Tuning the direct growth of Agseeds into bimetallic Ag@Cu nanorods on surface functionalized electrochemically reduced graphene oxide: enhanced nitrite detection

S. E. Jeenaa, P. Gnanaprakasama, Arun Dakshinamurthyb and T. Selvaraju*a
aDepartment of Chemistry, Karunya University, Coimbatore 641114, India. E-mail: selvaraju@karunya.edu; Fax: +91-422-2615615; Tel: +91-422-2614149
bDepartment of Nanoscience and Nanotechnology, Karunya University, Coimbatore 64114, India

Received 31st March 2015 , Accepted 21st May 2015

First published on 21st May 2015


Abstract

We describe a facile and an efficient method for the direct growth of bimetallic Ag@Cu nanorods (Ag@Cu NRDs) on electrochemically reduced graphene oxide (ERGO) nanosheets by a seed mediated growth approach. Surface functionalization of ERGO is a key step carried out using L-tryptophan that directs the growth of one dimensional Ag@Cu NRDs on its surface. The generated bimetallic Ag@Cu NRDs on L-tryptophan functionalized ERGO (Ag@Cu NRDs/L-ERGO) modified electrodes were characterised by SEM, XRD, EDAX, FT-IR and Raman spectroscopy. Anodic stripping voltammetry analysis and electrochemical impedance spectra were used to understand the electrochemical properties of the synthesized nanostructured material. The modified electrode enabled the electrooxidation of nitrite ions (NO2) with high sensitivity and a good detection limit of 1 nM. Importantly, the modified electrodes were successfully tested for the analyses of generated NO2 in urinary tract infection (UTI) affected patient's urine samples. In addition, the modified electrode was satisfactorily used in the electrooxidation of NO2 generated from in vitro biochemical reduction of nitrate ions using microbial species such as Escherichia coli (E. coli).


1. Introduction

A network of one dimensional (1D) metallic nanomaterials such as nanorods or nanowires on graphene surface has been considered as a promising material for new generation supercapacitors, fuel cells, sensors and optical devices.1–3 Growing of bimetallic nanoparticles on a graphene surface greatly alters its physical and chemical properties.4,5 Compared to other nanostructures, 1D nanostructure such as nanorods and nanowires possess unique characteristics which result in intriguing material characteristics due to synergistic effects such as high optical transparency, specific surface area, good mechanical strength, electrical and thermal conductivity.6 Lee et al. constructed transparent and stretchable electrodes using graphene–metal nanowire hybrid structures.7 Liu et al. developed conductive films using graphene–silver nanowire sandwich like structures.8

Although various metallic nanostructures have been generated on the graphene surface through different techniques like hybridisation, in situ reduction, microwave assisted synthesis and hydrothermal process,7–11 very few reports are available on the direct growth of 1D nanostructures on graphene surface. Kim et al. studied the direct growth of gold nanorods on graphene surface by seedless and seed mediated method.12 Surface functionalization and uniform dispersion in aqueous medium are significant challenges in the growth of 1D nanostructure on graphene surface. Kim et al. have used long chain organic molecules like pyrine ethylene glycol amine and decylpyrine for surface functionalization of reduced graphene oxide (RGO) nanosheets which in turn grow as gold nanorods.13 In seed-mediated growth method, high immobilisation of seed density on RGO surface leads to the growth of spherical nanostructures.13 On the other hand, tuning the surface functionalization on graphene could result in low immobilisation of seed density which in turn paves a way to uniformly generate 1D nanostructure such as nanorods or nanowires on its surface. In the present investigation, two key aspects were taken into consideration for simple and effective growth of bimetallic nanorods on ERGO surface. (1) The poor dispersion quality of RGO in wet chemical method was overcomed by electrochemical reduction of graphene oxide as ERGO on the surface of electrode and (2) in the next step, a simple amino acid L-tryptophan would be introduced to carryout stringent functionalization on ERGO surface. It enabled the immobilization of low density Agseeds on ERGO surface. Currently, various techniques were used to synthesize graphene nanosheets including epitaxial growth,14 chemical vapour deposition,15 mechanical exfoliation16 and chemical reduction of graphene oxide.17 Importantly, in order to scale up the RGO on electrode surface, ERGO was considered as an efficient tool for the synthesis of better quality graphene sheets.18,19 Accordingly, ERGO nanosheets possess low oxygen moieties and better conductivity compared to chemically reduced GO sheets.20

Transition metals like Ag, Au, Pt, Pd, and Cu were commonly used in the formation of heterogeneous catalysts due to its high catalytic efficiency.21,22 Yin et al. developed graphene and Pd-based bimetallic nanocomposites for electrocatalytic oxygen reduction reaction.23 Feng et al. synthesised PtAg alloy–graphene hybrid composites for studying the catalytic activity in methanol electrooxidation.24 Even though various metal nanostructures loaded graphene were extensively studied, the direct growth of bimetallic Ag@Cu NRDs on ERGO nanosheets is not yet explored. In the present work, a simple and an efficient seed mediated growth method was developed for the direct growth of bimetallic Ag@Cu NRDs on L-tryptophan functionalized ERGO.

Nitrite ion (NO2) is commonly present in food, soil and physiological systems.25,26 It is used as a preservative in food industries and an overdose is reported to seriously affect human health. It is formed as carcinogenic nitrosamines in the presence of amine which causes blue body and gastric cancer.27,28 Excess amount of NO2 secreted in the urine samples is known to result in urinary tract infection (UTI). Thus, a positive NO2 test in the urine sample is considered as a crucial indication of UTI.29 Different techniques were developed for the detection of NO2 including spectrophotometry,30 polarography,31 chromatography32 and electrochemical methods.33,34 Electrochemical detection of NO2 has been considered as the most efficient tool with high sensitivity, reproducibility and detection limit. Liu et al. studied Au–Fe(III) nanoparticle modified glassy carbon electrode for electrochemical sensing of NO2 with a detection limit of 2.0 × 10−7 M.33 Lin et al. used poly(3,4-ethylenedioxythiophene) modified SPCEs for NO2 with the detection limit of 0.96 μM.34 Liu et al. used gold nanoparticles for the electrochemical detection of NO2.35 Kalimuthu et al. studied electropolymerized film of functionalized thiadiazole modified glassy carbon electrode for the detection of NO2.36 Pal et al. used Ag nanoparticles modified electrode for the oxidation of NO2.37 Ning et al. used PAMAM dendrimer-stabilized Ag nanoparticles for the electrochemical sensing of NO2 with a detection limit of 0.4 μM.38 Gao et al. used Ag nanorods for the electrooxdiation of NO2.39 Jayabal et al. developed Au/Ag bimetallic nanorods for the electrochemical sensing of NO2.40 Teymourian et al. used Fe3O4 magnetic nanoparticles/reduced graphene oxide nanosheets as a biosensor for NO2 detection.41 Nevertheless, bimetallic nanorods or nanowires decorated reduced graphene oxide nanosheets were used in the development of NO2 sensor. In the present investigation, bimetallic Ag@Cu NRDs decorated L-tryptophan functionalized ERGO electrode exhibits an efficient electrochemical sensing of NO2 with high sensitivity and good detection limit in wide linear range. The interference of various cations and anions were studied in the detection of NO2. In addition, hospital urine samples of UTI affected patients were analysed for the effective practice of NO2 assay at bimetallic Ag@Cu NRDs decorated L-tryptophan functionalized ERGO electrode. Finally, in vitro biochemical conversion of NO3 to NO2 was carried out using E. coli to study NO2 generation at the modified electrode.

2. Experimental section

2.1. Chemicals

Graphite powder (300 mesh) was purchased form Alfa Aesar, UK. Sodium nitrate, sulphuric acid, hydrogen peroxide, sodium borohydride, copper nitrate, L-tryptophan, sodium hydroxide, sodium nitrite, disodium hydrogen phosphate and monosodium dihydrogen phosphate were purchased form Merck, India. Cetyl trimethyl ammonium bromide (CTAB) was obtained from Himedia, India, and trisodium citrate, ascorbic acid (AA) and silver nitrate were obtained from Sigma-Aldrich, USA. All reagents were of analytical grade and used without further purification. All solutions were prepared using deionized (DI) water.

2.2. Characterizations

In order to exhibit the structural and morphological properties, bimetallic Ag@Cu NRDs decorated ERGO nanosheets were grown on Indium tin oxide (ITO) coated electrode surface. Prior to the growth process, ITO plates were cleaned and pretreated in H2O, NH4OH and H2O2 with the ratio of 5[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1. Surface morphology of bimetallic Ag@Cu NRDs on ERGO nanosheet was confirmed by SEM (VEGA 3 TESCAN, USA) analysis. X-ray diffraction (XRD) patterns were studied using Shimadzu XRD 6000 (Japan) with Cu-Kα radiation (λ = 1.5418 Å). EDAX mapping analyses were carried out using energy dispersive X-ray analyser (Bruker, Germany). FT-IR (Jasco 4000 series, Japan) analyses were carried out to study the surface functionalization of graphene nanosheets using L-tryptophan. Raman spectra were recorded using Horiba-Jobin, LabRAM HR equipped with Argon (514 nm) ion laser excitation source.

2.3. Electrochemical measurements

All electrochemical measurements were carried out using CHI 660D electrochemical workstation (CH Instruments, USA). The electrochemical cell consist of a three electrode system, Ag@Cu/L-ERGO modified glassy carbon electrode (0.07 cm2) serves as a working electrode, platinum wire as a counter electrode and Ag/AgCl filled with saturated KCl solution was used as a reference electrode.

2.4. Synthesis of L-tryptophan functionalized ERGO electrode

Graphite oxide was prepared by modified hummers method.42,43 Briefly, 2 g of natural graphite powder was mixed with 1 g of NaNO3 and 25 mL of conc. H2SO4 and stirred for 30 min at 5 °C, followed by stepwise addition of 7 g KMnO4 over a period 1 h. The reaction mixture was stirred further for another 2 h at 35 °C, followed by the addition of 92 mL of DI water slowly into the reaction mixture that eventually turned into brown colour. Finally, 10 mL H2O2 (30%) and 140 mL of warm DI water were added. The filtrate was collected by washing with 5% HCl and DI water and dried overnight at 50 °C. The synthesised graphite oxide should be dispersed in a fine concentration of 0.2 mg mL−1 in DI water under ultrasonication for 3 h and thereby resulted in exfoliation of graphene oxide (GO). The well dispersed GO thus synthesized is stable in DI water for more than 6 months.

Glassy carbon electrodes (GCEs) were polished well with alumina (0.3 micron) using buehler cloth and sonicated in ethanol–water mixture for 5 min. Then, 10 μL of exfoliated GO was drop casted on GCE surface and dried at room temperature for 1 h. GO was electrochemically reduced in 0.05 M phosphate buffer (pH 5).18,44 Using three electrode cell set up, 30 successive cycles were performed in the potential range between 0.0 and −1.5 V at a scan rate of 50 mV s−1. Further, the ERGO modified electrode was dipped in L-tryptophan (1 mg mL−1) for 24 h at room temperature for surface functionalization.

2.5. Synthesis of bimetallic Ag@Cu NRDs on L-tryptophan functionalized ERGO electrode

Bimetallic Ag@Cu NRDs were grown on functionalized ERGO surface by seed mediated growth method. First, Ag seeds (Agseeds) solution was prepared by adding 0.6 mL of 10 mM NaBH4 to an aqueous solution containing 0.25 mM AgNO3 and 0.25 mM trisodium citrate. L-tryptophan functionalized ERGO/GCE was dipped in Agseeds solution for 30 min and dried for 1 h. In the next step, Agseeds/ERGO/GCEs were dipped into growth solution containing 80 mM CTAB for 3 min. Subsequently, 1 mL of 10 mM Cu(NO3)2, 2 mL of 100 mM AA, 0.4 mL of 1 M NaOH were added to the growth solution for the anisotropic growth of bimetallic Ag@Cu NRDs on the electrode surface. A gradual change in colour of the growth solution from colourless to bright yellow over a period of 15 min indicated the generation of bimetallic Ag@Cu NRDs on functionalized ERGO surface. Importantly, it is possible to control the aspect ratio of bimetallic NRDs generated by varying the concentration of Agseeds on ERGO surface.45,46 Thus, low loading Agseeds on functionalized ERGO surface would tend to grow long uniform bimetallic NRDs. Accordingly, the growth process was repeated to increase the loading density of bimetallic Ag@Cu NRDs on functionalized ERGO surface.

3. Results and discussion

The direct growth of bimetallic Ag@Cu NRDs on ERGO surface by seed mediated growth method is outlined in Scheme 1. It explained the 1D growth of Agseeds into bimetallic Ag@Cu NRDs. It was carried out by the interaction of crystal faces of Agseeds with the soft template CTAB and the anisotropic reduction of Cu2+ ions using mild reducing agent such as AA. The direct growth of 1D NRDs or nanowires on RGO surfaces is a challenge due to poor surface functionalization methodologies. Seed immobilisation density on graphene surface is altered by effective surface functionalization. In the present work, the direct growth of Agseeds in the presence and absence of functionalizing agent were compared. L-tryptophan was used as the surface functionalizing agent paving a way for the direct growth of bimetallic Ag@Cu NRDs at ERGO surface. L-Tryptophan plays a vital role in the coordination of metal ions by the electrostatic interaction, reduction of metal ions and in the direct growth of bimetallic NRDs at ERGO surface. Fig. 1 shows the SEM images of ERGO (A), Ag@Cu nanoparticles on ERGO (B) and bimetallic Ag@Cu NRDs generation on L-tryptophan functionalized ERGO (C) nanosheets decorated electrodes. Fig. 1A represents the sheet like morphology of ERGO nanosheets. Fig. 1B clearly shows the formation of spherical nanostructures on graphene surface. In the absence of L-tryptophan, Agseeds grew densely as Ag@Cu spherical nanoparticles on bare ERGO surface. On the other hand, the presence of L-tryptophan on ERGO surface facilitated the successful growth of Agseeds into Ag@Cu bimetallic NRDs. Fig. 1C confirmed the formation of 1D Ag@Cu NRDs on L-tryptophan functionalized ERGO surface with an average nanorods diameter as 180 ± 2 nm and length as ∼1 μm.
image file: c5ra05730b-s1.tif
Scheme 1 Schematic representation of 1D growth of bimetallic Ag@Cu NRDs on L-tryptophan functionalized ERGO nanosheets at GCE surface.

image file: c5ra05730b-f1.tif
Fig. 1 SEM images of ERGO (A), Ag@Cu nanoparticles decorated at ERGO in the absence of L-tryptophan (B) and Ag@Cu NRDs decorated at L-ERGO (C).

Functionalization of ERGO surface with L-tryptophan was confirmed by FT-IR spectroscopy. Fig. 2 shows the FT-IR spectra of GO (a), ERGO (b), L-ERGO (c) and L-tryptophan (d). The broad band observed for GO at 3461 cm−1 corresponds to O–H stretching vibration. Sharp bands at 1730 cm−1, 1620 cm−1 1224 cm−1 and 1048 cm−1 were due to stretching vibrations of C[double bond, length as m-dash]O, C[double bond, length as m-dash]C, epoxy groups and C–O group, respectively. Subsequently, the corresponding bands for oxygen functionalities and epoxy groups were absent which confirmed the electrochemical reduction of GO (a) into ERGO (b). Importantly, FT-IR spectral tool inferred the pure form of electrochemically reduced graphene nanosheets compared to chemically reduced graphene oxide nanosheets.20 The C[double bond, length as m-dash]C stretching occur at 1650 cm−1 and the broad bands at 1085 cm−1and 950 cm−1 were responsible for C–H vibrations. L-Tryptophan functionalized ERGO (c) shows intense bands at 3604 cm−1, 3401 cm−1, and 1700 cm−1 corresponding to O–H stretching, N–H stretching, and C[double bond, length as m-dash]O stretching, respectively. The C[double bond, length as m-dash]C stretching and C–H bending vibrations occur at 1514 cm−1 and 680 cm−1. A weak band at 2920 cm−1 is responsible for the stretching vibrations of CH2 group. All these bands were correlated with the FT-IR spectra of L-tryptophan (d). Thus, FT-IR spectral results confirmed the N–H stretching which validates the effective functionalization of L-tryptophan on ERGO surface (c). The ERGO surface intrinsically does not contain heteroatom like nitrogen, and it must therefore solely be derived from N–H functionalization of L-tryptophan where Agseeds immobilize and direct the seed mediated growth of bimetallic Ag@Cu NRDs.


image file: c5ra05730b-f2.tif
Fig. 2 FT-IR transmittance spectra of GO (a), ERGO (b), L-ERGO (c) and L-tryptophan (d).

EDAX spectra and elemental mapping analyses were used to confirm the coexistence of bimetallic Ag and Cu on the surface of ERGO. Fig. 3 shows the EDAX pattern and elemental mapping analyses of bimetallic Ag and Cu on L-ERGO surface. Briefly, in the first step, Agseeds were immobilised on L-tryptophan functionalized ERGO surface which corresponds to the identification of metallic Ag (Fig. 3A). In the next step, using mild reducing agent, the existing Agseeds tend to mediate the growth of Cu on its surface. The existence of Cu on Ag surface was confirmed with the identification of Cu in mapping analysis (Fig. 3B). Fig. 3C clearly confirmed the coexistence of Ag and Cu on bimetallic Ag@Cu NRDs on ERGO surface. Finally, EDAX spectra (Fig. 3D) authenticated the presence of bimetallic Ag and Cu at NRDs on functionalized ERGO surface with a weight percentage ratio of ∼10[thin space (1/6-em)]:[thin space (1/6-em)]1. Further, XRD pattern (Fig. 4) justified the presence of bimetallic Ag and Cu at L-tryptophan functionalized ERGO surface. Graphene oxide shows an intense diffraction at 10.1° whereas it is minimised and a new diffraction is observed at 23.9° for ERGO surface or at Ag@Cu NRDs decorated L-ERGO surface. The change in the diffraction pattern is due to the complete reduction of GO with an interlayer spacing of 3.7 Å.47 In addition, the XRD pattern at 17.1°, 20.6°, 27.5° corresponds to Ag with (1 1 1), (2 0 0) and (2 2 0) facets and at 20.6° (1 1 1), 23.5° (2 0 0) and 38.2° (3 1 1) corresponds to the metallic Cu. All diffraction patterns exhibited good correlation with JCPDS no. 011167 and 011242, respectively.


image file: c5ra05730b-f3.tif
Fig. 3 EDAX elemental mapping analyses of Agseeds (A) or Cu at Ag@Cu (B) or bimetallic Ag@Cu at L-ERGO (C), and EDAX spectra of Ag@Cu NRDs on L-ERGO nanosheets coated electrode (D).

image file: c5ra05730b-f4.tif
Fig. 4 XRD patterns of ERGO (a), Ag@Cu NRDs/L-ERGO (b) nanosheets coated electrodes. Inset: XRD patterns of GO.

Raman spectroscopy is an effective and non-destructive method to characterize the crystal structure and defects in graphene based materials. D and G band in the Raman spectra are responsible for the κ-point phonons of the A1g symmetry and E2g vibration mode of sp2 carbon atoms. The intensity ratios of D and G bands (ID/IG) are the measure of defects on graphene layers. Fig. 5 shows the Raman spectra of ERGO and bimetallic Ag@Cu NRDs decorated L-ERGO. GO shows sharp bands at 1350 cm−1 and 1594 cm−1 corresponding to D and G with an ID/IG ratio of 0.90 (Fig. 5 (inset)). The ratio increased to 1.42 due to RGO formation, reflecting the rise in defect concentration as well as disorderliness at RGO surface.48,49 On the other hand, the ID/IG ratio decreased to 1.03 at bimetallic Ag@Cu NRDs decorated L-ERGO with sharp bands at 1345 cm−1 and 1586 cm−1. This is due to the growth of bimetallic Ag@Cu NRDs on ERGO producing a healing effect which reflects in the increase of sp2 domain character. 2D and S3 bands at 2681 cm−1 and 2943 cm−1 confirmed better graphitization process. In addition, the intensity of 2D band was attributed to the layer stacking in graphene.50


image file: c5ra05730b-f5.tif
Fig. 5 Raman spectra of ERGO (a) and Ag@Cu NRDs/L-ERGO (b) modified electrodes. Inset: Raman spectra of GO.

Anodic stripping voltammetry (ASV) is the most effective, reproducible and sensitive method for the detection of trace amounts of metal ions at the electrode surface. The existence of metallic Ag and Cu at L-ERGO modified electrode was analyzed by ASV. Fig. 6 shows the cyclic voltammograms (CVs) of Agseeds/L-ERGO/GCE (a) and bimetallic Ag@Cu NRDs/L-ERGO/GCE (b) in 0.5 M H2SO4. The anodic peak appeared at 0.23 V corresponding to Ag oxidation at Agseeds/L-ERGO/GCE (a) and at 0.12 V corresponding to bimetallic Cu and Ag oxidation at Ag@Cu NRDs/L-ERGO/GCE (b), which coincides with the stripping potential of Cu which is in the range of 0.1 V.51 Hence, the decrease in the anodic stripping potential at Fig. 6b was attributed to the presence of Cu at Ag surface.52 Thus, the presence of Cu at the bimetallic NRDs surface tends to decrease the over potential of Ag metal dissolution and a concurrent oxidation of bimetallic Cu and Ag with decreased oxidation potential. This result confirms the presence of bimetallic Ag and Cu at L-ERGO surface.


image file: c5ra05730b-f6.tif
Fig. 6 CVs of Agseeds/L-ERGO (a) and Ag@Cu NRDs/L-ERGO/GCE (b) in 0.5 M H2SO4 at a scan rate of 50 mV s−1.

Electrochemical impedance spectroscopy is an effective tool for studying the interfacial properties of modified electrodes. Fig. 7 shows the Nyquist plot for ERGO/GCE (a), Agseeds/L-ERGO/GCE (b) and Ag@Cu NRDs/L-ERGO/GCE (c) in the electrolyte containing 0.5 mM K4Fe(CN)6 and 0.5 M KCl. All the impedance data are fitted with Randle's equivalent circuit. Ag@Cu bimetallic NRDs decorated at L-ERGO (c) shows a very low charge transfer resistance (Rct) compared to ERGO (a) and Agseeds decorated L-ERGO electrode (b). This indicates the importance of bimetallic NRD formation at graphene surface resulting in enhancement of electron transfer kinetics between the electrolyte and electrode interface with increase in conductivity.


image file: c5ra05730b-f7.tif
Fig. 7 Electrochemical impedance spectra of ERGO (a), Agseeds/L-ERGO (b) and Ag@Cu NRDs/L-ERGO (c) in 0.5 mM K4Fe(CN)6. The data's are fitted with Randle's equivalent circuit.

3.1. Electrooxidation of NO2 at bimetallic Ag@Cu NRDs/L-ERGO/GCE

The bimetallic modified electrodes were tested for the electrochemical sensing of NO2 by investigating the electrooxidation of NO2 in phosphate buffer (PB), pH 6. Fig. 8 shows the CVs at Ag@Cu NRDs/L-ERGO/GCE in the presence (a) and absence (b) of 1 mM NO2 in PB at a scan rate of 50 mV s−1. In the presence of NO2, Ag@Cu NRDs/L-ERGO/GCE (a) shows an intense electrooxidation peak at 0.78 V with an onset potential at 0.46 V due to the oxidation of NO2 to NO3 through a two electron process.53 On the other hand, in the absence of NO2, no such oxidative behaviour was observed (Fig. 8b). Further, for comparison, bare GCE (a), L-ERGO/GCE (b) and Agseeds/L-ERGO/GCE (c) were used to investigate the electrooxidation of NO2 under identical conditions as displayed in Fig. 8 (inset). Bare GCE shows a broad oxidation peak at 0.92 V with an increase in the overpotential towards NO2 oxidation, Fig. 8 (inset a). Interestingly, after L-ERGO deposition GCE shows an increased peak current towards NO2 electrooxidation with a broad peak at 1.05 V (Fig. 8 (inset b)). On the other hand, no noticeable difference in the peak current towards NO2 electrooxidation was observed after Agseeds was introduced at L-ERGO/GCE surface (Agseeds/L-ERGO/GCE), Fig. 8 (inset c). A new oxidative peak at 0.26 V which corresponds to the dissolution of Agseeds in PB buffer was observed in addition to NO2 oxidation. Scheme 2 illustrates the unfavorable and favorable conditions for electrooxidation of NO2 at Agseeds/L-ERGO/GCE (A) and Ag@Cu NRDs/L-ERGO/GCE (B), respectively. Scheme 2(A) shows the poor electrooxidation of NO2 at Agseeds modified electrode which is due to electrostatic repulsive behaviour between the surface plasmon of Agseeds and NO2 ions. On the other hand, in the case of bimetallic Ag@Cu NRDs, the heterometallic bonding interactions (ligand effect) between the metals resulted in the modification of surface plasmon of Agseeds.54 It greatly enhanced the catalytic effect of Ag@Cu NRDs on ERGO for the electrooxidation of NO2 which is depicted in Scheme 2(B). Thus, seed mediated growth approach of Cu coated Ag as bimetallic Ag@Cu NRDs generation is absolutely essential at ERGO surface to minimize the electrostatic repulsion and to subsequently enhance the electrooxidative behaviour towards NO2 oxidation. Accordingly, Ag@Cu bimetallic NRDs decorated ERGO electrode shows an increased electrochemical activity for the electrooxidation of NO2 with decrease in overpotential of 0.26 V compared to L-ERGO/GCE or Agseeds/L-ERGO/GCE.
image file: c5ra05730b-f8.tif
Fig. 8 Cyclic voltammetric response of Ag@Cu NRDs/L-ERGO/GCE in the presence (a) and absence of 1 mM NO2 in PB, pH 6. Inset: CVs of bare GCE (a), L-ERGO/GCE (b) and Agseeds/L-ERGO/GCE (c) at 1 mM NO2 in PB.

image file: c5ra05730b-s2.tif
Scheme 2 Schematic representation of electrooxidation of NO2 at Agseeds/L-ERGO/GCE (A) and Ag@Cu/L-ERGO/GCE (B).

The potential scan rate study is an important factor that might affect the kinetics of the electrode process. Therefore, the effect of scan rate on the electrooxidation of NO2 at Ag@Cu NRDs/L-ERGO/GCE was studied. Fig. 9 shows the CVs of Ag@Cu NRDs/L-ERGO/GCE in PB containing 1 mM NO2 under different scan rate from 0.01 to 0.1 mV s−1. It shows a positive shift in the oxidation potential with respect to increase in scan rate. Fig. 9 (inset) shows a linear relationship between the anodic peak current and the square root of scan rate with the linear regression equation Ipa = 85.72 (V s−1) + 13.09. The linear increase in the anodic peak current with respect to the square root of scan rate indicates that the electrooxidation of NO2 at Ag@Cu NRDs/L-ERGO/GCE is a diffusion controlled process. It is used to calculate the number of electrons transferred in the electrooxidation process with Randles–Sevcik equation.

 
Ip = 0.4463(F3/RT)1/2Anα3/2Do1/2Coν1/2 (1)


image file: c5ra05730b-f9.tif
Fig. 9 CVs of Ag@Cu NRDs/L-ERGO/GCE in 1 mM NO2 in PB, pH 6 at 0.01 (a), 0.03 (b), 0.05 (c), 0.07 (d) and 0.09 (e) V s−1. Inset: plot of the peak currents vs. square root of scan rates.

Here nα the no. of electrons transferred, Do is the diffusion coefficient of the electroactive species, Co is the concentration of NO2 and ν1/2 is the square root of scan rate. Hence, the number of electrons transferred is estimated as 2.1 using diffusion coefficient as 2.1 × 10−5 cm2 s−1 for NO2 electrooxidation in 0.1 M PB.55,56 The oxidation of NO2 to NO3 involves two step reactions as eqn (2) and (3).34,36

 
2NO2 ⇌ 2NO2 + 2e (2)
 
2NO2 + H2O → NO3 + NO2 + 2H+ (3)

The oxidation of NO2 to NO2 is the rate determining step. The electron transfer coefficient (α) of the reaction can be calculated from eqn (4)

 
image file: c5ra05730b-t1.tif(4)
where nα is the number of electrons transferred in the rate determining step. Since it is a two electron process the estimated value of α is 0.674.

The effect of pH on the electrooxidation of NO2 with respect to anodic peak current and anodic peak potential were investigated between pH 4 and 8 at Ag@Cu NRDs/L-ERGO/GCE. Fig. 10 shows the variations in the anodic peak current with respect to pH from 4 to 8. Fig. 10 (inset) shows the plot of current vs. pH where a slight decrease and linear increase in anodic peak current was observed from pH 4 to 5 and up to pH 6. Further increase in pH resulted in a drastic decrease in the anodic peak current. On the other hand, there was no change in the anodic peak potential between pH 4 and 5, whereas the oxidation potential shifted to a less positive range at pH 6. From pH 6 to 8, a linear increase in the oxidation potential was observed with a linear regression equation y = 0.0155x + 0.6978 (R2 = 0.9914). The number of exchanged protons were calculated from the equation dE/dpH = 0.059X/αnα. Here, α is the electron transfer coefficient and nα is the no. of electrons. From the slope, the no. of protons is estimated as 0.3. It indicates that in acidic media, there is no change in peak potential and in basic media, the estimated proton range is very low. Thus, pH 6 offers an appropriate proton environment for the electrocatalytic oxidation of NO2 due to the electrocatalytic effect of Ag@Cu NRDs decorated L-ERGO nanosheets. The number of protons calculated from the slope of peak potential vs. pH is inconsistent with theoretical two proton oxidation reaction of NO2. Therefore, the electrooxidation of NO2 is speculated as a kinetically controlled process as discussed elsewhere.57,58 Thus, pH 6 was found to be optimal for enhanced electrooxidation of NO2 at Ag@Cu NRDs/L-ERGO/GCE. Subsequent NO2 electrooxidation studies were carried out in PB at pH 6.


image file: c5ra05730b-f10.tif
Fig. 10 CVs of Ag@Cu NRDs/L-ERGO/GCE at different pH for 1 mM NO2 [pH: 4 (a), 5 (b), 6 (c), 7 (d) and 8 (e)]. Inset: plot of pH vs. anodic peak current.

Ag@Cu NRDs/L-ERGO/GCE was used to study the linear range of NO2 assay and detection limit. Fig. 11A shows the chronoamperometric current response for the successive addition of an aliquot of NO2 after every 40 s in PB. It indicates the steady increase in the oxidation current for [NO2] from 0.1 mM to 0.9 mM. Fig. 11A (inset) shows the linear relationship between the oxidative current vs. [NO2] with a correlation coefficient of 0.998. The detection limit of NO2 was estimated as 1 × 10−9 M with a signal to noise ratio 3 (S/N = 3) at Ag@Cu NRDs/L-ERGO modified electrode (Fig. 11B). The interference study of a 100 fold higher concentration of different cations and anions such as Li+, Na+, K+, Mg2+, Ca2+, PO43−, SO42−, NO3−, Cl and CO3 were carried out in the presence of 0.1 mM NO2 using chronoamperometry. None of the aforementioned ions interfered in the electrochemical sensing of NO2 at Ag@Cu NRDs/L-ERGO/GCE. The reproducibility and stability of the Ag@Cu NRDs/L-ERGO/GCE were studied by cyclic voltammetry. After 20 successive cycles, Ag@Cu NRDs/L-ERGO/GCE shows a very good response toward NO2 electrooxidation with a relative standard deviation of 3.3%. Further, the modified GCE was stored and studied after 3 weeks at room temperature and exhibit a similar current response toward NO2 electrooxidation. This indicates that the modified electrode possess good reproducibility and stability.


image file: c5ra05730b-f11.tif
Fig. 11 (A) Chronoamperometric response of Ag@Cu NRDs/L-ERGO/GCE for 0.1–0.9 mM [NO2] in PB, pH 6. The electrode potential is fixed at −0.85 V. Inset plot shows the [NO2] vs. anodic peak current. (B) CVs of modified electrode in the presence (a) and absence (b) of 1 nM NO2 in PB.

3.2. Real sample analysis

UTI is one of the most common bacterial infection affecting both men and women. E. coli (80–85%) is the prominent bacteria for the cause of UTI and it converts the urinary nitrate into NO2. A positive NO2 test is therefore an indication of UTI. Thus the modified electrode should enable the analysis of hospital urine samples of UTI affected patients for detecting the presence of NO2 at Ag@Cu NRDs/L-ERGO/GCE. Prior to the test, the hospital urine samples were centrifuged and supernatant solutions were used for analyses. Fig. 12 shows the electrooxidation signal of NO2 in the urine samples of UTI affected patient (a) compared with non infected urine sample (b) at Ag@Cu NRDs/L-ERGO/GCE. UTI patient sample shows a broad oxidation in the range of 0.6–0.9 V in Fig. 12a whereas no such oxidation was observed in the control experiment (Fig. 12b). In order to validate the result, differential pulse voltammetry was used to confirm the electrooxidation corresponding to NO2. Fig. 12 (inset a) shows a noticeable oxidation peak at 0.75 V due to NO2 oxidation in the urine sample of UTI patient whereas no such oxidation peak was observed in the case (Fig. 12 (inset b)) of non infected person. The concentration of nitrite at UTI infected samples was calculated as 0.1 mM through chronoamperometry and 90 ± 4% recovery of nitrite was obtained with urine samples. The decrease in the recovery is due to the matrix effect that exist in urine samples. The modified electrode was successfully used in the rapid screening of UTI affected urinary samples, proving it to be a highly sensitive electrochemical sensor for NO2 detection.
image file: c5ra05730b-f12.tif
Fig. 12 Cyclic voltammetric responses of Ag@Cu NRDs/L-ERGO/GCE for hospital urine sample of UTI patient (a) and a urine sample of non-infected person (b) in PB, pH 6. Inset: differential pulse voltammograms of same urine samples (a) and (b).

In addition, E. coli bacteria were used for in vitro biochemical reduction of sodium nitrate to sodium nitrite and the generated NO2 was analyzed at Ag@Cu NRDs/L-ERGO/GCE. 5 × 107 cfu mL−1 of E. coli was added into 10 mL of 5 mM NO3 in PB and incubated for 20 h at room temperature.29 The control experiment did not contain E. coli. The solutions were centrifuged and the supernatants were subjected for electrochemical analyses. Fig. 13 shows the cyclic voltammetric response of supernatant solutions obtained in the presence (a) and absence (b) of E. coli. An oxidative peak was observed at 0.72 V in the case of supernatant containing E. coli (a) which is due to the electrooxidation of NO2. Thus, the biochemical reduction of nitrate into nitrite ions in the presence of E. coli was confirmed. On the other hand, there was no such oxidation peak was observed in the absence of E. coli for 5 mM NO3. In addition, in the absence of 5 mM NO3 under identical conditions of PB and E. coli was incubated for 20 h. Then the resultant supernatant was subjected for electrochemical analysis, Fig. 13c. The control experiments didn't show any characteristic cyclic voltammetric responses. Fig. 13 (inset) shows a characteristic DPV peak corresponding to electrooxidation of NO2 generated in the presence of E. coli. These investigations permit us to validate the successful utilization of bimetallic Ag@Cu NRDs decorated L-ERGO electrodes in the sensing of NO2 at hospital urine samples.


image file: c5ra05730b-f13.tif
Fig. 13 Cyclic voltammetric responses of biochemically generated NO2 due to the presence (a) and absence (b) of E. coli containing 5 mM NO3 and in the absence of 5 mM NO3 (c) under identical conditions of PB and E. coli at Ag@Cu NRDs/L-ERGO/GCEs. Inset: differential pulse voltammogram of biochemically generated NO2 due to the presence of E. coli.

4. Conclusions

Ag@Cu bimetallic NRDs were successfully grown on ERGO surface by simple seed mediated method. Importantly, a dual role for L-tryptophan was identified; (1) as a functionalizing agent for controlling the Agseeds density on ERGO surface and (2) in the direct growth of bimetallic Ag@Cu NRDs on the surface of ERGO. The modified electrode was characterized with ease using SEM, XRD, EDAX, FT-IR and Raman spectroscopy. Ag@Cu NRDs/L-ERGO modified electrodes were used for the electrooxidation of NO2 with amplified current response and good detection limit was achieved as low as 1 nM. In addition, the overpotential for NO2 oxidation was decreased at the modified electrode compared to bare GCE. It indicates the synergistic effect produced at the coexistence of bimetallic Ag@Cu NRDs on L-ERGO surface which drastically enhanced the electrochemical and catalytic effect of graphene nanosheets towards NO2 detection with a fast electron transfer rate. The real time application of the modified electrode was satisfactorily proved in the analyses of NO2 containing hospital urine samples of UTI patients with ease. Further, the modified electrode could be commercialized by generating the patterns of low cost bimetallic Ag@Cu NRDs at L-ERGO surface based screen printed electrode for analyzing the hospital urine sample of UTI in short duration of time with very high accuracy.

Acknowledgements

Financial support by DST-SERB, New Delhi (File no. SR/FT/CS-44/2011 dated 04.05.2012) is gratefully acknowledged. We thank Karunya University, Coimbatore for providing instrumental facilities. This work is dedicated to Professor R. Ramaraj's 60th birthday and his pioneering contribution in the fields of Photochemistry & Electrochemistry.

References

  1. R. Zou, G. He, K. Xu, Q. Liu, Z. Zhang and J. Hu, J. Mater. Chem. A, 2013, 1, 8445 CAS.
  2. C. Xu, J. Sun and L. Gao, J. Mater. Chem., 2012, 22, 975 RSC.
  3. J. M. Lee, J. Yi, W. W. Lee, H. Y. Jeong, T. Jung, Y. Kim and W. I. I. Park, Appl. Phys. Lett., 2012, 100, 061107 CrossRef PubMed.
  4. M. Wang, J. Huang, M. Wang, D. Zhang, W. Zhang, W. Li and J. Chen, Electrochem. Commun., 2013, 34, 299 CrossRef PubMed.
  5. C. H. Liu, R. H. Liu, Q. J. Sun, J. B. Chang, X. Gao, Y. Liu, S. T. Lee, Z. H. Kang and S. D. Wang, Nanoscale, 2015, 7, 6356 RSC.
  6. J. L. Xie, C. X. Guo and C. M. Li, Energy Environ. Sci., 2014, 7, 2559 CAS.
  7. M. S. Lee, K. Lee, S. Y. Kim, H. Lee, J. Park, K. H. Choi, H. K. Kim, D. G. Kim, D. Y. Lee, S. W. Nam and J.-U. Park, Nano Lett., 2013, 13, 2814 CrossRef CAS PubMed.
  8. B. T. Liu and H. L. Kuo, Carbon, 2013, 63, 390 CrossRef CAS PubMed.
  9. Z. Luo, L. Yuwen, B. Bao, J. Tian, X. Zhu, L. Weng and L. Wang, J. Mater. Chem., 2012, 22, 7791 RSC.
  10. H. M. A. Hassan, V. Abdelsayed, A. E. R. S. Khder, K. M. AbouZeid, J. Terner, M. S. E. Shall, S. I. A. Resayes and A. A. E. Azhary, Mater. Chem., 2009, 19, 3832 RSC.
  11. K. Liu, L. Liu, Y. Luo and D. Jia, J. Mater. Chem., 2012, 22, 20342 RSC.
  12. Y. K. Kim, H. K. Na, Y. W Lee, H. Jang, S. W. Han and D. H. Min, Chem. Commun., 2010, 46, 3185 RSC.
  13. Y. K. Kim, H. K. Na and D. H. Min, Langmuir, 2010, 26, 13065 CrossRef CAS PubMed.
  14. C. Berger, Z. M. Song, X. B. Li, X. S. Wu, N. Brown, C. Naud, D. Mayou, T. B. Li, J. Hass, A. N. Marchenkov, E. H. Conrad, P. N. First and W. A. D. Heer, Science, 2006, 312, 1191 CrossRef CAS PubMed.
  15. S. Park and R. S. Ruoff, Nat. Immunol., 2009, 4, 217 CAS.
  16. K. S. Novoselov, A. K. Geim, S. V. Morozov, D. Jiang, Y. Zhang, S. V. Dubonos, V. Grigorieva and A. A. Firsov, Science, 2004, 306, 666 CrossRef CAS PubMed.
  17. S. Stankovich, D. A. Dikin, G. H. B. Dommett, K. M. Kohlhaas, E. J. Zimney, E. A. Stach, R. D. Piner, S. T. Nguyen and R. S. Ruoff, Nature, 2006, 442, 282 CrossRef CAS PubMed.
  18. M. A. Raj and S. A. John, J. Phys. Chem. C, 2013, 117, 4326 CAS.
  19. X. Xu, D. Huang, K. Cao, M. Wang, S. M. Zakeeruddin and M. Gratzel, Sci. Rep., 2013, 3, 1489 Search PubMed.
  20. Z. Wang, X. Zhou, J. Zhang, F. Boey and H. Zhang, J. Phys. Chem. C, 2009, 113, 14071 CAS.
  21. S. Xie, M. Jin, J. Tao, Y. Wang, Z. Xie, Y. Zhu and Y. Xia, Chem.–Eur. J., 2012, 18, 14974 CrossRef CAS PubMed.
  22. J. Chen, B. Wiley, J. McLellan, Y. Xiong, Z. Y. Li and Y. Xia, Nano Lett., 2005, 5, 2058 CrossRef CAS PubMed.
  23. H. Yin, S. Liu, C. I. Zhang, J. Bao, Y. Zheng, M. Han and Z. Dai, ACS Appl. Mater. Interfaces, 2014, 6, 2086 CAS.
  24. L. Feng, G. Gao, P. Huang, X. Wang, C. Zhang, J. Zhang, S. Guo and D. Cui, Nanoscale Res. Lett., 2011, 6, 551 CrossRef PubMed.
  25. C. L. Walters, Oncology, 1980, 37, 289 CrossRef CAS PubMed.
  26. Y. Zhang, R. Yuan, Y. Q. Chai, W. J. Li, X. Zhong and H. A. Zhong, Biosens. Bioelectron., 2011, 26, 3977 CrossRef CAS PubMed.
  27. W. Lijinsky and S. S. Epstein, Nature, 1970, 225, 21 CrossRef CAS PubMed.
  28. E. Morcos and N. P. Wiklund, Electrophoresis, 2001, 22, 2763 CrossRef CAS.
  29. S. Carlsson, M. Govoni, N. P. Wiklund, E. Weitzberg and J. O. Lundberg, Antimicrob. Agents Chemother., 2003, 47, 3713 CrossRef CAS.
  30. M. Grau, U. B. H. Cotta, P. Brouzos, C. Drexhage, T. Rassaf, T. Lauer, A. Dejam, M. Kelm and P. Kleinbongard, J. Chromatogr., 2007, 851, 106 CAS.
  31. Z. Zhao and X. Cai, J. Electroanal. Chem., 1988, 252, 361 CrossRef CAS.
  32. L. E. Vanatta, J. Chromatogr. A, 2008, 1213, 70–76 CrossRef CAS PubMed.
  33. T. S. Liu, T. F. Kang, L. P. Lu, Y. Zhang and S. Y. Cheng, J. Electroanal. Chem., 2009, 632, 197 CrossRef CAS PubMed.
  34. C. Y. Lin, V. S. Vasantha and K. C. Ho, Sens. Actuators, B, 2009, 140, 51 CrossRef CAS PubMed.
  35. Y. Liu and H. Y. Gu, Microchim. Acta, 2008, 162, 101 CrossRef CAS.
  36. P. Kalimuthu and S. A. John, Electrochem. Commun., 2009, 11, 1065 CrossRef CAS PubMed.
  37. M. Pal and V. Ganesan, Analyst, 2010, 135, 2711 RSC.
  38. D. Ning, H. Zhang and J. Zheng, J. Electroanal. Chem., 2014, 717, 29 CrossRef PubMed.
  39. P. Gao, M. Zahang, H. Hou and Q. Xiao, Mater. Res. Bull., 2008, 43, 531 CrossRef CAS PubMed.
  40. S. Jayabal and R. Ramaraj, Electrochim. Acta, 2013, 88, 51 CrossRef CAS PubMed.
  41. H. Teymourian, A. Salimi and S. Khezrian, Biosens. Bioelectron., 2013, 49, 1 CrossRef CAS PubMed.
  42. D. C. Marcano, D. V. Kosynkin, J. M. Berlin, A. Sinitskii, Z. Sun, A. Slesarev, L. B. Alemany, W. Lu and J. M. Tour, ACS Nano, 2010, 4, 4806 CrossRef CAS PubMed.
  43. P. Gnanaprakasam and T. Selvaraju, RSC Adv., 2014, 4, 24518 RSC.
  44. S. Cheemalapati, S. Palanisamy and S. M. Chen, Int. J. Electrochem. Sci., 2013, 8, 3953 CAS.
  45. N. R. Jana, L. Gearheart and C. J. Murphy, Chem. Commun., 2001, 617 RSC.
  46. K. Aslan, Z. Leonenko, J. R. Lakowicz and C. D. Geddes, J. Phys. Chem., 2005, 109, 3157 CrossRef CAS PubMed.
  47. M. O Danilov, I. A. Slobodyanyuk, I. A. Rusetskii and G. Y. Kolbasov, J. Nanostructure Chem., 2013, 3, 49 CrossRef.
  48. H. L. Guo, X. F. Wang, Q. Y. Qian, F. B. Wang and X. H. Xia, ACS Nano, 2009, 3, 2653 CrossRef CAS PubMed.
  49. Z. Wang, S. Wu, J. Zhang, P. Chen, G. Yang, X. Zhou, Q. Zhang, Q. Yan and H. Zhang, Nanoscale Res. Lett., 2012, 7, 161 CrossRef PubMed.
  50. V. C. Tung, M. J. Allen, Y. Yang and R. B. Kaner, Nat. Immunol., 2009, 4, 25 CAS.
  51. M. Kopanica and F. Vydra, J. Electroanal. Chem., 1971, 31, 175 CrossRef CAS.
  52. J. S. Easow and T. Selvaraju, Electrochim. Acta, 2013, 112, 648 CrossRef CAS PubMed.
  53. S. M. Chen, J. Electroanal. Chem., 1998, 457, 23 CrossRef CAS.
  54. K. Tedsree, T. Li, S. Jones, C. W. A. Chan, K. M. K. Yu, P. A. J. Bagot, E. A. Marquis, G. D. W. Smith and S. C. E. Tsang, Nat. Nanotechnol., 2011, 6, 302 CrossRef CAS PubMed.
  55. W. J. R. Santos, A. L. Sousa, C. S. R. Luz, F. S. Damos, L. T. Kubota, A. A. Tanaka and S. M. C. N. Tanaka, Talanta, 2006, 70, 588 CrossRef CAS PubMed.
  56. M. H. Pournaghi-Azar and H. Dastangoo, J. Electroanal. Chem., 2004, 567, 211 CrossRef CAS PubMed.
  57. X. Chen, F. Wang and Z. Chen, Anal. Chim. Acta, 2008, 623, 213 CrossRef CAS PubMed.
  58. Y. Cui, C. Yang, W. Zeng, M. Oyama, W. Pu and J. Zhang, Anal. Sci., 2007, 23, 1421 CrossRef CAS.

This journal is © The Royal Society of Chemistry 2015
Click here to see how this site uses Cookies. View our privacy policy here.