DOI:
10.1039/C6RA10470C
(Paper)
RSC Adv., 2016,
6, 59907-59918
A highly sensitive non-enzymatic hydrogen peroxide and hydrazine electrochemical sensor based on 3D micro-snowflake architectures of α-Fe2O3†
Received
22nd April 2016
, Accepted 9th June 2016
First published on 10th June 2016
Abstract
In the present work, well crystalline 3D micro-snowflake structured α-Fe2O3 has been successfully synthesized on a large scale via a simple hydrothermal reaction by hydrolysis of a K3Fe(CN)6 precursor. The structure, composition, purity and morphology of the synthesized α-Fe2O3 samples are examined using powder X-ray diffraction (PXRD), field emission scanning electron microscopy (FESEM), transmission electron microscopy (TEM), energy dispersive X-ray spectroscopy (EDS), Fourier transform infrared (FTIR) spectroscopy and Mössbauer spectroscopy. The FESEM and TEM images reveal that the sample exhibits a micro-snowflake like shape having six-fold symmetry with symmetric branching along each arm consisting of a long central trunk and secondary branches. The 3D micro-snowflake structured α-Fe2O3 embedded ITO electrode exhibits high selectivity and sensitivity for electrochemical probing of hydrogen peroxide (H2O2) and hydrazine (N2H4) with a very low detection limit in a wide linear range. Amperometric measurements show a sensitivity of 7.16 μA mM−1 cm−2 in a wide linear range from 0.1 to 5.5 mM with the lowest detection limit of 0.01 mM (S/N = 3) towards H2O2 sensing. The sample also exhibits a sensitivity of 24.03 μA mM−1 cm−2 in the linear range between 50 μM and 1340 μM with the lowest detection limit of 5 μM towards hydrazine detection. The excellent electrochemical activity of the sample is rendered to the presence of a large number of catalytic sites in the sample due to its 3D micro-snowflake like architecture. Good reproducibility, stability and selectivity suggest its suitability for the fabrication of H2O2 and hydrazine sensors.
1. Introduction
Recently, iron based nanoparticles have emerged as an area of considerable interest owing to their striking nanoscale physicochemical properties which are markedly different from their bulk counterparts.1–9 Equipped with the advantages of nanoscale materials (small size, high surface to volume ratio, size and shape dependent properties), iron based nanoparticles have been successfully used in diverse technological fields mainly due to their novel optical, electronic and magnetic properties. These materials are profitably used in magnetic data storage devices, electronic devices and magnetic resonance imaging, and also as optical sensors, gas sensors, magnetically recoverable catalysts and photo-catalysts.1–5,10 In this context, it is pertinent to mention that iron oxide (Fe3O4 and α-Fe2O3) nanoparticles are biocompatible, non-toxic, chemically stable, low cost, abundant materials and have been successfully used in electrochemical sensing systems such as enzyme sensors, ethanol sensors, humidity sensors and DNA sensors.11–15
Hydrogen peroxide (H2O2) is extensively used in clinical, biopharmaceutical, environmental and industrial fields.16 Oxidation of glucose produces H2O2.17 Thus, detection of H2O2 is of immense importance in the field of biological science because it helps in the diagnosis of diseases like diabetes.18 On the other hand, hydrazine (N2H4) and its derivatives are generally used in fuel cells, explosives, rocket propellants, insecticides and plant growth regulators.19 Although these two materials are applied in diverse technological sectors, it has been found that they have adverse effects towards health and the environment.19 H2O2 is a strong oxidizing agent which causes corrosion of eyes and skin and may cause irreversible cell damage. It can also trigger cell proliferation, which may culminate into various diseases like cancer, cardiovascular and neurodegenerative disorders.20 Again, hydrazine is a carcinogenic, hepatotoxic and mutagenic material which may cause severe damage to the liver, kidneys and central nervous system.19,20 Therefore, sensitive and accurate detection of a small quantity of H2O2 and N2H4 is of great scientific importance. Up to now, numerous techniques have been proposed for the fruitful detection of H2O2 and N2H4.19,21,22 However, nonenzymatic amperometric sensing is the prime choice due to its high sensitivity, short response time and low cost of instrumentation. Furthermore, as nonenzymatic biosensors remove the impediments of enzyme-based sensors (like complex fabrication processes and low chemical/thermal stability), a lot of effort has been devoted to developing new enzyme free, stable and reproducible biosensors.23
It may be noted that α-Fe2O3 nanostructures show size and shape dependent properties, which have triggered lots of research into the controlled synthesis of α-Fe2O3 nanostructures.24–27 Up to now, a variety of α-Fe2O3 structures, such as 1D, 2D, and 3D have been developed. In 3D architectures α-Fe2O3 often exhibits more interesting electrical, optical, magnetic and catalytic properties, which are markedly superior compared to 2D or 1D architectures.24–27 The α-Fe2O3 nanorod array displays excellent selectivity and stability for the electrochemical detection of H2O2,28 whereas the α-Fe2O3 nanowire array exhibits excellent catalytic performance towards the oxidation of glucose.23 The α-Fe2O3 nanorings have been found to be effective in sensing of H2O2.29 Conversely, to the best of our knowledge there is no report on hydrazine sensing by α-Fe2O3 nano/micro structures. This issue has inspired us to synthesize a snowflake like 3D architecture of α-Fe2O3 and check its performance as a nonenzymatic H2O2 and N2H4 sensor.
In this paper, 3D micro-snowflake structured α-Fe2O3 has been synthesized using a facile one step solvothermal method. The sample has been thoroughly characterized by powder X-ray diffraction (PXRD), high resolution transmission electron microscopy (HRTEM), field emission scanning electron microscopy (FESEM), Fourier transform infrared (FTIR) spectroscopy and Mössbauer spectroscopy. The growth mechanism of the 3D micro-snowflake structured α-Fe2O3 has been followed by varying the reaction parameters. Further, detailed electrochemical analysis has been performed to study the electrocatalytic activity of the samples towards sensing of H2O2 and N2H4 in alkaline medium. It has been shown that the sample can be used as an amperometric probe for highly sensitive and selective detection of H2O2 and N2H4.
2. Materials and methods
2.1. Materials
Potassium ferricyanide (K3[Fe(CN)6]) was purchased from Loba Chemie and used without further purification for synthesis purposes. Other relevant reagents (potassium ferrocyanide (K4[Fe(CN)6]), potassium chloride (KCl), sodium hydroxide (NaOH), potassium hydroxide (KOH), hydrogen peroxide (H2O2), hydrazine (N2H4), glucose, ascorbic acid (AA), dopamine (DA), uric acid (UA), NaBr and NaCl) used in the experiments were also purchased from Loba Chemie and Mark Germany. Doubly distilled water was used as the solvent for all the experiments.
2.2. Synthesis of 3D micro-snowflake structured α-Fe2O3
0.2640 g of K3[Fe(CN)6] was dissolved in 80 mL deionized water. The aqueous solution was then transferred into a 100 mL Teflon-lined autoclave and the hydrothermal reaction was conducted at 180 °C for 24 (S1), 36 (S2), 48 (S3) and 96 (S4) hours. Then the autoclave was allowed to cool down to room temperature and the precipitates were washed several times in deionized water and ethanol by vigorous centrifugation. Afterward, the collected samples were dried at 70 °C for 10 hours in a vacuum oven. It may be noted that we have reported the H2O2 sensing property of the sample S1 in our earlier study.30 Herein, the sensing property of sample S4 has been reported and the other samples have been synthesized only to study the morphological changes taking place with the increase of reaction time.
2.3. Characterization of micro-snowflake structured α-Fe2O3
The powder X-ray diffraction (PXRD) patterns of the samples were recorded using a Bruker D8 Advanced Diffractometer using Cu Kα (λ = 1.54184 Å) radiation at ambient temperature (21 °C). The in depth structural and microstructural studies have been performed by using high resolution transmission electron microscopic (JEOL, 2100) and field emission scanning electron microscopic (FEI, INSPECT F50) techniques. The presence of constituent elements in sample S4 was probed using a BRUKER EDS system attached to the FESEM equipment. The FTIR spectrum of S4 was recorded using a Perkin Elmer spectrometer (Spectrum Two) equipped with an attenuated total reflectance (ATR) attachment in the wave number range 500–4000 cm−1. The room temperature Mössbauer spectrum of S4 was recorded in transmission geometry using constant acceleration drive (CMTE-250) with a 15 mCi 57Co source in a Rh matrix. A natural iron sample was used for calibration. The Mössbauer spectrum of the sample was fitted using the Recoil program.31 The solid circles represent experimental data points, whereas the solid lines represent the simulated pattern.
2.4. Fabrication of working electrodes and electrochemical experiments
Electrochemical characterizations (cyclic voltammetry and amperometry) were performed to evaluate the electro-catalytic ability of S4 towards sensing of H2O2 and N2H4. Electrochemical impedance spectroscopy (EIS) was performed to evaluate the charge transfer properties of the working electrode at the electrode/electrolyte interface. All electrochemical characterizations were performed in a standard three electrode configuration, with α-Fe2O3 coated ITO (Indium Tin Oxide coated glass) as the working electrode, Pt wire as the counter electrode and Ag/AgCl as the reference electrode. To prepare the working electrode, firstly the ITO sheets (1 × 1 cm2) were cleaned by ultrasonication in distilled water, ethanol and acetone for 10 minutes each and subsequently dried in air. 10 mg of S4 was dispersed in chitosan solution (1 wt% in acetic acid) and ultrasonicated for 1 hour to achieve a homogeneous suspension. 10 μL of the suspension was dropped on to the ITO coated glass and dried overnight for further measurements. The cyclic voltammetry (CV) and amperometry studies were performed in 0.1 M NaOH aqueous solution. EIS measurements were carried out in an electrolyte comprising of 5 mM of K4Fe(CN)6/K3Fe(CN)6 (1
:
1 mixture) as the redox probe with 0.1 M KCl as the supporting electrolyte.
3. Results and discussion
3.1. Structural study
In the powder X-ray diffraction (PXRD) patterns of all the samples (S1, S2, S3 and S4) only the characteristic diffraction peaks for the α-Fe2O3 phase (JCPDS no. 33-0664) have been observed and no additional peaks due to possible impurity phases have been noticed. This confirms that S1, S2, S3 and S4 are single phase α-Fe2O3. It is well known that the Rietveld refinement of powder X-ray patterns is a trustworthy tool for structural and microstructural characterization of nanometric oxides.1,5,6 It may be noted that α-Fe2O3 has a rhombohedrally centered hexagonal structure with a close-packed oxygen lattice in which two thirds of the octahedral sites are occupied by Fe3+ ions.32 To determine the crystal structure of S4 precisely we have fitted the PXRD pattern of the sample using the Rietveld refinement method using the GSAS program with a EXPGUI interface.33,34 The results are presented in Fig. 1. The detail of the methodology is reported in our earlier works.1,5,6 The final Rietveld refinement converged with an excellent agreement between the experimental and simulated patterns. The values of RP, RWP and χ2 are 0.02, 0.025 and 0.72, respectively. The results indicate that the sample is well crystalline α-Fe2O3 and it crystallizes in the hexagonal R
c space group. The miller indices and the space group obtained for the sample are in good agreement with the JCPDS database (JCPDS no. 33-0664). The values of the structural and microstructural parameters, fractional coordinates and occupancy of different ions are listed in Tables 1 and 2.
 |
| Fig. 1 (a) The final Rietveld plot (output file of GSAS) of the α-Fe2O3 sample, showing the difference (blue color line) between the experimental (black color line) and the simulated pattern (red color line). (b) Unit cell of the sample. | |
Table 1 Structural parameters of the sample obtained using the Rietveld refinement method
Crystal system |
Space group |
Lattice parameter (Å) |
Volume V [Å3] |
Crystalline size (nm) |
Microstrain |
Hexagonal |
R c |
a = b = 5.03564(4), c = 13.7589(2) |
302.151(0.01) |
18.2(0.1) |
3 × 10−3 (2 × 10−5) |
Table 2 Fractional coordinates and occupancy of different ions obtained using the GSAS program
Ions |
x |
y |
z |
Occupancy (±0.003) |
Fe (1) |
0 |
0 |
0.14451 |
0.025 |
Fe (2) |
0 |
0 |
0.35530 |
0.021 |
O (1) |
0.3051 |
0 |
0.25 |
0.025 |
O (2) |
−0.3061 |
−0.3061 |
0.25 |
0.024 |
3.2. FESEM, EDS and FTIR study on S4
The microstructural property and surface morphology of S4 has been investigated using field emission scanning electron microscopy (FESEM). The FESEM images of S4 both at high and low magnification are depicted in Fig. 2. The FESEM images clearly establish the formation of 3D micro-snowflake structures with dendritic arms. The individual dendritic arms consist of a long central trunk with a length of ∼5–6 μm with secondary branches with a length of ∼1–2 μm. The branches are uniformly distributed on the both sides of the central trunk.
 |
| Fig. 2 FESEM images of the 3D micro-snowflake structured α-Fe2O3 sample. | |
The presence of the constituent elements in the sample has been verified using an energy dispersive X-ray study (EDS). The EDS spectrum, illustrating the well resolved peaks originating from the constituent atoms in the energy range from 0 to 8 keV, is shown in Fig. S1.† The Au Lα peak in the EDS spectrum can be attributed to the gold coating of the sample prior to the EDS study. The EDS spectrum clearly indicates the presence of all the constituent elements (Fe, O) of α-Fe2O3.
The FTIR spectrum of the sample is shown in Fig. S2.† The strong peak at about 664 cm−1 appears due to the stretching vibrational mode of the Fe–O bond.1 The broad absorption bands centered around 3400 and 1633 cm−1 are assigned to the stretching vibration of O–H bonds of the free or absorbed water molecules in the sample.35
3.3. Mössbauer spectroscopic study
Magnetic ordering and hyperfine properties of iron containing materials can be easily investigated with the help of 57Fe Mössbauer spectroscopic studies.1,2,5,6,36 The room temperature Mössbauer spectrum (Fig. S3†) of S4 has been fitted using the Lorentzian site analysis method of the Recoil program. The Mössbauer spectrum at 300 K is best fitted by a single sextet with the value of the isomer shift (IS) and quadrupole splitting (QS) at 0.46 and 0.004 mm s−1, respectively. The value of the hyperfine field obtained from the fitting is 58.5 T. The value of the IS suggests that only Fe3+ ions are present in the sample.5,6 It may be noted that the hyperfine field of the present sample is substantially higher compared to its counterparts reported earlier.24,25 These provide clear evidence in favour of the presence of only a α-Fe2O3 phase in S4 and no traces of γ-Fe2O3 or Fe3O4 have been found in the Mössbauer spectrum of the sample.
3.4. Growth mechanism of 3D micro-snowflake structured α-Fe2O3
It is well known that the synthesis route, reaction parameters and use of growth controlling agents (surfactants) play a major role in determining the shape, size, porosity and specific surface area of the synthesized anisotropic complex 3-dimensional architectures of α-Fe2O3.24 To develop a standard, simple, reliable and high yielding reaction mechanism for the synthesis of complex architectures of α-Fe2O3 is a big challenge. Up to now, many types of complex architectures of α-Fe2O3 such as nanoparticles, nanocages, cantaloupe like structures, hollow spheres, dendrites and snowflakes have been synthesized by different chemical routes.12,23–27
Wang et al. have described a one pot hydrothermal route for the synthesis of a single crystalline dendritic micropine structure of α-Fe2O3.25 Recently, Bharathi et al. have shown that a hydrothermal reaction without surfactant leads to the formation of 3-D dendritic α-Fe2O3 structures, whereas the use of CTAB and PEG as surfactants during the reaction results in single and double layered snowflake structures of six-fold symmetry, respectively.24 According to Zhen et al., NP-9 surfactant converts dendritic micropines of α-Fe2O3 to hexagonal micro-snowflakes.37 Liu et al. have shown how the morphology of α-Fe2O3 changes from snowflakes to starfish then flowers and finally to hexagonal plates upon increasing the concentration of octylamine.38 Sun et al. prescribed a hydrothermal reaction in the presence of benzoic acid for precise tailoring of α-Fe2O3 dendritic structures.39 The influence of reaction time and pH of the hydrothermal chamber on the growth of different α-Fe2O3 structures have been studied by Zhang et al.40 They have shown that in a basic pH (∼12), the dendritic α-Fe2O3 structures convert into snowflake like microstructures with enhanced reaction time. Further increase in the reaction time promotes more crystal growth and forms paired microplates and dumbbell-like or spindle-like microstructures.40
In the present case we have synthesized a 3D architecture of micro-snowflake structured α-Fe2O3 from K3[Fe(CN)6] using a hydrothermal technique in normal pH without using any surfactant. The reaction time has been varied to study the evolution of the structure and morphology of the sample. We have carried out the hydrothermal reaction for different time periods (24 (S1), 36 (S2), 48 (S3) and 96 (S4) hours) by maintaining the reaction temperature at 180 °C. The FESEM micrographs of the samples are given in Fig. 3. The FESEM micrograph of S1 suggests that this sample possesses a 3D dendritic micropine like structure with a single symmetric branched arm. This structure consists of a single central trunk with secondary branches on both sides of the central trunk. The morphology of S2 is similar to that of S1. However, S2 has more fractal branches than S1. This suggests that fractal branch formation increases with the increase of reaction time (from 24 to 36 hours). The FESEM images of S2, S3 and S4 suggest that with the increase of reaction time the micropine like structures get attached to form micro-snowflakes of α-Fe2O3 (sample S3) with symmetric branching along each arm. This micro-snowflake structure has six-fold symmetry.
 |
| Fig. 3 Schematic diagram of the formation of 3D micro-snowflake structured α-Fe2O3 (above) and FESEM images of the synthesized samples S1, S2, S3 and S4 (below). | |
Faster crystal growth rate along one of the six crystallographic directions compared to the other five crystallographic directions results in the formation of the dendritic micropine like structure of α-Fe2O3. The TEM micrograph of S4 (Fig. 4) shows a clear and well defined dendritic fractal structure with a prolonged trunk with ordered branches on both sides of it. The length of the main trunk is 5.37 μm, whereas the length of the branches range from 1.25–1.6 μm. Actually, the formation of desired microstructures of α-Fe2O3 depends upon the dissociation rate of [Fe(CN)6]3− ions in the hydrothermal reaction. S. Bharati et al. have shown that the use of metal salts in the hydrothermal reaction, which readily dissociate in aqueous solution (FeCl3, Fe(NO3)3), results in irregular shaped nanoparticles.24 On the contrary, the [Fe(CN)6]3− ion is highly stable at room temperature. Hydrothermal treatment of the aqueous solution of K3[Fe(CN)6] at a high temperature (180 °C) for an extended time results in the dissociation of [Fe(CN)6]3− ions into Fe3+ ions, which upon further hydrolysis get converted into FeOOH/Fe(OH)3. Due to the high temperature hydrothermal reaction the FeOOH/Fe(OH)3 further dissociates into α-Fe2O3.24,25
 |
| Fig. 4 (a) TEM micrograph of each symmetric branch of micro-snowflake structured α-Fe2O3, (b) TEM micrograph of micro-snowflake structured α-Fe2O3, (c) HRTEM showing atomic planes with 0.25 nm spacing, and (d) SAED pattern of the sample. | |
α-Fe2O3 crystallizes in a hexagonal closed pack structure. The Fe3+ and O2− ions arrange alternately in a regular fashion, parallel to the c direction, i.e., [0001] direction. It may be noted that each layer is terminated with Fe3+ and O2− ions, thus developing an ionic charge on the surface. The c plane, being a polar surface, neutralizes the ionic charge by adsorbing molecules present in the hydrothermal chamber and the growth along the [0001] direction gets terminated. On the other hand, growth proceeds along the other six crystallographically equivalent directions ±[10
0], ±[1
00], and ±[0
10] which give rise to the formation of hexagonally shaped particles.24,25 The stacking of α-Fe2O3 nanoparticles along the [1
00] direction leads to the formation of the main trunk, while the formation of branches can be assigned to the slower crystallization rate along the equivalent [10
0] and [0
10] directions. By extending the reaction time to 48 hours, the complete snowflake like structure of α-Fe2O3 appears.24,25 The prolonged reaction time promotes attachment to form the micro-snowflake structure from micropines and this gives rise to a single-layered snowflake like structure. The HRTEM image (Fig. 4(c)) taken at the periphery of the branch shows clear lattice fringes, which indicate that the sample is crystalline in nature. The lattice spacing of 2.5 Å corresponds to the {11
0} crystal planes. The selected area electron diffraction (SAED) pattern of S4 (Fig. 4(d)) clearly indicates that the sample is single crystalline in nature. The central spot of the SAED pattern corresponds to [0001] plane. The SAED pattern suggests that the branches are formed along the six crystallographically equivalent directions ±[10
0], ±[1
00], and ±[0
10].
4. Electrochemical sensing study
4.1. Electrochemical activity and EIS study
The electrocatalytic activity of S4 was evaluated using both cyclic voltammetry and EIS techniques. It may be noted that the intensity of the oxidation and reduction peak markedly increases (the current density at the oxidation potential for ITO and S4 coated ITO are 100 and 200 μA cm−2, respectively) and the difference between the oxidation and reduction potential decreases (the redox potential for ITO and S4 coated ITO are 0.8 and 0.075 V, respectively) upon coating the ITO substrate with S4 (Fig. 5). This was further confirmed by examining the interfacial charge transfer resistance at the electrode-electrolyte interface using EIS. Fig. 6 displays the electrochemical impedance spectra of bare ITO and S4 coated ITO in the frequency range between 0.1 Hz and 100 kHz. The plot consists of a semicircular arc in the high frequency region followed by a linear part. The diameter of the semicircular region is a measure of the charge transfer resistance at the electrode/electrolyte interface and the linear part corresponds to a diffusion limited process. It is evident from this figure that the diameter of the semicircular arc in the impedance spectrum of the S4 coated ITO working electrode is smaller than that of bare ITO and this in turn indicates that the interfacial resistance at the electrode-electrolyte interfaces for the S4 coated ITO working electrode is less than that of bare ITO. Thus, the S4 coated ITO working electrode provides a lower interfacial resistance in the path of charge transport compared to that of bare ITO and it favors the movement of charges through it. It seems that S4 coated ITO may show good electrocatalytic activity.
 |
| Fig. 5 Cyclic voltammetry in 5 mM K3Fe(CN)6/K4Fe(CN)6 redox probe (1 : 1) with 0.1 M KCl as supporting electrolyte. | |
 |
| Fig. 6 Electrochemical impedance spectra (EIS) for both electrodes. | |
4.2. Electrochemical sensing of H2O2
Fig. 7 depicts the CV plot for the reduction of H2O2 at the S4 coated ITO working electrode conducted in 0.1 M NaOH solution as the electrolyte between the potential range of 0 and −1 V at a scan rate of 0.1 V s−1. According to an early report, a sharp reduction peak appears at −0.82 V in the CV plot of a α-Fe2O3 nanorod coated stainless steel (SS) substrate with 1 M KOH as the electrolyte due to the reduction of α-Fe2O3.41 Thus, in the CV plot of the bare S4 embedded ITO electrode the peak at −0.8 V can be attributed to the reduction of α-Fe2O3. It is noteworthy that with increasing molar concentrations of H2O2 from 0.1 to 0.2 mM, the current density at the reduction potential of −0.8 V increases markedly compared to that of the bare S4 coated ITO electrode indicating the enhancement of the electrochemical response of S4 coated ITO upon addition of H2O2 in the electrolyte. Whenever H2O2 is added in an alkaline solution it is converted to either HO2− or O22−. In the present case the working electrode is actually detecting the presence of H2O2 in the electrolyte (NaOH) by sensing the subsequent reduction of these anionic byproducts. The α-Fe2O3 coating on the working electrode may therefore be identified as an electrocatalyst responsible for the indirect detection of H2O2. The effect of scan rate on the reduction of 0.1 mM H2O2 is shown in Fig. S4.† The inset of Fig. S4† shows that the current density at the reduction potential increases linearly with scan rate, thereby indicating a surface confined process.42
 |
| Fig. 7 Cyclic voltammetry (CV) study with different molar concentrations of H2O2. | |
Amperometric current density (J) versus time (t) (J–t) measurements were performed to evaluate the H2O2 sensing performance of S4 coated ITO at the reduction potential of −0.8 V (Fig. 8). A H2O2 stock solution of different molar concentrations was added successively to a homogeneously stirred NaOH solution at regular intervals of 50 s. In this experiment stable current steps were obtained as shown in Fig. 8. The corresponding calibration curve shown in the inset of Fig. 8 can be fitted with the equation J = 7.16 μA mM−1 cm−2CH2O2 − 4.8968 with a correlation coefficient R2 = 0.99434. The value of sensitivity calculated from the slope of the calibration plot is obtained as 7.16 μA mM−1 cm−2. Further, the lowest detection limit was estimated to be 0.01 mM at S/N = 3.
 |
| Fig. 8 Amperometric J–t curve of the sample for the H2O2 sensing study, linear response of H2O2 concentration with the current values (inset). | |
The analytical performance (sensitivity, detection limit and linear response range) of the 3D micro-snowflake structured α-Fe2O3 is compared with previously reported H2O2 sensors and the results are shown in Table 3. From Table 3 it is evident that the sensitivity of the proposed H2O2 sensor is the second best.17,18,20,43–59 Moreover, according to our previous work30 the sensitivity of sample S1 is 0.8 μA mM−1 cm−2 for H2O2 sensing. But in these 3D micro-snowflake structured α-Fe2O3 architectures the higher sensitivity value of the system can be attributed to the additional catalytic sites available due to the three dimensional morphology of the structure resulting in favorable dissociation of H2O2 leading to higher current density and improved sensitivity.
Table 3 Comparison of the analytical performance of the proposed H2O2 biosensor with other H2O2 biosensors reported previously
Electrode |
Sensitivity |
Linear range |
Detection limit |
Reference |
FeS nanosheet/GC |
— |
0.5–150 μM |
9.2 × 10−8 M |
43 |
NiFe2/OMC |
4.29 μA mM−1 cm−2 |
6.2–42 710 μM |
0.24 μM |
17 |
Pt/CNT |
1.4 μA mM−1 cm−2 |
5.0–25 000 μM |
1.5 μM |
44 |
Pt/OMC/Au |
6.83 μA mM−1 cm−2 |
— |
42 μM |
45 |
OMC/Au |
0.071 μA mM−1 cm−2 |
— |
0.73 μM |
45 |
Mesoporous Pt |
2.8 μA mM−1 cm−2 |
20–40 000 μM |
4.5 μM |
46 |
Ag@CuNW |
— |
1–10 mM |
3 μM |
47 |
MnO2 microspheres |
— |
10–150 mM |
2.0 μM |
48 |
Mn2O3 hollow spheres |
— |
0.1–126.5 μM |
0.07 μM |
49 |
Ag NP/MB biocomposite |
0.357 μA μM−1 cm−2 |
1 μM to 3 mM |
0.088 μM |
50 |
Ag/SnO2/GCE |
— |
0.01 to 35 mM |
5 μM |
51 |
Ag/RGO/ITO |
— |
0.1 to 100 mM |
1.6 μM |
52 |
PB-graphene/GCE |
0.53 μA μM−1 cm−2 |
10–1440 μM |
3 μM |
53 |
PANI–PS–sulphonate-carboxylated graphene modified graphite |
201.4 nA μM−1 cm−2 |
25–350 μM |
1 μM |
54 |
Fe2O3/GC |
2.5 μA μM−1 cm−2 |
10–150 μM |
7 μM |
55 |
Au NPs–N–GQDs |
182.22 μA mM−1 cm−2 |
0.25–13 327 μM |
0.12 μM |
56 |
Ta/Fe–Phen–CFS |
— |
0.1–100 μM |
68 nM |
20 |
Fe3C/NG |
9.44 (μA L−1) mM−1 |
50 μM L−1 to 15 mM L−1 |
0.035 Mm L−1 |
18 |
MnO2/CF2 |
0.054 μA μM−1 cm−2 |
2.5–2055 μM |
0.12 μM |
57 |
Co3O4–NWs/CF |
230 nA μM−1 cm−2 |
0.01–1.4 mM |
1.4 μM |
58 |
Cu2O/GNs |
— |
0.3–7.8 mM |
20.8 μM |
59 |
3D micro-snowflake structured α-Fe2O3 |
7.16 μA mM−1 cm−2 |
0.1–5.5 mM |
0.01 mM |
This work |
4.3. Electrochemical hydrazine sensing
Electrochemical hydrazine (N2H4) sensing studies were conducted in 0.1 M NaOH electrolyte. Fig. 9 shows the comparative CV in the absence and presence of 10 mM N2H4. Electro-oxidation of N2H4 was observed at 0.1 V in the background NaOH electrolyte. The cyclic voltammograms of 10 mM N2H4 at different scan rates using S4 coated ITO as the working electrode are depicted in Fig. S5.†
 |
| Fig. 9 Comparative CV in the presence and absence of 10 mM hydrazine. | |
The amperometric studies for N2H4 were performed at the operating potential of 0.1 V corresponding to the oxidation peak current in the CV and the results are presented in Fig. 10. In this amperometric J–t curve stable steps are observed upon successive addition of 50 μM N2H4 (Fig. 10). The current density versus concentration curve (calibration plot) is linear in nature and fitted by the regression equation J = 0.02403Chydrazine − 1.19176 with correlation coefficient R2 = 0.9843. The value of the sensitivity obtained from the calibration curve is 24.03 μA mM−1 cm−2 and the lowest detection limit is estimated to be 5 μM at S/N = 3. The limit of detection (LOD) value is quite impressive in comparison with other noble metal based N2H4 sensing platforms. The amperometric measurements remain stable despite the evolution of N2 indicating a satisfactory response for hydrazine detection. The analytical performance (sensitivity, detection limit and linear response range) of the 3D micro-snowflake structured α-Fe2O3 was compared with previously reported hydrazine sensors and is presented in Table 4. From Table 4 it is clear that the proposed N2H4 sensor shows excellent electrochemical activity compared to previously reported hydrazine sensors.60–68
 |
| Fig. 10 Amperometric J–t curve of the sample for the hydrazine sensing study, and linear response of hydrazine concentration with the current values. | |
Table 4 Comparison of the analytical performance of the proposed hydrazine biosensor with other hydrazine biosensors reported previously
Electrode |
Sensitivity |
Linear range (μM) |
Detection limit (μM) |
Reference |
Ag nanocubes/GCE |
— |
5–460 |
1.1 |
60 |
Pd–TiO2 |
0.55 mA mM−1 cm−2 |
1000–20 000 |
23 |
61 |
CuS/RGO |
7.96 μA mM−1 |
1–1000 |
0.3 |
62 |
Graphene-Bi/GCE |
1.49 μA mM−1 cm−2 |
0.02–280 |
0.005 |
63 |
Pd/carbon black/GCE |
— |
0–500 |
13.4 |
64 |
Au NPs/graphite |
— |
25–1000 |
3.07 |
65 |
Au NPs/ZnO-MWCNT/GCE |
0.0428 μA μM−1 cm−2 |
0.5–1800 |
0.15 |
66 |
Au NPs/poly(BCP)/CNT/GCE |
— |
0.5–1000 |
0.1 |
67 |
Ag dendritics/ITO |
20.81 μA mM−1 cm−2 |
100–1700 |
0.5 |
68 |
3D micro-snowflake structured α-Fe2O3 |
24.03 μA mM−1 cm−2 |
50–1340 |
5 |
This work |
4.4. Interference study
The most important consideration for any sensing platform is its selectivity in the presence of other coexisting electroactive species. The same was tested for H2O2 reduction and N2H4 oxidation in the present case and the results are shown in Fig. 11(a) and (b), respectively. The H2O2 reduction selectivity was investigated in the presence of ascorbic acid (AA), dopamine (DA), uric acid (UA) and glucose solutions. The amperometric J–t plot indicates that the α-Fe2O3 coated ITO electrode shows a high response towards H2O2 and almost negligible response with the addition of AA, DA, UA and glucose. The selectivity of the α-Fe2O3 coated ITO electrode towards H2O2 sensing may be due to the fact that when AA, DA, UA and glucose are added in the NaOH electrolyte either no anionic species are evolved in the solution, as in the case of H2O2, or there is no reduction peak for AA, DA, UA and glucose at about the working potential of −0.8 V. The selectivity of N2H4 oxidation was evaluated in the presence of AA, DA, UA, glucose, Br−, Cl− and K+. From the amperometric J–t plot, it can be inferred that the response of the α-Fe2O3 coated ITO electrode towards N2H4 oxidation is highly selective with almost no change in the current density upon the addition of glucose, Br−, Cl− and K+. However, a slight increase in current has been observed for the addition of AA which may be ascribed to the tendency of AA to oxidize at low potentials. The experimental results suggest that the tolerance level of α-Fe2O3 microstructures for Br−, Cl− and K+ ions is high.59 It may be noted that the oxidation of DA, UA and glucose usually takes place at higher potential.59 As we have carried out the measurements at a low working potential (0.1 V) the oxidation of DA, UA and glucose does not interfere with amperometric response of N2H4. The cause behind the selectivity of the α-Fe2O3 coated electrode for the oxidation of N2H4 can also be due to the reason that the oxidation of N2H4 in alkaline medium is a kinetic-controlled reaction while the oxidation of the other interferences used is a diffusion-controlled electrochemical process.59 If these are facts, then the better performance of N2H4 oxidation can be attributed to the presence of a large number of active catalytic sites due to the large surface area of the 3D snowflake architecture.
 |
| Fig. 11 Selectivity in the presence of (a) H2O2 and (b) hydrazine with other co-existing analytes. | |
4.5. Reproducibility and long term stability
The reproducibility of the sensing performance for both H2O2 and N2H4 was evaluated by recording the amperometric response of 5 identically prepared working electrodes. Fig. 12(a) and 13(a) show the plots of the sensitivity values for the electrodes towards H2O2 and N2H4 and the relative standard deviation for H2O2 and N2H4 are 1.62% and 1.81%, respectively. The long term stability was studied over a period of four weeks by performing the electrochemical tests after each week and the results for H2O2 and N2H4 are shown in Fig. 12(b) and 13(b), respectively. The study on H2O2 has revealed that the current density is 80% of its initial value after the first round of usage and drops to 60% of its initial value after four weeks. The N2H4 oxidation study also shows good long term stability with the oxidation peak current retaining 60% of its initial value at the end of four weeks. Hence, the S4 coated working electrode shows acceptable reproducibility and good long term stability towards the reduction of H2O2 and oxidation of N2H4.
 |
| Fig. 12 (a) The reproducibility of the sensing performance for H2O2, (b) reproducibility and good long term stability towards H2O2. | |
 |
| Fig. 13 (a) The reproducibility of the sensing performance for hydrazine, (b) reproducibility and good long term stability towards hydrazine. | |
5. Conclusion
Herein, we have successfully synthesized 3D micro-snowflake structured α-Fe2O3 by a simple hydrothermal reaction without using any surfactant and followed the growth mechanism with the variation of reaction time. We have employed PXRD, FESEM, TEM, FTIR spectroscopy and Mössbauer spectroscopy to characterize the sample. The crystal structure of S4 has been analyzed by Rietveld refinement of the PXRD pattern of the sample. FESEM and TEM images show that S4 exhibits a micro-snowflake like shape having six-fold symmetry with symmetric branching along each arm consisting of a long central trunk and secondary branches. The results of the PXRD, FTIR and Mössbauer spectroscopic studies together confirmed that this micro-snowflake like structure is formed by pure, crystalline and single phase α-Fe2O3. The sample displayed excellent electrochemical sensing performance towards H2O2 reduction (sensitivity 7.16 μA mM−1 cm−2 in a wide range from 0.1 mM to 5.5 mM) and N2H4 oxidation (sensitivity 24.03 μA mM−1 cm−2 in the linear range between 50 μM and 1340 μM).
In summary, we have shown that the 3D micro-snowflake structured α-Fe2O3 synthesized by simple hydrothermal decomposition of K3[Fe(CN)6] without using any surfactant can be used for highly selective, sensitive and stable amperometric sensing of H2O2 and N2H4 in the presence of common coexisting electroactive interferences. We have examined how different morphological forms evolve for microstructured α-Fe2O3 synthesized by this hydrothermal route with the variation of reaction time by keeping the temperature and reactant concentration the same. We have shown that the morphology of α-Fe2O3 can be gradually tailored from micropines to micropines with fractal branches and finally to 3D micro-snowflake structures simply by changing the reaction time without using any growth controlling agent (surfactant). It may be noted that the electrochemical performance of the S4 modified ITO working electrode for H2O2 sensing is second best among all the H2O2 sensing electrodes reported to date. Moreover, to the best of our knowledge, the electrochemical N2H4 sensing activity of α-Fe2O3 is explored for the first time. As an ideal enzyme-less sensing material, the sample has good stability and selectivity against common coexisting interferences. Hence, α-Fe2O3 coated ITO working electrodes could be suitable platforms for enzyme-less H2O2 and N2H4 sensing.
Acknowledgements
Authors (S. Majumder and B. Saha) gratefully acknowledge UGC, Govt. of India for providing research fellowship. S. Dey and R. Mondal gratefully acknowledge CSIR, Govt. of India and DST, Govt. of India, respectively, for providing fellowship. The UPE (JU) and DSA (Physics Department, JU) program of UGC, the PURSE and FIST programs (Physics Department, JU) of DST, Government of India are also acknowledged. We are thankful to Dr Sourav Das, Associate Professor, Department of Chemistry, JU for helpful discussions.
References
- S. Majumder, S. Dey, K. Bagani, S. K. Dey, S. Banerjee and S. Kumar, Dalton Trans., 2015, 44, 7190 RSC.
- S. Dey, S. K. Dey, B. Ghosh, V. R. Reddy and S. Kumar, Mater. Chem. Phys., 2013, 138, 833 CrossRef CAS.
- A. Tiano, G. C. Papaefthymiou, C. S. Lewis, J. Han, C. Zhang, Q. Li, C. Shi, A. M. M. Abeykoon, S. J. L. Billinge, E. Stach, J. Thomas, K. Guerrero, P. Munayco, J. Munayco, R. B. Scorzelli, P. Burnham, A. J. Viescas and S. S. Wong, Chem. Mater., 2015, 27, 3572 CrossRef CAS.
- M. M. Shirolkar, R. Das, T. Maity, P. Poddar and S. K. Kulkarni, J. Phys. Chem. C, 2012, 116, 19503 CrossRef CAS.
- S. Dey, S. K. Dey, K. Bagani, S. Majumder, A. Roychowdhury, S. Banerjee, V. R. Reddy and S. Kumar, Appl. Phys. Lett., 2014, 105, 063110 CrossRef.
- S. Dey, S. K. Dey, B. Ghosh, P. Dasgupta, A. Poddar, V. R. Reddy and S. Kumar, J. Appl. Phys., 2013, 114, 093901 CrossRef.
- S. Majumder, S. K. Jana, K. Bagani, B. Satpati, S. Kumar and S. Banerjee, Opt. Mater., 2015, 40, 97 CrossRef CAS.
- V. Šepelák, A. Feldhoff, P. Heitjans, F. Krumeich, D. Menzel, F. J. Litterst, I. Bergmann and K. D. Becker, Chem. Mater., 2006, 18, 3057 CrossRef.
- S. Bedanta and W. Kleeman, J. Phys. D: Appl. Phys., 2009, 42, 013001 CrossRef.
- S. K. Jana, S. Majumder, B. Satpati, S. K. Mishra, R. K. Srivastava and S. Banerjee, RSC Adv., 2015, 5, 37729 RSC.
- P. V. Adhyapak, U. P. Mulik, D. P. Amalnerkar and I. S. Mulla, J. Am. Ceram. Soc., 2013, 96, 731 CrossRef CAS.
- Y. Wang, J. L. Cao, M. G. Yu, G. Sun, X. D. Wang, H. Bala and Z. Y. Zhang, Mater. Lett., 2013, 100, 102 CrossRef CAS.
- J. Wang, H. Gao, F. Sun, Q. Hao and C. Xu, Biosens. Bioelectron., 2013, 42, 550 CrossRef CAS PubMed.
- B. Xu, D. Zheng, W. Qiu, F. Gao, S. Jiang and Q. Wang, Biosens. Bioelectron., 2015, 72, 175 CrossRef CAS PubMed.
- N. Li, X. Hao, B. H. Kang, Z. Xu, Y. Shi, N. B. Li and H. Q. Luo, Biosens. Bioelectron., 2016, 77, 525 CrossRef CAS PubMed.
- M. Y. Wang, T. Shen, M. Wang, D. E. Zhang, Z. W. Tong and J. Chen, Sens. Actuators, B, 2014, 190, 645 CrossRef CAS.
- D. Xiang, L. Yin, J. Ma, E. Guo, Q. Li, Z. Li and K. Liu, Analyst, 2015, 140, 644 RSC.
- R. Zhang and W. Chen, Sci. Bull., 2015, 60, 522 CrossRef CAS.
- Y. You, Y. Yang and Z. Yang, J. Solid State Electrochem., 2013, 17, 701 CrossRef CAS.
- R. Zhang, S. He, C. Zhang and W. Chen, J. Mater. Chem. B, 2015, 3, 4146 RSC.
- B. Saha, S. K. Jana, S. Majumder, B. Satpati and S. Banerjee, Electrochim. Acta, 2015, 174, 853 CrossRef CAS.
- K. C. Fang, C. P. Hsu, Y. W. Kang, J. Y. Fang, C. C. Huang, C. H. Hsu, Y. F. Huang, C. C. Chen, S. S. Li, J. A. Yeh, D. J. Yao and Y. L. Wang, Biosens. Bioelectron., 2014, 55, 294 CrossRef CAS PubMed.
- X. Cao and N. Wang, Analyst, 2011, 136, 4241 RSC.
- S. Bharathi, D. Nataraj, M. Seetha, D. Mangalaraj, N. Ponpandian, Y. Masuda, K. Senthil and K. Yong, CrystEngComm, 2010, 12, 373 RSC.
- M. Cao, T. Liu, S. Gao, G. Sun, X. Wu, C. Hu and Z. L. Wang, Angew. Chem., Int. Ed., 2005, 44, 4197 CrossRef CAS PubMed.
- Z. Liu, B. Lv, D. Wu, Y. Zhu and Y. Sun, CrystEngComm, 2012, 14, 4074 RSC.
- Y. Jiao, Y. Liu, F. Qu and X. Wu, CrystEngComm, 2014, 16, 575 RSC.
- X. Liu, J. Liu, Z. Chang, L. Luo, X. Lei and X. Sun, RSC Adv., 2013, 3, 8489 RSC.
- X. Hu, J. C. Yu, J. Gong, Q. Li and G. Li, Adv. Mater., 2007, 19, 2324 CrossRef CAS.
- S. Majumder, B. Saha, S. Dey, K. Bagani, M. K. Roy, S. K. Jana, S. Kumar and S. Banerjee, AIP Conf. Proc., 2015, 1665, 050117 CrossRef.
- K. Lagarec and D. G. Rancourt, Recoil-Mössbauer Spectral Analysis Software for Windows, University of Ottawa Press, Ottawa, 1998 Search PubMed.
- L. Machala, J. Tuček and R. Zbořil, Chem. Mater., 2011, 23, 3255 CrossRef CAS.
- A. C. Larson and R. B. Von Dreele, General structure analysis system (GSAS), Los Alamos National Laboratory, Report LAUR, 2000, pp. 86–748 Search PubMed.
- B. H. Toby, J. Appl. Crystallogr., 2001, 34, 210 CrossRef CAS.
- R. Suresh, R. Prabu, A. Vijayaraj, K. Giribabu, A. Stephen and V. Narayan, Mater. Chem. Phys., 2012, 134, 590 CrossRef CAS.
- S. Dey, R. Gomes, R. Mondal, S. K. Dey, P. Dasgupta, A. Poddar, V. R. Reddy, A. Bhaumik and S. Kumar, RSC Adv., 2015, 5, 78508 RSC.
- K. He, C. Y. Xu, L. Zhen and W. Z. Shao, Mater. Lett., 2008, 62, 739 CrossRef CAS.
- Z. Liu, C. Y. Chiang, W. Li and W. Zhou, Chem. Commun., 2015, 51, 9350 RSC.
- G. Sun, B. Dong, M. Cao, B. Wei and C. Hu, Chem. Mater., 2011, 23, 1587 CrossRef CAS.
- X. Zhang, C. Sui, J. Gong, Z. Su and L. Qu, J. Phys. Chem. C, 2007, 111, 9049 CrossRef CAS.
- D. Sarkar, A. Shukla and D. D. Sharma, ACS Energy Lett., 2016, 1, 82 CrossRef CAS.
- J. Wang, D. F. Thomas and A. Chen, Anal. Chem., 2008, 80, 997 CrossRef CAS PubMed.
- Z. Dai, S. Liu, J. Bao and H. Ju, Chem.–Eur. J., 2009, 15, 4321 CrossRef CAS PubMed.
- Z. Wen, S. Ci and J. Li, J. Phys. Chem. C, 2009, 113, 13482 CrossRef CAS.
- X. Jiang, Y. Wu, X. Mao, X. Cui and L. Zhu, Sens. Actuators, B, 2011, 153, 158 CrossRef CAS.
- S. A. G. Evans, J. M. Elliott, L. M. Andrews, P. N. Bartlett, P. J. Doyle and G. Denuault, Anal. Chem., 2002, 74, 1322 CrossRef CAS PubMed.
- J. S. Easow and T. Selvaraju, Electrochim. Acta, 2013, 112, 648 CrossRef CAS.
- L. Zhang, Z. fang, Y. Ni and G. Zhao, Int. J. Electrochem. Sci., 2009, 4, 407 CAS.
- C. Cheng, Y. Huang, N. Wang, T. Jiang, S. Hu, B. Zheng, H. Yuan and D. Xiao, ACS Appl. Mater. Interfaces, 2015, 7, 9526 Search PubMed.
- S. Palanisamy, C. Karuppiah, S. M. Chen, R. Emmanuel, P. Muthukrishnan and P. Prakash, Sens. Actuators, B, 2014, 202, 177 CrossRef CAS.
- Y. E. Miao, S. He, Y. Zhong, Z. Yang, W. W. Tjiu and T. Liu, Electrochim. Acta, 2013, 99, 117 CrossRef CAS.
- A. M. Golsheikh, N. M. Huang, H. M. Lim, R. Zakaria and C. Y. Yin, Carbon, 2013, 62, 405 CrossRef.
- Y. Jiang, X. Zhang, C. Shan, S. Hua, Q. Zhang, X. Bai, L. Dan and L. Niu, Talanta, 2011, 85, 76 CrossRef CAS PubMed.
- S. Prasannakumar, R. Manjunath, C. Nethravathi, G. S. Suresh, M. Rajamathi and T. V. Venkatesha, Port. Electrochim. Acta, 2012, 30, 371 CrossRef CAS.
- A. K. Dutta, S. K. Maji, D. N. Srivastava, A. Mondal, P. Biswas, P. Paul and B. Adhikary, J. Mol. Catal. A: Chem., 2012, 360, 71 CrossRef CAS.
- J. Ju and W. Chen, Anal. Chem., 2015, 87, 1903 CrossRef CAS PubMed.
- S. He, B. Zhang, M. Liu and W. Chen, RSC Adv., 2014, 4, 49315 RSC.
- M. Liu, S. He and W. Chen, Nanoscale, 2014, 6, 11769 RSC.
- M. Liu, R. Liu and W. Chen, Biosens. Bioelectron., 2013, 45, 206 CrossRef CAS PubMed.
- Y. Wang, X. Yang, J. Bai, X. Jiang and G. Fan, Biosens. Bioelectron., 2013, 43, 180 CrossRef CAS PubMed.
- Q. Yi, F. Niu and W. Yu, Thin Solid Films, 2011, 519, 3155 CrossRef CAS.
- Y. J. Yang, W. Li and X. Wu, Electrochim. Acta, 2014, 123, 260 CrossRef CAS.
- R. Devasenathipathy, V. Mani and S. M. Chen, Talanta, 2014, 124, 43 CrossRef CAS PubMed.
- J. Panchompoo, L. Aldous, C. Downing and A. Crossley, Electroanalysis, 2011, 23, 1568 CrossRef CAS.
- M. A. Aziz and A. N. Kawde, Talanta, 2013, 115, 214 CrossRef PubMed.
- C. Zhang, G. Wang, Y. Ji, M. Liu, Y. Feng, Z. Zhang and B. Fang, Sens. Actuators, B, 2010, 150, 247 CrossRef CAS.
- S. Kocak and B. Aslısen, Sens. Actuators, B, 2014, 196, 610 CrossRef CAS.
- R. Sivasubramanian and M. V. Sangaranarayanam, Sens. Actuators, B, 2015, 213, 92 CrossRef CAS.
Footnotes |
† Electronic supplementary information (ESI) available. See DOI: 10.1039/c6ra10470c |
‡ These authors (S. Majumder and B. Saha) contributed equally. |
|
This journal is © The Royal Society of Chemistry 2016 |
Click here to see how this site uses Cookies. View our privacy policy here.