Open Access Article
This Open Access Article is licensed under a
Creative Commons Attribution 3.0 Unported Licence

One-pot sonochemical synthesis of reduced graphene oxide uniformly decorated with ultrafine silver nanoparticles for non-enzymatic detection of H2O2 and optical detection of mercury ions

A. Moradi Golsheikh *ab, N. M. Huang *a, H. N. Lim c and Rozalina Zakaria d
aLow Dimensional Materials Research Centre (LDMRC), Physics Department, Faculty of Science, University of Malaya, 50603 Kuala Lumpur, Malaysia. E-mail: huangnayming@um.edu.my; amir_moradi_g@ymail.com; Fax: +60 379674146; Tel: +60 122091008
bNanotechnology Laboratory, Esfarayen University of Technology, Esfarayen 9661998195, North Khorasan, Iran
cDepartment of Chemistry, Faculty of Science, University Putra Malaysia, 43400 UPM Serdang, Selangor Darul Ehsan, Malaysia
dPhotonics Research Centre, Physics Department, Faculty of Science, University of Malaya, 50603 Kuala Lumpur, Malaysia

Received 19th June 2014 , Accepted 6th August 2014

First published on 7th August 2014


Abstract

Reduced graphene oxide (rGO) uniformly decorated with silver nanoparticles (AgNPs) was synthesized through the simple ultrasonic irradiation of an aqueous solution containing a silver ammonia complex (Ag(NH3)2OH) and graphene oxide (GO). The results of X-ray diffraction, Fourier-transform infrared transmission spectroscopy, X-ray photoelectron spectroscopy and Raman spectroscopy confirmed the simultaneous formation of cubic-phase AgNPs and the reduction of GO through an ultrasonication process. The size of the nanoparticles could be tuned by adjusting the volume ratio of the precursors and the ultrasonic irradiation time. Transmission electron microscopy images showed a uniform distribution of ultrafine spherical AgNPs with a narrow size distribution on the rGO sheets, which could only be achieved using the silver ammonia complex, rather than silver nitrate, as the precursor. The average particle size of the silver with the narrowest size distribution was 4.57 nm. The prepared AgNPs–rGO modified glassy carbon electrode exhibited notable electrocatalytic activity toward the non-enzymatic detection of H2O2 with a wide linear range of 0.1–70 mM (R2 = 0.9984) and a detection limit of 4.3 μM. Furthermore, the prepared AgNPs–rGO composite was employed for the spectral detection of Hg2+ ions and showed a detection limit of 20 nM.


1. Introduction

Graphene, a one-atom-thick sheet of sp2-bonded carbon atoms, has unique electronic, mechanical and thermal properties.1–3 Thus, it has potential applications in many advanced technologies, including nanoelectronics, sensors, capacitors and composites.4–7 At present, graphene nanosheets are prepared using a variety of techniques, including the micromechanical exfoliation of graphite, chemical vapour deposition, electrochemical reduction of graphene oxide (GO), epitaxial growth and thermal or chemical reduction of graphite oxide.1,8–12 Recently, a sonochemical method was reported as a facile, fast and cost-effective route for the reduction of GO to form graphene nanosheets.13

Graphene's favourable characteristics, including its high electrical conductivity, large surface-to-volume ratio and excellent chemical tolerance, render it a distinguishable matrix for composites. In view of this, metal-nanoparticle-decorated graphene composites have been the focus of research in recent years because of their multifunctional abilities. Among these, silver-nanoparticle (AgNP)-decorated graphene composites are consistently some of the most frequently researched composites because they are effective in various applications, including surface-enhanced Raman scattering (SERS) substrates, glucose sensors and hydrogen peroxide sensors.14–16 These AgNP-decorated graphene composites are prepared using different strategies: (1) the in situ reduction of silver ions on the surface of GO and simultaneous reduction of GO to graphene,16,17 (2) in situ reduction of silver ions on the surface of pre-formed reduced GO18,19 and (3) decoration of pre-formed reduced GO with pre-synthesized AgNPs.15,20 The latter two strategies involve multiple steps, which make the preparation process complex. In addition, AgNP-decorated graphene composites normally require chemical reducing and/or stabilizing agents (such as ethylene glycol,21 hydrazine,22 and poly(diallydimethylammonium chloride) (PDDA)23), some of which are toxic.

In recent years, a sonochemical method has proven to be a versatile and promising technique for the synthesis of a variety of nanostructures such as metals,24–26 metal oxides27,28 and metal sulphides.28,29 The chemical effects of ultrasound irradiation arise from the acoustic cavitation phenomenon. When a liquid is irradiated with ultrasound, bubbles are created, which allow the ultrasonic energy to accumulate during growth. The unstable gas–liquid interface subsequently collapses, releasing the stored energy within a very short time. These cavitational implosions generate localized hotspots with a high temperature of 5000 K, pressure of 1000 bar and heating and cooling rate of 1010 K s−1.30 These extreme conditions are suitable for reducing GO and metal ions. Recently, a sonochemical strategy has been used for the synthesis of graphene-based inorganic nanocomposites such as metals,31,32 bimetals33 and metal oxides.34–36 Herein, we report a fast, one-step, cost-effective and environmentally friendly synthesis of rGO uniformly decorated with ultrafine AgNPs using a sonochemical method that does not involve chemical reducing and/or stabilizing agents.

2. Experimental

2.1 Materials

Graphite flakes were purchased from Ashbury Inc. (NJ, USA). Sulphuric acid (H2SO4, 98%), potassium permanganate (KMnO4, 99.9%), hydrogen peroxide (H2O2, 30%), hydrochloric acid (HCl, 37%) and sodium hydroxide (NaOH, 99.99%) were purchased from Merck. Silver nitrate (AgNO3, 99.7%) was purchased from Systerm, Malaysia. An ammonia solution (NH3, 25%) was obtained from Sigma-Aldrich. Distilled water was used throughout the sample preparation.

2.2 Preparation of AgNPs–rGO composite

GO was prepared using a simplified Hummers' method.37 A silver–ammonia [Ag(NH3)2OH] solution was prepared by adding ammonia (1 w/v%) to a silver nitrate solution (50 mM) until the complete absence of precipitates was observed. The concentration of the obtained Ag(NH3)2OH was approximately 40 mM. The freshly prepared Ag(NH3)2OH solution was mixed with an aqueous solution of GO (1.0 mg mL−1) at GO-to-Ag(NH3)2OH volume ratios of 8, 4 and 2 (the samples were labelled AgNPs–rGO-1, AgNPs–rGO-2 and AgNPs–rGO-3, respectively) and stirred for 5 min to ensure homogeneity. The resulting solutions were exposed to acoustic cavitation by using an ultrasonic horn (Misonix Sonicator S-4000, USA, 20 kHz) immersed directly into the solutions at a 60% amplitude for 5 min, with a 3 s pulse and 5 s relaxation cycle. The products were centrifuged and washed with distilled water three times and were finally redispersed in water. To investigate the effect of the ultrasonic irradiation time, AgNPs–rGO-4 and AgNPs–rGO-5 were prepared by using ultrasonic irradiation times of 15 and 30 min, respectively, for the same solution as AgNPs–rGO-2. The ultrasonic reaction was conducted without cooling. Thus, temperatures of approximately 65, 75 and 80 °C were achieved at the end of the 5, 15 and 30 min reaction times, respectively. For comparison, AgNPs–rGO-6 was prepared under the same conditions as AgNPs–rGO-2 using an AgNO3 solution (0.04 M) instead of the Ag(NH3)2OH solution (0.04 M).

2.3 Preparation of modified electrode

A glassy carbon electrode (GCE) was polished using 1.0- and 0.3 μm alumina powders and then sonicated in ethanol and distilled water for 1 min each. Finally, the electrode was dried in a stream of nitrogen, and 5 μL of a suspension of the AgNPs/rGO composite was dropped onto the pre-cleaned GCE and allowed to dry in air at room temperature. An electrochemical sensing application of AgNPs–rGO was conducted on a potentiostat/galvanostat (Versastat 3 Applied Research Princeton, USA) using a three-electrode system, with a modified GCE (with an active area of 0.07 cm2) as the working electrode, a platinum wire as the counter electrode and a saturated calomel electrode (SCE) as the reference electrode.

2.4 Spectral detection of Hg2+ ions

The spectral detection of Hg2+ ions using the AgNPs/rGO composite was performed using a UV-visible spectrometer (Thermo Scientific Evolution). The absorption spectra of the AgNPs/rGO composite were recorded after adding different concentrations of Hg2+ ions from 0.1 to 100 μM. To monitor the absorption spectra, a freshly prepared solution of Hg2+ ions was added to 4 mL of the AgNPs/rGO composite, shaken well, and allowed a constant resting time (8 min). For selective detection, 100 μM of the analyte (Hg2+, Fe2+, K+, Mn2+, Pb2+, Zn2+, Cd2+, Cu2+ and Ni2+ ions) was added to the AgNPs/rGO composite, and the changes in the absorption spectra were recorded. The limits of detection (LODs) for both hydrogen peroxide and mercury ions were calculated using the following equation:
 
LOD = 3S/b(1)
where S is the standard deviation of the blank, and b is the slope of the calibrated curve.

2.5 Characterization

The crystal phase, morphology and microstructure of the samples were characterized using the X-ray powder diffraction (XRD; Philips X'pert system using Cu Kα radiation), a UV-visible spectrometer (Thermo Scientific Evolution), a high-resolution transmission electron microscope (HRTEM, JEOL JEM-2100F), a Fourier-transform infrared spectrometer (FTIR; Perkin Elmer System 2000 series spectrophotometer, USA), X-ray photoelectron spectroscopy (XPS) (measurements were taken using synchrotron light at beamline no. 3.2 of the Siam Photon Laboratory in the Synchrotron Light Research Institute, Thailand) and a Raman spectrometer (Renishaw inVia Raman microscope using laser excitation at λ = 514 nm).

3. Results and discussion

Fig. 1 shows the XRD patterns of pristine GO (a), a sample holder (b) and AgNPs–rGO prepared at different reaction times (c–e) (5, 15 and 30 min, respectively). The pristine GO has a sharp peak at 10.8°, which is assigned to the (002) inter-planar spacing of 0.82 nm.38 Meanwhile, the XRD pattern of the sample holder has a broad peak at about 2θ = 13°, which is observed in the XRD patterns of all the samples. As shown in Fig. 1(c)–(e), the XRD patterns of AgNPs–rGO composites exhibit broad peaks at 38.1°, 44.3°, 64.2° and 77.6°, which can be indexed to the (111), (200), (220) and (311) planes of the cubic Ag crystal, respectively (PDF card no: 00-004-0783). The calculated lattice spaces of the (111), (200), (220) and (311) planes are 2.36, 2.05, 1.44 and 1.23 Å, respectively. The average crystallite sizes of the AgNPs–rGO composites prepared at 5, 15 and 30 min, calculated using the size strain plot method,39 are 5.6, 6.6 and 8.3 nm, respectively. As the ultrasonic irradiation time increases, the intensity of the peaks and the average crystallite size increase, indicating the formation of AgNPs with a larger average size. Meanwhile, the (002) peak of GO disappeared after the sonication process. This is attributed to the growth of AgNPs on the surface of the graphene sheets, which prevented the restacking of the graphene sheets.38
image file: c4ra05998k-f1.tif
Fig. 1 XRD patterns of pristine GO (a), sample holder (b) and AgNPs–rGO prepared at different ultrasonic irradiation times: 5 min (c), 15 min (d) and 30 min (e).

Fig. 2A shows the UV-Vis absorption spectra of GO and AgNPs–rGO. In the UV-Vis spectrum of GO, two characteristic peaks can be observed at 229 and 304 nm, which are attributed respectively to π–π* transitions of aromatic C[double bond, length as m-dash]C bonds and n–π* transitions of C[double bond, length as m-dash]O bonds.40 After ultrasonic irradiation of the Ag(NH3)2OH and GO solution, those peaks disappeared, and two new peaks appeared at 251 and 396 nm, which correspond to the reduction in GO and surface plasmon resonance (SPR) of the spherical AgNPs, respectively.41,42 The inset of Fig. 2A shows a photograph of the solution of Ag(NH3)2OH and GO before and after ultrasonic irradiation treatment, in which the colour changes from yellow brown to black. Fig. 2B(a)–(d) shows the effect of the ultrasonic irradiation time on the formation of the AgNPs–rGO composites. After 5 min of ultrasonic irradiation, a slight absorption band at 396 nm appears. Meanwhile, the peak can be clearly observed at 15 min, indicating the presence of spherical AgNPs. A further increase in the reaction time induced a heightened peak intensity at 396 nm and the “red enlargement” of this peak. The enhanced peak intensity and broadness imply the increased concentration of silver, which is attributed to the increased amount and size of the AgNPs, as proven in the subsequent discussion on the morphology of the composite observed using the HRTEM.


image file: c4ra05998k-f2.tif
Fig. 2 (A) UV-Vis absorption spectra of GO and AgNPs–rGO (inset shows photograph of solution of GO and Ag(NH3)2OH before and after ultrasonic irradiation) and (B) time evolution of UV-Vis absorption spectra of AgNPs–rGO.

Fig. 3 shows the FTIR spectra of pristine GO (a) and AgNPs–rGO composites that were prepared at different ultrasonic irradiation times (b–d) (5, 15 and 30 min, respectively). For GO, the broad peak centred at 3227 cm−1 is attributed to the O–H stretching vibrations, whereas the peaks at 1735, 1626, 1381 and 1225 cm−1 are assigned to the C[double bond, length as m-dash]O stretching, sp2-hybridized C[double bond, length as m-dash]C group and O–H bending, C–OH stretching and C–O–C stretching, respectively.43 Meanwhile, the peaks at 1169 and 1046 cm−1 can be attributed to the C–O vibration of the epoxy or alkoxy groups.44 For the AgNPs–rGO composites, the peak at 1597 cm−1 is assigned to the sp2-hybridized C[double bond, length as m-dash]C group. The peak at 1169 cm−1 for GO is absent for the AgNPs–rGO composites, and the intensity of the peaks, related to the oxygen-containing groups, decreases for the AgNPs–rGO composites, indicating the reduction of GO during the ultrasonic irradiation process. In addition, Fig. 3 shows that the peak intensity of the oxygen-containing groups decreases gradually with an increase in the ultrasonic irradiation time, which indicates the stepwise removal of oxygen-containing groups.


image file: c4ra05998k-f3.tif
Fig. 3 FTIR spectra of pristine GO (a) and AgNPs–rGO prepared at different ultrasonic irradiation times: 5 min (b), 15 min (c) and 30 min (d).

The XPS spectra further proved the reduction of GO after ultrasonic treatment. As shown in Fig. 4a, the C1 band of pristine GO can be fitted to four deconvulated components, centred at 284.5, 286.2, 287.8 and 289 eV, which are assigned to the non-oxygenated ring C, the C in C–O, the C in C[double bond, length as m-dash]O and the C in C(O)O, respectively.45,46 In comparison to the pristine GO, the C1 band of the AgNPs–rGO composites shows that the peak intensity of the oxygenated carbonaceous bands decreased, indicating the reduction of GO after ultrasonic treatment. In addition, the peak intensity of the oxygenated carbonaceous bands gradually decreases with an increase in the ultrasonic irradiation time, which is consistent with the FTIR results.


image file: c4ra05998k-f4.tif
Fig. 4 XPS spectra of pristine GO (a) and AgNPs–rGO prepared at different ultrasonic irradiation times: 5 min (b), 15 min (c) and 30 min (d).

Fig. 5 shows the Raman spectra for pristine GO (a) and the AgNPs–rGO composites with different ultrasonic irradiation times (b–d) (5, 15 and 30 min, respectively). The Raman spectrum of GO shows two peaks at 1350 and 1603 cm−1, which correspond to the well-known D and G bands, respectively. The D band is ascribed to the breathing mode of A1g symmetry, involving phonons near the K zone boundary. Meanwhile, the G band is assigned to the E2g mode of sp2-bonded carbon atoms.47 The D and G band positions and intensity ratios of I(D)/I(G) for the GO and AgNPs–rGO composites, prepared at different ultrasound irradiation times, are summarized in Table 1. In comparison to the pristine GO, the Raman spectra of the AgNPs–rGO composites show that the G band shifted to lower wavenumbers, which can be attributed to the reduction in GO during the ultrasonic irradiation process.48,49 In addition, the intensity ratio of I(D)/I(G) for an AgNPs–rGO composite decreases with an increase in the ultrasonic irradiation time to 15 min compared to that of GO, which can be attributed to the formation of sp2 bonds during the ultrasonic irradiation process.50 However, when the ultrasonic irradiation time was further increased to 30 min, the intensity ratio of I(D)/I(G) increased to a slightly higher value of 0.95, compared to that of GO (0.91), which can be assigned to the formation of further defects in the rGO sheets. A possible explanation is that the prolonged reaction time is conducive to the increased formation of AgNPs on the rGO, thus overriding the sp2 hybridized carbon. Moreover, the intensity peaks of the D and G bands were enhanced in the case of AgNPs–rGO as a result of the SERS effect of the AgNPs. It is well-known that electromagnetic and chemical enhancements are the origins of the SERS effect. Electromagnetic enhancement can be associated with the excitation of localized surface plasmon resonances (LSPR) of noble metal nanoparticles (such as Ag and Au), with a large enhancement factor (up to 1012), whereas chemical enhancement can be attributed to the formation of charge-transfer complexes through chemical interaction, with a minor enhancement factor (up to 100).51,52 As shown in Fig. 5, the intensity peaks of the D and G bands for the AgNPs–rGO composite increased from 1.3 to 4 times compared to that of GO with the increase in the ultrasonic irradiation time, which indicates the increase in AgNPs on the surface of the graphene nanosheets as the ultrasonic irradiation time increased.52


image file: c4ra05998k-f5.tif
Fig. 5 Raman spectra of pristine GO (a) and AgNPs–rGO prepared at different ultrasonic irradiation times: 5 min (b), 15 min (c) and 30 min (d).
Table 1 D and G peak positions and intensity ratios of I(D)/I(G) (obtained by Raman analysis) of GO and AgNPs–rGO composites prepared at different ultrasonic irradiation times
  GO AgNPs–rGO (5 min) AgNPs–rGO (15 min) AgNPs–rGO (30 min)
D band (cm−1) 1350 1360 1363 1360
G band (cm−1) 1603 1594 1593 1591
I(D)/I(G) 0.91 0.88 0.87 0.95


Fig. 6 shows the typical TEM images and size distribution histograms of the AgNPs–rGO prepared by using solutions with GO (1.0 mg mL−1) to Ag(NH3)2OH (0.04 M) volume ratios of 8 (a and b), 4 (c and d) and 2 (e and f), as well as a solution with a GO (1.0 mg mL−1) to AgNO3 (0.04 M) volume ratio of 4 (g and h) at an ultrasonic irradiation time of 5 min. As shown in Fig. 6(a) and (b), ultrafine spherical AgNPs with a mean size of 4.56 nm and a narrow size distribution are anchored and well distributed on the surface of the rGO sheets. Fig. 6(c)–(f) shows that the increased amount of Ag(NH3)2OH is commensurate with the mean size and size distribution. The type of Ag precursor appears to affect the size and coverage density of the nanoparticles on the surface of the graphene. When AgNO3 was used as a precursor of Ag, the mean particle size increased to 9.96 nm, and the size distribution widened, as shown in Fig. 6(g) and (h). Moreover, the density of the particles was reduced tremendously compared to those prepared by using Ag(NH3)2OH as the precursor of Ag, as a result of the clumping of nanoparticles. This implies that the ammonia molecules acted somewhat as a template for the nucleation and growth of AgNPs at a constraint geometry. The ammonia molecules critically controlled the particle size and size distribution of the AgNPs.


image file: c4ra05998k-f6.tif
Fig. 6 TEM images and size distribution diagrams of AgNPs–rGO prepared using solutions with GO (1.0 mg mL−1) to Ag(NH3)2OH (0.04 M) volume ratios of 8 (a and b), 4 (c and d) and 2 (e and f), as well as solution with GO (1.0 mg mL−1) to AgNO3 (0.04 M) volume ratio of 4 (g and h), with same ultrasonic irradiation time of 5 min.

Fig. 7 shows typical TEM images and size distribution diagrams of the AgNPs–rGO prepared by using the solution with the GO (1.0 mg mL−1) to Ag(NH3)2OH (0.04 M) volume ratio of 4 at different ultrasonic irradiation times of 15 min (a and b) and 30 min (c and d). As shown in Fig. 7, as the ultrasonic irradiation time increases from 5 min to 15 and 30 min, the mean size increases from 5.67 nm to 7.82 and 12.68 nm, respectively, and the size distribution broadens. Fig. 7(e) shows an HRTEM image of AgNPs anchored on the surface of the rGO sheet. The measured lattice fringe spaces of 0.23 and 0.20 nm are attributed to the (111) and (200) planes of silver crystal, respectively. In retrospect, the planes identified by TEM agree with those deduced from XRD.


image file: c4ra05998k-f7.tif
Fig. 7 TEM images and size distribution diagrams of AgNPs–rGO prepared by using solution with GO (1.0 mg mL−1) to Ag(NH3)2OH (0.04 M) volume ratio of 4 at different ultrasonic irradiation times of 15 min (a and b) and 30 min (c and d), and HRTEM image of AgNPs anchored on surface of graphene sheet (e).

The formation mechanism of the AgNPs–rGO via a sonochemical reaction is shown in Fig. 8. Dispersed GO sheets in water are negatively charged as a result of the ionization of the carboxyl and hydroxyl groups on the surface of the GO (Fig. 8(a)).53 This causes the positively charged [Ag(NH3)2]+ ions to be adsorbed on the negatively charged GO sheets by electrostatic attraction (Fig. 8(b)). When an aqueous solution is irradiated with ultrasound, bubbles are created and accumulate ultrasonic energy while growing. Rapid collapse of the bubbles leads to the creation of localized hotspots with a high temperature of 5000 K, pressure of 1000 bar and heating and cooling rate of 1010 K s−1. Under these conditions, highly reactive H and OH radicals can be generated by the pyrolysis of water molecules.25

 
H2O)))))) H˙ + OH˙(2)


image file: c4ra05998k-f8.tif
Fig. 8 Schematic illustration of formation mechanism for AgNPs–rGO composite via ultrasonic irradiation.

The generated H˙ radicals act as reducing species, resulting in reduction mechanisms:25,31

 
Ag+ + H˙ → Ag0 + H+(3)
or
 
2[Ag(NH3)2]+ + H˙ → 2Ag0 + (NH4)+ + NH3(4)
and
 
GO + reducing species of H˙ → rGO(5)

The GO and adsorbed [Ag(NH3)2]+ or Ag+ ions are simultaneously reduced to graphene and AgNPs by the ultrasonic irradiation process (Fig. 8(c)). In addition, GO is a thermally unstable material above 200 °C. Hence, the high temperature generated during ultrasonic irradiation could reduce GO to graphene.13,54 When the concentration of the [Ag(NH3)2]+ ions or the ultrasonic irradiation time increased, the free [Ag(NH3)2]+ or Ag+ ions in the aqueous solution continued to grow on the previously formed AgNPs, as well as nucleating on the rGO sheets, which explained the increased particle size and density on the rGO sheets. As can be seen in Fig. 6, the sonochemical reaction can be enhanced by using the silver ammonia complex instead of silver nitrate. The two plausible reasons are given as follows: (1) the negatively charged GO is enhanced through neutralization by the alkaline Ag(NH3)2OH to attract more [Ag(NH3)2]+ ions than AgNO3, leading to the creation of more initial nucleation sites;17 and (2) the ammonia can scavenge the OH radicals that are produced from pyrolysis to provide a reducing condition.55

AgNPs are typically known to exhibit high catalytic activity for the reduction of hydrogen peroxide. To investigate the effect of the ultrasonication time on the electrocatalytic activity of the AgNPs–rGO towards hydrogen peroxide reduction, cyclic voltammetry (CV) measurements of the AgNPs–rGO/GCE electrodes prepared at different ultrasonication times were conducted in a 0.2 M phosphate buffer solution (PBS) at pH 6.5 in the presence of 1 mM H2O2. As shown in Fig. 9, all the AgNPs–rGO/GCE electrodes exhibited a notable cathodic peak for the reduction of H2O2 in comparison to the bare GCE. As can be seen in Fig. 9, the catalytic activity of the AgNPs–rGO/GCE electrode was enhanced by an increase in the ultrasonication time used to prepare the composite from 5 min to 15 min. A plausible reason is that a higher-density of AgNPs could be achieved on the surface of rGO by increasing the ultrasonication time, which enhanced the electrocatalytic activity of the composite. In addition, a further increase in the ultrasonication time to 30 min decreased the electrocatalytic activity of the composite towards the reduction of H2O2. A plausible reason is that, although a higher density of AgNPs could be achieved on the surface of the rGO, it consisted of larger AgNPs. These results are consistent with those of a previous study,56 which implied that the electrocatalytic activity of silver nanoparticles decreased with increasing particle size and increased with increasing particle density. The reduction peak of all the AgNPs–rGO/GCE electrodes shifted to a positive potential as the size and density of the AgNPs increased.33,57


image file: c4ra05998k-f9.tif
Fig. 9 CV values of various electrodes in 0.2 M PBS (pH 6.5) in presence of 1.0 mM H2O2: (a) bare GCE and AgNPs–rGO/GCE prepared by using different ultrasonication times of (b) 5 min, (c) 15 min and (d) 30 min.

Fig. 10 shows the amperometric current–time response of the AgNPs–rGO-4/GCE electrode at −0.4 V in an N2-saturated 0.2 M PBS buffer (pH: 6.5) upon successive step changes in the H2O2 concentration. When an aliquot of H2O2 was injected in the stirring PBS solution, the reduction current rapidly changed to reach a steady-state value within 3 s, indicating a fast amperometric response behaviour. The inset shows the corresponding calibration curve of the AgNPs–rGO-4/GCE electrode. The current response of the AgNPs–rGO-4/GCE electrode is estimated to be linear within the H2O2 concentration range of 0.1 to 70 mM (R2 = 0.9984), whereas the limit of detection is estimated to be 4.3 μM, based on a signal-to-noise ratio of three. The relative standard deviation (RSD) of the current response to 5 mM H2O2 at −0.4 V is 5.6% for five successive measurements. Based on the results of a comparative analysis of different types of electrodes listed in Table 2, the present AgNPs–rGO/GCE electrode is demonstrated to be capable of affording a favourable detection limit and linear range for sensing H2O2.


image file: c4ra05998k-f10.tif
Fig. 10 Steady-state response of AgNPs–rGO-4/GCE to successive injections of H2O2 into stirred 0.2 M PBS (pH 6.5) with applied potential of −0.4 V. The inset is the corresponding calibration curve.
Table 2 Comparison of results from this work and literature regarding performance of H2O2 assays
Type of electrode Limit of detection (μM) Linear range (mM) References
Ag NPs–MWCNT/Au electrode 0.5 0.05–17 58
Ag NPs–GN–R/GCE 28 0.1–40 22
AgNP–PMPD/GCE 4.7 0.1–30 59
MnO2NWs–Ag/GCE 0.24 0.1–4 60
AgNP/GO/ssDNA/AuE 1.9 0.1–20 61
GN–AgNWs/GCE 9.3 0.02–5.34 32
Ag NPs–NFs/GCE 62 0.1–80 62
AgNPs–rGO/ITO 5 0.1–100 16
AgNP/rGO–benzylamine/GCE 31.3 0.1–100 15
PDDA–rGO/AgNPs/GCE 35 0.1–41 23
Ag NPs/3DG 14.9 0.03–16.21 63
AgNPs/rGO/GCE 4.3 0.1–70 This work


In order to evaluate the sensitivity and application of the AgNPs–rGO composite to the determination of Hg2+ ions, the UV-vis absorbance peak related to the surface plasmon resonance (SPR) of silver nanoparticles was monitored during the addition of different concentrations of an aqueous solution of Hg2+ ions at room temperature. As shown in the inset of Fig. 11a, the colour of AgNPs/rGO gradually changes from brown-yellow to colourless with the increase in the Hg2+ ion concentrations. Fig. 11a shows the UV-vis absorbance values for the AgNPs–rGO composite in the absence and presence of different concentrations of Hg2+ ions, where the intensity of the absorbance peak decreases with an increasing concentration of Hg2+ ions. Accompanied by the extinction intensity of the peak, the SPR peak position shifts toward shorter wavelengths. This phenomenon can be explained by the redox reaction between the Hg2+ ions and Ag atoms. Because the redox potential of the Hg2+/Hg0 (0.85 V) is higher than that of the Ag+/Ag0 (0.8 V), Hg2+ ions oxidize the metallic Ag atoms, which produces metallic Hg atoms and Ag+ ions. According to the previous reports, the blue shift in the SPR peak can be related to the adsorption of Hg onto Ag NPs.64,65Fig. 11b shows a calibration curve related to the difference in the absorbance peak of the AgNPs–rGO composite, which is subject to the concentration of Hg2+ ions. The intensity of the absorbance peak rapidly decreases with an increase in the concentration of Hg2+ ions up to 10 μM. Above this concentration, the intensity of the absorbance peak slowly decreases with good linearity. A plausible reason for the two linear ranges is that with the increase in the Hg2+ ion concentration above 10 μM, the formation of a metallic Hg layer around the Ag nanoparticles may occur, which decreases the sensitivity and produces the second linear range.65 The estimated limit of detection is 20 nM, based on a signal-to-noise ratio of three. Further experiments were carried out to investigate the selective sensing of Hg2+ ions by AgNPs–rGO in the presence of other environmentally relevant metal ions (Fe2+, K+, Mn2+, Pb2+, Zn2+, Cd2+, Cu2+ and Ni2+). As shown in Fig. 11c, the SPR peak of AgNPs/rGO was dramatically quenched only by the addition of Hg2+, and no significant changes in the SPR peak were observed by the addition of other metal ions, which demonstrated the selective sensing ability of the prepared AgNPs/rGO composite toward Hg2+ ions.


image file: c4ra05998k-f11.tif
Fig. 11 (a) Absorbance responses of AgNPs–rGO composite for different concentrations of Hg2+ ions from 0.1 to 100 μM. The inset shows the gradual colour change of the AgNPs–rGO solution with the increase in the Hg2+ ion concentration. (b) The corresponding calibration curve for the Hg2+ ion detection. (c) The relative absorbance change in AgNPs–rGO in the presence of Hg2+ ions and some other common metal ions.

4. Conclusion

The ultrasonic irradiation of an aqueous solution containing a silver ammonia complex (Ag(NH3)2OH) and graphene oxide (GO) is a simple strategy for producing rGO uniformly decorated with ultrafine AgNPs with a narrow size distribution. It was established that the silver ammonia complex is the key to the successful synthesis of well-formed composites. The particle size of the AgNPs embedded on the rGO sheets could be easily tuned by adjusting the ultrasonic radiation time or the amount of Ag(NH3)2OH. The prepared AgNPs–rGO composite exhibited a promising potential application to the enzyme-less electrochemical detection of hydrogen peroxide and the SPR optical sensing of mercury ions.

Acknowledgements

This work was supported by the UMRG Program Grant (RP007C/13AFR) by University of Malaya and a High Impact Research Grant from the Ministry of Higher Education of Malaysia (UM.C/625/1/HIR/MOHE/SC/21).

References

  1. K. S. Novoselov, A. K. Geim, S. V. Morozov, D. Jiang, Y. Zhang, S. V. Dubonos, I. V. Grigorieva and A. A. Firsov, Science, 2004, 306, 666–669 CrossRef CAS PubMed.
  2. C. Lee, X. Wei, J. W. Kysar and J. Hone, Science, 2008, 321, 385–388 CrossRef CAS PubMed.
  3. A. A. Balandin, S. Ghosh, W. Bao, I. Calizo, D. Teweldebrhan, F. Miao and C. N. Lau, Nano Lett., 2008, 8, 902–907 CrossRef CAS PubMed.
  4. S. Gilje, S. Han, M. Wang, K. L. Wang and R. B. Kaner, Nano Lett., 2007, 7, 3394–3398 CrossRef CAS PubMed.
  5. M. Zhou, Y. Zhai and S. Dong, Anal. Chem., 2009, 81, 5603–5613 CrossRef CAS PubMed.
  6. C. Liu, Z. Yu, D. Neff, A. Zhamu and B. Z. Jang, Nano Lett., 2010, 10, 4863–4868 CrossRef CAS PubMed.
  7. X. Huang, X. Qi, F. Boey and H. Zhang, Chem. Soc. Rev., 2012, 41, 666–686 RSC.
  8. A. Guermoune, T. Chari, F. Popescu, S. S. Sabri, J. Guillemette, H. S. Skulason, T. Szkopek and M. Siaj, Carbon, 2011, 49, 4204–4210 CrossRef CAS PubMed.
  9. H.-L. Guo, X.-F. Wang, Q.-Y. Qian, F.-B. Wang and X.-H. Xia, ACS Nano, 2009, 3, 2653–2659 CrossRef CAS PubMed.
  10. C. Berger, Z. Song, X. Li, X. Wu, N. Brown, C. Naud, D. Mayou, T. Li, J. Hass, A. N. Marchenkov, E. H. Conrad, P. N. First and W. A. de Heer, Science, 2006, 312, 1191–1196 CrossRef CAS PubMed.
  11. H. C. Schniepp, J.-L. Li, M. J. McAllister, H. Sai, M. Herrera-Alonso, D. H. Adamson, R. K. Prud'homme, R. Car, D. A. Saville and I. A. Aksay, J. Phys. Chem. B, 2006, 110, 8535–8539 CrossRef CAS PubMed.
  12. S. Stankovich, D. A. Dikin, R. D. Piner, K. A. Kohlhaas, A. Kleinhammes, Y. Jia, Y. Wu, S. T. Nguyen and R. S. Ruoff, Carbon, 2007, 45, 1558–1565 CrossRef CAS PubMed.
  13. K. Krishnamoorthy, G.-S. Kim and S. J. Kim, Ultrason. Sonochem., 2013, 20, 644–649 CrossRef CAS PubMed.
  14. Y. Zhang, S. Liu, L. Wang, X. Qin, J. Tian, W. Lu, G. Chang and X. Sun, RSC Adv., 2012, 2, 538–545 RSC.
  15. S. Liu, J. Tian, L. Wang and X. Sun, Carbon, 2011, 49, 3158–3164 CrossRef CAS PubMed.
  16. A. Moradi Golsheikh, N. M. Huang, H. N. Lim, R. Zakaria and C.-Y. Yin, Carbon, 2013, 62, 405–412 CrossRef CAS PubMed.
  17. Y. Zhang, X. Yuan, Y. Wang and Y. Chen, J. Mater. Chem., 2012, 22, 7245–7251 RSC.
  18. J. Li and C.-y. Liu, Eur. J. Inorg. Chem., 2010, 2010, 1244–1248 CrossRef.
  19. J. Ma, J. Zhang, Z. Xiong, Y. Yong and X. S. Zhao, J. Mater. Chem., 2011, 21, 3350–3352 RSC.
  20. J. Liu, S. Fu, B. Yuan, Y. Li and Z. Deng, J. Am. Chem. Soc., 2010, 132, 7279–7281 CrossRef CAS PubMed.
  21. B. Yu, J. Feng, S. Liu and T. Zhang, RSC Adv., 2013, 3, 14303–14307 RSC.
  22. S. Liu, J. Tian, L. Wang, H. Li, Y. Zhang and X. Sun, Macromolecules, 2010, 43, 10078–10083 CrossRef CAS.
  23. X. Liu, X. Xu, H. Zhu and X. Yang, Anal. Methods, 2013, 5, 2298–2304 RSC.
  24. S. K. Tripathy, A. Mishra, S. K. Jha, R. Wahab and A. A. Al-Khedhairy, Anal. Methods, 2013, 5, 1456–1462 RSC.
  25. V. G. Pol, D. N. Srivastava, O. Palchik, V. Palchik, M. A. Slifkin, A. M. Weiss and A. Gedanken, Langmuir, 2002, 18, 3352–3357 CrossRef CAS.
  26. M. Darroudi, A. Khorsand Zak, M. R. Muhamad, N. M. Huang and M. Hakimi, Mater. Lett., 2012, 66, 117–120 CrossRef CAS PubMed.
  27. A. Khorsand Zak, W. H. a. Majid, H. Z. Wang, R. Yousefi, A. Moradi Golsheikh and Z. F. Ren, Ultrason. Sonochem., 2013, 20, 395–400 CrossRef CAS PubMed.
  28. J. Geng, X.-D. Jia and J.-J. Zhu, CrystEngComm, 2011, 13, 193–198 RSC.
  29. Y. Wu, X. Hao, J. Yang, F. Tian and M. Jiang, Mater. Lett., 2006, 60, 2764–2766 CrossRef CAS PubMed.
  30. J. H. Bang and K. S. Suslick, Adv. Mater., 2010, 22, 1039–1059 CrossRef CAS PubMed.
  31. K. Vinodgopal, B. Neppolian, I. V. Lightcap, F. Grieser, M. Ashokkumar and P. V. Kamat, J. Phys. Chem. Lett., 2010, 1, 1987–1993 CrossRef CAS.
  32. S. Chandra, S. Bag, R. Bhar and P. Pramanik, J. Nanopart. Res., 2011, 13, 2769–2777 CrossRef CAS.
  33. S. Anandan, A. Manivel and M. Ashokkumar, Fuel Cells, 2012, 12, 956–962 CrossRef CAS.
  34. Q.-P. Luo, X.-Y. Yu, B.-X. Lei, H.-Y. Chen, D.-B. Kuang and C.-Y. Su, J. Phys. Chem. C, 2012, 116, 8111–8117 CAS.
  35. G. Park, L. Bartolome, K. G. Lee, S. J. Lee, D. H. Kim and T. J. Park, Nanoscale, 2012, 4, 3879–3885 RSC.
  36. S. Zhu, J. Guo, J. Dong, Z. Cui, T. Lu, C. Zhu, D. Zhang and J. Ma, Ultrason. Sonochem., 2013, 20, 872–880 CrossRef CAS PubMed.
  37. N. M. Huang, H. N. Lim, C. H. Chia, M. A. Yarmo and M. R. Muhamad, Int. J. Nanomed., 2011, 6, 3443–3448 CrossRef CAS PubMed.
  38. G. Xie, J. Cheng, Y. Li, P. Xi, F. Chen, H. Liu, F. Hou, Y. Shi, L. Huang, Z. Xu, D. Bai and Z. Zeng, J. Mater. Chem., 2012, 22, 9308–9314 RSC.
  39. M. A. Tagliente and M. Massaro, Nucl. Instrum. Methods Phys. Res., Sect. B, 2008, 266, 1055–1061 CrossRef CAS PubMed.
  40. J. I. Paredes, S. Villar-Rodil, A. Martínez-Alonso and J. M. D. Tascón, Langmuir, 2008, 24, 10560–10564 CrossRef CAS PubMed.
  41. B. Su, D. Tang, Q. Li, J. Tang and G. Chen, Anal. Chim. Acta, 2011, 692, 116–124 CrossRef CAS PubMed.
  42. A. M. Abdel-Mohsen, R. Hrdina, L. Burgert, G. Krylová, R. M. Abdel-Rahman, A. Krejčová, M. Steinhart and L. Beneš, Carbohydr. Polym., 2012, 89, 411–422 CrossRef CAS PubMed.
  43. C. Cheng, S. Nie, S. Li, H. Peng, H. Yang, L. Ma, S. Sun and C. Zhao, J. Mater. Chem. B, 2013, 1, 265–275 RSC.
  44. V. H. Pham, T. V. Cuong, S. H. Hur, E. Oh, E. J. Kim, E. W. Shin and J. S. Chung, J. Mater. Chem., 2011, 21, 3371–3377 RSC.
  45. D. R. Dreyer, S. Park, C. W. Bielawski and R. S. Ruoff, Chem. Soc. Rev., 2010, 39, 228–240 RSC.
  46. S. Pei, J. Zhao, J. Du, W. Ren and H.-M. Cheng, Carbon, 2010, 48, 4466–4474 CrossRef CAS PubMed.
  47. A. C. Ferrari and J. Robertson, Phys. Rev. B: Condens. Matter Mater. Phys., 2000, 61, 14095–14107 CrossRef CAS.
  48. T. N. Lambert, C. A. Chavez, B. Hernandez-Sanchez, P. Lu, N. S. Bell, A. Ambrosini, T. Friedman, T. J. Boyle, D. R. Wheeler and D. L. Huber, J. Phys. Chem. C, 2009, 113, 19812–19823 CAS.
  49. A. M. Golsheikh, N. M. Huang, H. N. Lim, C. H. Chia, I. Harrison and M. R. Muhamad, Chem. Eng. J., 2013, 218, 276–284 CrossRef CAS PubMed.
  50. M. Shi, J. Shen, H. Ma, Z. Li, X. Lu, N. Li and M. Ye, Colloids Surf., A, 2012, 405, 30–37 CrossRef CAS PubMed.
  51. G. Baker and D. Moore, Anal. Bioanal. Chem., 2005, 382, 1751–1770 CrossRef CAS PubMed.
  52. Z. Xu, H. Gao and H. Guoxin, Carbon, 2011, 49, 4731–4738 CrossRef CAS PubMed.
  53. D. Li, M. B. Muller, S. Gilje, R. B. Kaner and G. G. Wallace, Nat. Nanotechnol., 2008, 3, 101–105 CrossRef CAS PubMed.
  54. K. Krishnamoorthy, M. Veerapandian, R. Mohan and S.-J. Kim, Appl. Phys. A: Mater. Sci. Process., 2012, 106, 501–506 CrossRef CAS.
  55. L. Huang, L. Li, W. Dong, Y. Liu and H. Hou, Environ. Sci. Technol., 2008, 42, 8070–8075 CrossRef CAS.
  56. A. Yu, Q. Wang, J. Yong, P. J. Mahon, F. Malherbe, F. Wang, H. Zhang and J. Wang, Electrochim. Acta, 2012, 74, 111–116 CrossRef CAS PubMed.
  57. F. W. Campbell, S. R. Belding, R. Baron, L. Xiao and R. G. Compton, J. Phys. Chem. C, 2009, 113, 9053–9062 CAS.
  58. W. Zhao, H. Wang, X. Qin, X. Wang, Z. Zhao, Z. Miao, L. Chen, M. Shan, Y. Fang and Q. Chen, Talanta, 2009, 80, 1029–1033 CrossRef CAS PubMed.
  59. J. Tian, H. Li, W. Lu, Y. Luo, L. Wang and X. Sun, Analyst, 2011, 136, 1806–1809 RSC.
  60. Q. Han, P. Ni, Z. Liu, X. Dong, Y. Wang, Z. Li and Z. Liu, Electrochem. Commun., 2014, 38, 110–113 CrossRef CAS PubMed.
  61. W. Lu, G. Chang, Y. Luo, F. Liao and X. Sun, J. Mater. Sci., 2011, 46, 5260–5266 CrossRef CAS.
  62. J. Tian, S. Liu and X. Sun, Langmuir, 2010, 26, 15112–15116 CrossRef CAS PubMed.
  63. B. Zhan, C. Liu, H. Shi, C. Li, L. Wang, W. Huang and X. Dong, Appl. Phys. Lett., 2014, 104, 243704 CrossRef PubMed.
  64. A. Henglein and C. Brancewicz, Chem. Mater., 1997, 9, 2164–2167 CrossRef CAS.
  65. T. Morris, H. Copeland, E. McLinden, S. Wilson and G. Szulczewski, Langmuir, 2002, 18, 7261–7264 CrossRef CAS.

This journal is © The Royal Society of Chemistry 2014
Click here to see how this site uses Cookies. View our privacy policy here.