthesis and characterization of nanomaterials for application in cost-effective electrochemical devices. Sustainability 15(14), 10891 (2023).
Bloss, R. Multi-technology sensors are being developed for medical, manufacturing, personal health and other applications not previously possible with historic single-technology sensors. Sens. Rev. 37(4), 385–389 (2017).
Rivas, G. A. et al. Carbon nanotubes-based electrochemical (bio) sensors for biomarkers. Appl. Mater.Today 9, 566–588 (2017).
Deswal, R. et al. An improved amperometric sarcosine biosensor based on graphene nanoribbon/chitosan nanocomposite for detection of prostate cancer. Sens. Int. 3, 100174 (2022).
Zhou, Y. et al. Direct electrochemistry of sarcosine oxidase on graphene, chitosan and silver nanoparticles modified glassy carbon electrode and its biosensing for hydrogen peroxide. Electrochim. Acta 71, 294–301 (2012).
Reddy, G. R. et al. Engineering cationic vacancies on sphere-like zinc cobaltite microstructures via self-assembly of silkworm-like interconnected nanoparticles for battery-type supercapacitors. CrystEngComm 25(17), 2618–2628 (2023).
Kuchi, C. et al. Self-assembled three-dimensional intertwined zinc cobaltite nanocubes for high-performance supercapacitors: A solvothermal route. Mater. Sci. Semicond. Process. 142, 106453 (2022).
Xiang, C. et al. Simple synthesis of graphene-doped flower-like cobalt–nickel–tungsten–boron oxides with self-oxidation for high-performance supercapacitors. J. Mater. Chem. A 5(20), 9907–9916 (2017).
Siburian, R. et al. New route to synthesize of graphene nano sheets (2018).
Hadi, A. A. A. & AL-Kifaie, A. M. A. Preparation and study of the structural properties of SnO2 nanoparticles. J. Surv. Fish. Sci. 10(3), 6674–6677 (2023).
Mascagni, D. B. T. et al. Layer-by-layer assembly of functionalized reduced graphene oxide for direct electrochemistry and glucose detection. Mater. Sci. Eng. C 68, 739–745 (2016).
Malik, M., Chan, K. H. & Azimi, G. Quantification of nickel, cobalt, and manganese concentration using ultraviolet-visible spectroscopy. RSC Adv. 11(45), 28014–28028 (2021).
Kanta, U. et al. Preparations, characterizations, and a comparative study on photovoltaic performance of two different types of graphene/TiO2 nanocomposites photoelectrodes. J. Nanomater. 2017 (2017).
Ma, H. et al. Nickel cobalt hydroxide@ reduced graphene oxide hybrid nanolayers for high performance asymmetric supercapacitors with remarkable cycling stability. ACS Appl. Mater. Interfaces 8(3), 1992–2000 (2016).
Zheng, Y. et al. Multi-channeled hierarchical porous carbon incorporated Co3O4 nanopillar arrays as 3D binder-free electrode for high performance supercapacitors. Nano Energy 20, 94–107 (2016).
Yang, M. et al. Binder-free hydrogenated NiO–CoO hybrid electrodes for high performance supercapacitors. RSC Adv. 5(40), 31725–31731 (2015).
Shinde, N. M. et al. Wet chemical synthesis of WO3 thin films for supercapacitor application. Korean J. Chem. Eng. 32, 974–979 (2015).
Sun, L. Interfacing graphene with nanoparticles for energy and gas storage as well as sensors.
Cowen, T. & Cheffena, M. Molecularly imprinted polymer real-time gas sensor for ambient methanol vapor analysis developed using principles of sustainable chemistry. ACS Sustain. Chem. Eng. 11(29), 10598–10604 (2023).
Sun, L. Graphene composites for energy storage and sensors (2019).
Rebelo, T. S. et al. Sarcosine oxidase composite screen-printed electrode for sarcosine determination in biological samples. Anal. Chim. Acta 850, 26–32 (2014).
Lin, S. et al. Electrochemical detection of sarcosine and supercapacitor based on a new Ni–metal organic framework electrode material. Crystals 11(9), 1036 (2021).
Krzyczmonik, P. et al. Honeycomb-structured porous poly (3, 4-ethylenedioxythiophene) composite layers on a gold electrode. Thin Solid Films 565, 54–61 (2014).
Cernei, N. et al. Spectrometric and electrochemical analysis of sarcosine as a potential prostate carcinoma marker. Int. J. Electrochem. Sci 7(5), 4286–4301 (2012).
Xue, Z. et al. An organic indicator functionalized graphene oxide nanocomposite-based colorimetric assay for the detection of sarcosine. Nanoscale 8(10), 5488–5496 (2016).
Sheydaei, O., Khajehsharifi, H. & Rajabi, H. R. Rapid and selective diagnose of Sarcosine in urine samples as prostate cancer biomarker by mesoporous imprinted polymeric nanobeads modified electrode. Sens. Actuators B Chem. 309, 127559 (2020).
Rashedi, M. & Alizadeh, T. A novel non-enzymatic sensor for prostate cancer biomarker sensing based on electrocatalytic oxidation of sarcosine at nanostructured NiMn2O4 impregnated carbon paste electrode. Anal. Chim. Acta 1186, 339121 (2021).
Udayakumar, J. & Selvaraj, S. One-step nonenzymatic electrochemical sensor for the detection of sarcosine using nanozyme glutathione copper complex. Curr. Anal. Chem. 19(8), 577–584 (2023).
Farokhi, S., Roushani, M. & Saedi, Z. Fabrication of an electrochemical aptasensor for the determination of sarcosine based on synthesized CuCo2O4 nanosheets. Anal. Methods 15(37), 4938–4945 (2023).
Feng, J. et al. Sarcosine biosensor based on Pt/g-C3N4 nanocomposites with high electrocatalytic activity. ECS J. Solid State Sci. Technol. 11(4), 047001 (2022).
Acknowledgements
Authors acknowledge the Higher Education Commission (HEC) of Pakistan for this research.
Author information
Authors and Affiliations
Contributions
M.W. performed experiments and wrote the initial draft. S.S. prepared figures and wrote the initial draft. D.H. provided resources and validated the data. B.F. supervision, conceptualization, and validation. F.H. interpreted the results and reviewed the manuscript. S.T. visualized and validated results. M.M.R. visualized and validated results. A.Y. collected and prepared samples for analysis. S.M. reviewed and wrote the final draft. M.N.H. reviewed and wrote the final draft.
Corresponding authors
Ethics declarations
Competing interests
The authors declare no competing interests.
Additional information
Publisher’s note
Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Rights and permissions
Open Access This article is licensed under a Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 International License, which permits any non-commercial use, sharing, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if you modified the licensed material. You do not have permission under this licence to share adapted material derived from this article or parts of it. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by-nc-nd/4.0/.
About this article
Cite this article
Wasim, M., Shaheen, S., Fatima, B. et al. Non-enzymatic electrochemical detection of sarcosine in serum of prostate cancer patients by CoNiWBO/rGO nanocomposite. Sci Rep 14, 24240 (2024). https://doi.org/10.1038/s41598-024-74628-6
Received:
Accepted:
Published:
DOI: https://doi.org/10.1038/s41598-024-74628-6