DOI:
10.1039/C5RA11512D
(Paper)
RSC Adv., 2015,
5, 68964-68971
Ultra-sensitive polyaniline–iron oxide nanocomposite room temperature flexible ammonia sensor†
Received
16th June 2015
, Accepted 6th August 2015
First published on 6th August 2015
Abstract
A novel flexible, ultra-sensitive, selective, and room temperature operable polyaniline/iron-oxide (PAni/α-Fe2O3) nanocomposite ammonia (NH3) gas sensor was developed onto a flexible polyethylene terephthalate (PET) substrate by in situ polymerization process. The observations were recorded to 100 ppm fixed level for various gases including NO2, CH3OH, C2H5OH, NH3, and H2S through monitoring the change in resistance of the developed sensor. The flexible PAni/α-Fe2O3 nanocomposite sensor demonstrated better selectivity towards NH3 (response = 39% and stability = 74%). The synergistic response of the flexible PAni/α-Fe2O3 sensor was remarkable than that of the PAni and α-Fe2O3 alone; indicating the effective improvement in the performance of PAni flexible sensor on nanocomposite process. Moreover, the flexible sensor detected NH3 at low concentration (5 ppm) with a fast response (27 s) and very short recovery time (46 s). Further, PAni/α-Fe2O3 flexible sensor films were characterized by X-ray diffraction, field-emission scanning electron microscopy, UV-visible and Raman spectroscopy, Fourier transform infrared and X-ray photoelectron for structural analysis, morphological evolution, optical and surface related studies.
1. Introduction
Detecting ammonia (NH3) gas at low levels has attracted considerable attention due to its toxicity and its impact on living organisms. According to the US Environmental Protection Agency, a concentration of NH3 above 25 ppm creates health issues. Therefore, it is essential to develop an NH3 sensor which is reliable and inexpensive for detecting below the toxic limit. Recently, flexible substrate sensor has been attracting considerable attention1 due to its interesting characteristics, such as softness, lightness, flexibility and shock resistivity. In general, gas sensors require high temperature to operate effectively and efficiently (>200 °C). This is above the continuous operating temperature of most of the polymers, thus, limiting their direct application.2 So fabricating a sensor, operating at low-temperature i.e. room (≤45 °C), not only avoids the heating assembly but also makes the sensor setup simpler, portable, and cheaper.3 Most of the conducting polymers (e.g. polyaniline (PAni), polypyrrole, polythiophene etc.) demonstrate potential gas sensing properties.4–6 Their advantages include low cost, flexibility in design, tunability, low power consumption and excellent performance at room temperature7,8 etc. PAni is one of the scientifically potential polymers with versatile applications including gas sensors and emerged as a star material in the field electrochemical sciences because of its intriguing properties such as high conductivity, good environmental stability, ease of preparation methods etc. Furthermore, PAni can be facilely oxidised and reduced; making it suitable for gas sensing application.9,10
The presence of –NH– group in PAni favours the oxidation, reduction and various physical and chemical properties.11–13 Besides the sensing capabilities, it can also be used in rechargeable batteries,4,14 supercapacitors,15 light emitting diodes,16 corrosion protection systems,17 and electrochromic display devices.18 But, PAni suffers from a major disadvantage of poor chemical solubility in organic solvents and also hurdles while operating at higher temperatures, limiting its applications. Most semiconducting metal oxide-based gas sensors (tin oxide, zinc oxide, nickel oxide, tungsten oxide and iron oxide etc.) are widely used as a chemiresistive type gas sensors, alternative to conducting polymers due to their ability to detect various gases, high sensitivity, fast response, recovery time and low cost. Most of the metal oxides are n-type semiconductors, because electrons are naturally produced due to adsorption of O2− during oxide formation.19 The major disadvantage of metal oxide gas sensors is the need of high temperature. For example, iron oxide (Fe2O3) used for nitrogen dioxide (NO2) sensing required >250 °C for a better selectivity,20 a CdO–Fe2O3 ethanol sensor activated at 300 °C.21 This demands a low-temperature controlled heating system to operate the sensor, as sensor operation at elevated temperature causes gradual changes in the physical and chemical properties of sensor, which promotes deviation in gas sensing properties of sensor with respect to time.
To overcome this difficulty, new prospects can be opened by designing the flexible sensor based on organic–inorganic composite nanostructures. The inorganic materials are thermally more stable as compared to organic i.e. polymer matrices and can improve the sensing characteristics such as sensitivity, selectivity, stability of polymer.22–25 The most important advantage of organic–inorganic composite nanostructures is that, by combining individual materials different property profiles leads to produce desirable properties in composites. The physicochemical properties of organic–inorganic hybrid nanocomposites depend on the distribution of filler particles into polymer matrix as well as on interfacial bonding between the inorganic filler and polymer matrix. Khuspe et al. reported that the incorporation of metal oxides into polymer matrix greatly modify the microstructure and the mechanical properties of original polymer26 and improves the sensing performance of sensor.27
In the present study, PAni was functionalized with α-Fe2O3 to acclimatize its structural properties and to boost its gas sensing performance in order to induce stronger and faster chemical reactions between PAni sensor and target gas. It is worth noticing that the properties of α-Fe2O3 not only make them a new class of composites, but also make them capabling for reinforcement of fillers within a host PAni matrix.28 The effect of α-Fe2O3 filler particles on gas sensing properties of PAni was studied in terms of change in sensor resistance, sensor response, response/recovery time and stability etc. We believe, the unique physical and chemical properties of PAni/α-Fe2O3 nanocomposite enabling sensing activity at lower concentration of NH3 gas. PAni/α-Fe2O3 composite flexible sensor was fabricated by in situ polymerization process using aniline as monomer. In situ polymerization is an excellent approach to form homogenous mixture of two materials. This is simple, low-cost and time saving process and a fine clad of different materials with conducting polymers can be designed on flexible substrates.29
2. Experimental details
2.1 Chemicals used
Ferric chloride hexahydrate (FeCl3)·6H2O purchased from Aldrich was used for preparation of iron oxide, methanol was purchased from s-d fine chemicals, aniline as monomer and ammonium persulfate as an oxidant, all are AR grade purchased and were used as received. Polyethylene terephthalate (PET) substrates of the dimensions of 60 mm × 20 mm with ±80 μm thickness were purchased and used as it is without any physical and processing.
2.2 Fabrication of flexible PAni/α-Fe2O3 sensor films
Chemically adapted PAni/α-Fe2O3 nanocomposite flexible sensors were prepared on flexible PET by in situ polymerization process using aniline as monomer and ammonium persulfate as an oxidant in HCl solution; resulting procedure is as follows; 0.1 M APS (ammonium persulphate) [(NH4)2S2O8] was added in 0.2 M aniline (C6H5NH2) and 0.2 M HCl (hydrochloric acid) mixture. The resulting mixture was kept at room temperature for 3 h followed by continuous stirring. After polymerization, green precipitate was occurred and is nothing but the emeraldine salt form of PAni. To get emeraldine base form of PAni; de-doping of salt form takes place with 0.1 M NH4OH solution and at same time PET substrates were vertically dipped in the beaker. α-Fe2O3 nanoparticles were synthesized using sol–gel method and incorporated into PAni matrix during de-doping of PAni from 10 to 50 wt% (1–5 mass%). After 3 h, blue thin layer of PAni deposited on the PET substrate. The resulting film was washed in distilled water to eliminate the low molecular weight organic intermediates and dried at the ambient temperature and used for further studies. The synthesis process for the development of PAni/α-Fe2O3 nanocomposite on PET substrate is illustrated thorough Scheme 1.26,29
|
| Scheme 1 Schematic representation of synthesis route of PAni/Fe2O3 nanocomposite on PET substrate by in situ oxidative polymerization. | |
2.3 Characterizations and gas sensing measurements
X-ray diffraction (XRD) pattern was recorded on Rigaku Ultima-IV diffractometer using Cu-Kα radiation with a wavelength 1.5406 Å in continuous scan mode at an acceleration voltage of 40 kV and current of 40 mA. The composite formation and bonding characteristics were investigated by a Nicolet IS-10 (Thermo Fisher Scientific Instruments) Fourier transform infrared spectroscopy (FTIR) spectra. The surface morphology of flexible nanocomposite sensor film was analysed by field-emission scanning electron microscopy digital photoimages (Model: MIRA3 TESCAN, USA). The interaction of α-Fe2O3 with PAni was confirmed by UV-visible spectroscopy. The effect of addition of α-Fe2O3 on polyaniline was investigated by Raman spectroscopy (Model: Horiba, 228 Lab Ram Hr 800). The electrochemical impedance spectroscopy (Precision Impedance Analyzer 6500B) measurement was used to quantify the sensor response to the gases tested. X-ray photoelectron spectroscopy (XPS, VG, Multilab 2000, Thermo VG, Scientific UK) spectra was carried out to investigate the structure of core-level electron and to resolve binding energies of PAni/α-Fe2O3 nanocomposite. Gas sensing properties were studied using a home-made gas sensor unit equipped with computer-attached KEITHLEY 6514 Electrometer. The NH3, nitrogen dioxide (NO2), methanol (CH3OH), ethanol (C2H5OH) and liquefied petroleum gases (LPG) etc., have been purchased from space cryo gases Pvt. Ltd, Mumbai. The resultant sensitivity of composite flexible sensor was expressed according to following equation defining the response. |
| (1) |
where Ra and Rg are resistances of film sensor in air and target gas, respectively.
3. Results and discussion
3.1 Structural identification
Fig. 1 shows the XRD patterns of as-synthesized PAni, α-Fe2O3 and PAni/α-Fe2O3 nanocomposite flexible films. From the XRD pattern of pure PAni, single broad peak (Fig. 1(a)) in the range of 20–30° (2θ) was evidenced which was higher in enclosed area and intensity, but in the case of composite i.e. PAni/α-Fe2O3, several new planes were evidenced. The broad nature of XRD pattern clearly indicated amorphous behavior of synthesized PAni. XRD pattern of α-Fe2O3 is shown in Fig. 1(b). It is noteworthy that high intense XRD peak of PAni at 25.82° is disappeared for composite structure. Shift in XRD peak of PAni in composite structure is eccentric. However, there are few reported wherein shift of PAni XRD peak position is evidenced when composite is formed,30 but in present case we believe it could be due to a change of its structure when composite is formed and obtained small intense PAni XRD peak might be a shoulder of previously found broad intense peak. The other peaks of PAni/α-Fe2O3 composite were similar to the peaks observed in XRD pattern of α-Fe2O3,27 confirming the type of α-Fe2O3 crystal structure which was well-maintained after preparing composite in base media. The XRD result showed the presence of two different phases i.e. PAni and α-Fe2O3, supporting for the formation of PAni/α-Fe2O3 nanocomposite.
|
| Fig. 1 XRD patterns of; (a) PAni, (b) α-Fe2O3, and (c) PAni/α-Fe2O3 nanocomposite films. | |
3.2 Surface analysis
To identify the presence of α-Fe2O3 in the composite and to observe the association between the PAni and the α-Fe2O3, FTIR spectra of flexible PAni and PAni/α-Fe2O3 nanocomposite films were recorded. The peak at 569 cm−1, characteristic of the Fe–O stretching vibration, clearly supported for the presence of α-Fe2O3 in PAni matrix and shift of peak suggested the interaction of α-Fe2O3 with polyaniline.31–33 The peak observed (Fig. 2(a)) at wave numbers 672 cm−1 and 736 cm−1 were attributed to C–H out of plane bending vibrations.34 The peak at 856 cm−1 was the characteristic of C–C and C–H for benzenoid ring occurred from C–H out of plane bending vibration of benzene ring.35–38 The absorption peak at 1061 cm−1 was due to C–H bending.35 Also the peak observed at 1283 cm−1 was on account of C–N stretching of primary aromatic amine ring and the peak observed at 1490 cm−1 was assigned to CN stretching in aromatic ring.39 The peak at 1442 cm−1 was in favour of CC stretching mode of benzenoid ring which was shifted to 1450 cm−1 on addition of α-Fe2O3.40,41 The absorption peak at 1591 cm−1 was due to C–C stretching vibration42 and peak at 1723 cm−1 was due to CO bond stretching vibration.43 Comparing the FTIR spectra of PAni and PAni/α-Fe2O3 composite, it was concluded that the most of the peaks of PAni were shifted towards higher wavenumber side i.e. PAni/α-Fe2O3 composite (Fig. 2(b)) which could be due to an interaction between PAni and α-Fe2O3 (ref. 35) as when nanoparticles of metal oxide are incorporated into a polymer matrix so as to obtain chemically, environmentally and thermally stable composite material, the constituents demonstrate strong intermolecular interactions, resulting in change of electron density distribution followed FTIR shift.27
|
| Fig. 2 (a and b) FTIR, and (c and d) Raman spectrums of PAni, and PAni/α-Fe2O3 nanocomposite film surfaces. | |
3.3 Raman shift analysis
Chemical structures of PAni and PAni/α-Fe2O3 nanocomposite flexible films were studied by Raman spectroscopy. Raman spectra of pure PAni (Fig. 2(c)) showed the characteristic band at 1175 cm−1 due to C–H stretching,44 1458 cm−1 due to CN vibration of quinoid ring45 and band at 1553 cm−1 was lined to N–H bending.46 Raman spectra of PAni/α-Fe2O3 (Fig. 2(d)) composite noted a characteristic band of α-Fe2O3 at 283 cm−1. Bands of pure PAni at 1458 cm−1 and 1553 cm−1 shifted in the vicinity of 1304 cm−1 and 1578 cm−1 when α-Fe2O3 was introduced. The shifting of the band was due to a strong mutual interaction of hydrogen bonding between Fe–O of α-Fe2O3 and –NH– group of PAni.
3.4 Surface morphological studies
Fig. 3 shows the FE-SEM digital micrographs of PAni, α-Fe2O3, and PAni/α-Fe2O3 nanocomposite flexible films prepared by in situ polymerization process. FE-SEM micrograph of PAni (Fig. 3(a)) presents well-grown interconnected fibres of nearly identical diameter, covered uniformly the surface of flexible film. All fibres were porous in surface morphology. Fig. 3(b) shows the FE-SEM micrograph of Fe2O3 film prepared from a powder annealed at 700 °C. The particles were elongated-type with few reminiscence of the hexagonal corundum structure of α-Fe2O3. α-Fe2O3 nanoparticles were embedded within the net like structure consisting of PAni fibres in the case of PAni/α-Fe2O3 composite, suggesting a significant change in the original individual morphologies. As a consequence, PAni/α-Fe2O3 composite film (Fig. 3(d)) possessed both well-defined forms i.e. PAni and α-Fe2O3, implying a highly micro-porous structure and providing a path for insertion and extraction of gas molecules and ensuring high rate of reaction.
|
| Fig. 3 FESEM micrographs of; (a) PAni, (b) α-Fe2O3, and (c) PAni/α-Fe2O3 composite films. | |
3.5 Surface analysis
An XPS spectrum was employed to analyse the effect α-Fe2O3 and to understand the different electronic structures and chemical bonding information of PAni/α-Fe2O3 composite. XPS core level spectrum of PAni/α-Fe2O3 nanocomposite demonstrated C1s, O1s, N1s and Fe2p peaks, which were related to the binding energies of C, O, N, and Fe, respectively. Fig. 4(a) presents the C1s core-level spectrum of PAni/α-Fe2O3 composite; deconvoluted into three components at binding energies 283.8 eV, 285.85 eV and 287.7 eV, respectively. Peak assigned at binding energy 283.8 eV was due to C–H band.47 The peaks observed at binding energy 285.85 eV and 287.7 eV were attributed to C–N and CO band, respectively.47 Core-level O1s in XPS spectrum of PAni/α-Fe2O3 nanocomposite highlighted in Fig. 4(b). The spectrum deconvoluted into three components; peak observed at lower binding energy (530 eV) was the characteristic peak of α-Fe2O3, which was attributed to O–H band.48,49 The peak observed at 531.9 eV was assigned to surface oxygen due to chemisorbed water49,50 and the peak evidenced at higher binding energy (533.4 eV) was due to oxygen within the Fe–O–H bond because the estimated binding energy of O1s electron for adsorbed water is 533 eV.49 N1s core-level spectrum of PAni/α-Fe2O3 nanocomposite is shown in Fig. 4(c). The N1s core-level spectrum was situated around a binding energy of 399.4 eV, indicating the existence of benzenoid ring in the PAni/α-Fe2O3 nanocomposite as well as the prepared PAni could be in the neutral media with emeraldine form.51,52 Fig. 4(d) shows the core-level spectrum of Fe with two peaks located at binding energies of 710.85 eV (Fe3+ state) and 724.35 eV, respectively, and were well-matched with values reported in literature.49,53 These observed bands indicated the presence of inorganic phase of α-Fe2O3 in the composite.49,54
|
| Fig. 4 (a) C1s, (b) O1s, (c) N1s, and (d) Fe2p of PAni/α-Fe2O3 composite. | |
3.6 Gas sensing studies
3.6.1 Selectivity. High selectivity is an important parameter for excellent working of sensor. To confirm the selectivity of the sensor to a particular gas, chemo-resistive properties were studied for PAni and PAni/α-Fe2O3 nanocomposite sensors by exposing sequentially for different gases. Selectivity study of PAni/α-Fe2O3 nanocomposite sensor was performed for NH3, H2S, C2H5OH, CH3OH, and NO2 at 100 ppm concentration of each gas and given in Fig. 5(a). Resistance of PAni film sensor was not changed to C2H5OH and CH3OH gases, hence, the response was considered to be zero, confirming that these gases are not interacting with the PAni sensor. But when PAni sensor was exposed to NO2 and H2S gases, responses were increased by 1% and 0.5%, respectively; considerably lower as compared to NH3 (26%) gas. Selectivity study clearly supported that, PAni/α-Fe2O3 sensor was more selective to NH3 than other test gases and exhibited higher selectivity to NH3 when compared to pure PAni. After addition of α-Fe2O3 nanoparticles in PANI matrix, quinoid structure (higher conjugation) of PAni was transformed into benzoid structure (lower conjugation) and made it more selective to target gas (herein NH3).5
|
| Fig. 5 (a) Selectivity study of PAni and PAni/α-Fe2O3 sensor films, (b) change in resistance as a function of time of PAni/α-Fe2O3 sensor film upon 100 ppm exposure of NH3 gas, and (c) response of flexible composite sensor to different NH3 concentrations. | |
3.6.2 NH3 dependent response of PAni/α-Fe2O3 flexible sensor. NH3 gas sensing behaviour of PAni/α-Fe2O3 nanocomposite flexible sensor was determined by monitoring the electrical resistance. At first PAni/α-Fe2O3 sensor was exposed to air to measure the sensor resistance in air (Ra). Fig. 5(b) shows the response of flexible PAni/α-Fe2O3 nanocomposite sensor to NH3 gas at room temperature. From the response graph, it was observed that the sensor resistance was increased immediately with time when sensor exposed to NH3 gas. At the beginning when NH3 gas was injected into the chamber, the measured resistance of sensor increased rapidly, reached to its maximum value and became stable, which was used to define the resistance of sensor in gas (Rg). Recovery of sensor was achieved by opening the lead of chamber, leading to a strong gradual decrease of the measured resistance of PAni/α-Fe2O3 nanocomposite sensor until reaching to its initial value Ra. Decrease in resistance of sensor after exposure to air was due to a decrease of NH3 concentration.55 According to eqn (1), the response obtained for PAni/α-Fe2O3 nanocomposite flexible sensor was 39% for 100 ppm level of NH3 gas with short response (27 s) and (46 s) recovery times, defined as the time from exposure to NH3 to the peak in resistance and the time from exposure to pure air till the previous resistance, respectively.
3.6.3. Gas sensing mechanism. The gas sensing mechanism for PAni/α-Fe2O3 flexible sensor was straightforwardly concluded to stem from N–H group of PAni, by deprotonation and protonation.56 Fig. 5(b) shows that the resistance of PAni/α-Fe2O3 nanocomposite flexible sensor is increased on exposure to NH3 (electron donor) gas. Similar behaviour of PAni–metal oxide composite sensors for NH3 gas detection is found in literature for PAni–TiO2 (Tai et al.12), PAni–ZnO (Khan et al.13), PAni–SnO2 (Khuspe et al.27 and Geng et al.7) sensors. Being NH3 as a reducing agent, for p-type polyaniline fibres, resistance was increased upon interaction.During the process NH3, as a deprotonating agent, might interact with H+ ion on the PAni backbone, withdrawing protons from N–H sites followed transferring them to ammonium (NH4+) ions13 as shown in Scheme 2. After attaining optimum we found there was decrease in resistance before switching gas to its off level within 50–75 s, which was unusual, and might arise from an interaction of NH3 gas molecules with host material i.e. as NH3 is electron donating moiety thereby when it interacts with composite thin film which is p-type in nature; shift in its Fermi energy level towards conduction band position side with reduced resistivity is anticipated. The increase in electrical resistance was attributed to proton charge transfer from PAni to NH3 gas.9,10 Due to loss of H+ ion, surface depletion region was increased, thereby; the conductivity of PAni was decreased. When sensor film was exposed to air; could capture hydrogen from NH4+ and renovate the initial doping level of PAni.9,10,13 This added H+, form N–H bond and thus the resistance of sensor was decreased. However, the mechanism for NH3 gas molecules and α-Fe2O3 interactions in composite form was unprecedented. Our previous results showed that, the resistance of α-Fe2O3 remained unchanged upon exposure to NH3 gas,57 suggesting NH3 can be an inert gas for α-Fe2O3.
|
| Scheme 2 Possible sensing mechanism of NH3 gas with PAni/α-Fe2O3 film sensor. | |
3.6.4 Influence of NH3 concentration. The performance of PAni/α-Fe2O3 flexible sensor was determined by measuring the sensor response for different ammonia gas concentrations at room temperature (30 °C) (Fig. 5(c)), revealing to be sensitive as low as 5 ppm of NH3. From Fig. 5(c) it is revealed that, the response of PAni/α-Fe2O3 sensor was increased with rising NH3 gas concentration. At higher concentration of ammonia (100 ppm) it showed not only the maximum response (39%) but also very short response time (27 s), with increase of concentration the number of gas molecules interacting with sensor surface increases as well.
3.6.5 Reproducibility, response–recovery, stability and impedance studies. Reproducibility study of PAni/α-Fe2O3 nanocomposite flexible sensor revealed that the sensor maintained its initial amplitude with short response and recovery times for three cycles of NH3 gas exposure, indicating good stability of sensor [Fig. 6(a)]. Complete recovery of sensor resistance to its initial baseline value after exposition to air, evidencing favourable industrial and market quality of developed sensor. This was due to protonation and deprotonation of polyaniline. Fig. 6(b) shows response vs. recovery time graph of PAni/α-Fe2O3 flexible sensor to different NH3 gas concentrations. Very short response time (27 s) and recovery time (46 s) were noted at 100 ppm of NH3 gas concentration. Whereas, at the lowest concentration of NH3 gas (5 ppm), 76 s response time and 5 s recovery time were recorded, giving an indicator for the speed of the adsorption–desorption of NH3 gas on the sensor. From the response–recovery observations, it was evident that with increase of the NH3 gas concentration, response time was decreased drastically and recovery time was increased intensively; faster than previously reported literature values.35,58 The stability of the sensor was determined by measuring the response of the sensor for every five days for one month [Fig. 6(c)] by fixing the constant concentration (100 ppm) of NH3 gas, showing a decrease of response for the few first days until reach to an equilibrium level i.e. response after 15 days. Hence, PAni/α-Fe2O3 flexible sensor showed good (74%) long-term stability. The impedance spectra of PAni/α-Fe2O3 flexible sensor was measured in air and gas at room temperature and covers in Fig. 6(d). Impedance was measured in the frequency range of 20 Hz to 10 MHz; the plot of imaginary component (Z′′) against the real component (Z′) exhibited a perfect semicircle. The series resistance Rs was found by the value of real impedance at the highest frequency intercept. R1 and C1 component in the equivalent circuit correspond to the bulk properties while R2 and C2 component correspond to the grain boundary properties of the PAni/α-Fe2O3. Table 1 summarizes the values of impedance parameter in air and gas. When the gas was injected in to the chamber resistance R1 was increased and capacitance C1 was decreased.
|
| Fig. 6 (a) Reproducibility, (b) response vs. recovery variation, (c) stability of PAni/α-Fe2O3 nanocomposite sensor, and (d) impedance spectra (Z′ vs. Z′′) of PAni/α-Fe2O3 in presence of air and NH3 gas. Inset shows an equivalent circuit. | |
Table 1 Impedance parameters for PAni/α-Fe2O3 flexible nanocomposite sensor
Medium |
Parameter |
R0 (kΩ) |
R1 (kΩ) |
R2 (kF) |
C1 (nF) |
C2 (nF) |
Air |
3 |
109.80 |
216.66 |
1.23 |
0.21 |
Gas |
3 |
199.13 |
254.930 |
1.15 |
2.02 |
4. Conclusions
PAni/α-Fe2O3 composite ammonia sensor was successfully developed on PET substrate by in situ polymerization method. The resulting composite sensor was a mixture of fibers PAni and crystallites of α-Fe2O3. There was strong coupling between PAni and α-Fe2O3 as well as the formation of composite of PAni and α-Fe2O3. PAni/α-Fe2O3 nanocomposite flexible sensors was highly selective to NH3 gas at room temperature compared to other test gases. The sensor response was measured with the change of resistance which was linearly increased with an increase of NH3 concentration from 5 ppm to 100 ppm. The NH3 sensor showed highest response of 39% with a fast 27 s response time and 47 s recovery time at 100 ppm of NH3 gas. Impedance spectra measured in air and gas and was well-fitted by with two time constants in series. Due to the room temperature operation and low development cost, flexible substrate, developed composite sensor can find numerous applications in domestic and industrial market areas from the points of safely, ecology and economy.
Acknowledgements
Dr V. B. Patil would like thank DAE-BRNS, for financial support through the scheme no. 2010/37P/45/BRNS/1442. FJS would like to acknowledge Nanshan District Key Lab for Biopolymers and Safety Evaluation (No. KC2014ZDZJ0001A). The authors would like to extend their sincere appreciation to the Deanship of Scientific Research at King Saud University for funding this work through the Research Group No. RG-1436-034.
References
- G. Rizzo, A. Arena, N. Donato, M. Latino, G. Saitta, A. Bonavita and G. Neri, Thin Solid Films, 2010, 518, 7133 CrossRef CAS PubMed.
- S. G. Pawar, S. L. Patil, M. A. Chougule, B. T. Raut, S. A. Pawar, R. N. Mulik and V. B. Patil, J. Mater. Sci.: Mater. Electron., 2012, 23, 273 CrossRef CAS.
- A. Arena, N. Donato, G. Saitta, A. Bonavita, G. Rizzo and G. Neri, Sens. Actuators, B, 2010, 145, 488 CrossRef CAS PubMed.
- X. B. Yan, Z. J. Han, Y. Yang and B. K. Tay, Sens. Actuators, B, 2007, 23, 107 CrossRef PubMed.
- G. D. Khuspe, D. K. Bandgar, S. Sen and V. B. Patil, Synth. Met., 2012, 162, 1822 CrossRef CAS PubMed.
- S. T. Navale, A. T. Mane, G. D. Khuspe, M. A. Chougule and V. B. Patil, Synth. Met., 2014, 195, 228 CrossRef CAS PubMed.
- L. Geng, Y. Zhao, X. Huang, S. Wang, S. Zhang and S. Wu, Sens. Actuators, B, 2007, 120, 568 CrossRef CAS PubMed.
- S. T. Navale, A. A. Ghanwat and V. B. Patil, Measurement, 2014, 50, 363 CrossRef PubMed.
- R. G. Bavane, A. M. Mahajan, M. D. Shirsat and R. B. Gore, J. Adv. Phys., 2013, 3, 241 Search PubMed.
- M. Matsuguchi, J. Io, G. Sugiyama and Y. Sakai, Synth. Met., 2002, 128, 15 CrossRef CAS.
- D. N. Debarnot and F. P. Epaillard, Anal. Chim. Acta, 2003, 475, 1 CrossRef.
- H. Tai, Y. Jiang, G. Xie, J. Yu and X. Chen, Sens. Actuators, B, 2007, 125, 644 CrossRef CAS PubMed.
- A. A. Khan and M. Khalid, J. Appl. Polym. Sci., 2010, 117, 1601 CAS.
- A. Mirmohseni and R. Solhjo, Eur. Polym. J., 2003, 39, 219 CrossRef CAS.
- M. M. Khandpekar, R. K. Kushwaha and S. P. Pati, Solid-State Electron., 2011, 62, 156 CrossRef CAS PubMed.
- E. Menefee and Y. H. Pao, J. Chem. Phys., 1962, 35, 3472 CrossRef PubMed.
- C. K. Chiang, S. C. Gau, C. R. J. Fincher, Y. W. Park, A. G. MacDiarmid and A. J. Heeger, Appl. Phys. Lett., 1978, 33, 18 CrossRef CAS PubMed.
- R. Dupon, D. H. Whitmore and D. F. Shriver, J. Electrochem. Soc., 1981, 128, 715 CrossRef CAS PubMed.
- S. L. Patil, S. G. Pawar, A. T. Mane, M. A. Chougule and V. B. Patil, J. Mater. Sci.: Mater. Electron., 2010, 21, 1332 CrossRef CAS.
- G. Neria, A. Bonavitaa, S. Galvagnoa, P. Sicilianob and S. Capone, Sens. Actuators, B, 2002, 82, 40 CrossRef.
- X. Liu, Z. Xu, Y. Liu and Y. Shen, Sens. Actuators, B, 1998, 52, 270 CrossRef CAS.
- M. A. D. Paoli, R. J. Waltman, A. F. Diaz and J. Bargon, J. Chem. Soc., Chem. Commun., 1984, 15, 1015 RSC.
- S. E. Lindsey and S. B. Street, Synth. Met., 1984, 10, 67 CrossRef CAS.
- C. Cassignol, M. Cavarero, A. Boudet and A. Ricard, Polymer, 1999, 40, 1139 CrossRef CAS.
- S. G. Pawar, S. L. Patil, M. A. Chougule, B. T. Raut and P. R. Godse, IEEE Sens. J., 2011, 11, 2980 CrossRef CAS.
- S. Bai, Y. Zhao, J. Sun, Y. Tian, R. Luo, D. Li and A. Chen, Chem. Commun., 2012, 1 Search PubMed.
- G. D. Khuspe, S. T. Navale, D. K. Bandgar, R. D. Sakhare, M. A. Chougule and V. B. Patil, Electron. Mater. Lett., 2014, 10, 191 CrossRef CAS.
- Y. Wu, S. Xing, S. Jing, T. Zhou and C. Zhao, e-Polym., 2007, 103, 1 Search PubMed.
- J. Stejskal and I. Sapurina, Pure Appl. Chem., 2005, 77, 815 CrossRef CAS.
- S. Wang, Y. Hua, R. Zonga, Y. Tanga, Z. Chenb and W. Fana, Appl. Clay Sci., 2004, 25, 49 CrossRef CAS PubMed.
- C. Yang, J. Du, Q. Peng, R. Qiao, W. Chen, C. Xu, Z. Shuai and M. Gao, J. Phys. Chem. B, 2009, 113, 5052 CrossRef CAS PubMed.
- V. A. Khatil, S. B. Kondawar and V. A. Tabhane, Anal. Bioanal. Electrochem., 2011, 3, 614 Search PubMed.
- K. R. Reddy, K. P. Lee and A. I. Gopalan, J. Appl. Polym. Sci., 2007, 106, 1181 CrossRef CAS PubMed.
- R. Hawaldar, M. Kulkarni, S. Jadkar, U. Pal and D. Amalnerkar, J. Nano Res., 2009, 5, 79 CrossRef CAS.
- J. Husain, S. B. Chacradhar and M. V. N. Ambika Prasad, Int. J. Eng. Res. Ind. Appl., 2014, 4, 198 Search PubMed.
- N. G. Deshpande, Y. G. Gudage, R. Sharma, J. C. Vyas, J. B. Kim and Y. P. Lee, Sens. Actuators, B, 2009, 138, 76 CrossRef CAS PubMed.
- A. H. Elsayed, M. S. MohyEldin, A. M. Elsyed, A. H. Abo Elazm, E. M. Younes and H. A. Motaweh, Int. J. Electrochem. Sci., 2011, 6, 206 CAS.
- P. Kunzo, P. Lobotka, E. Kovacova, K. Chrissopoulou, L. Papoutsakis, S. H. Anastasiadis, Z. Krizanova and I. Vavra, Phys. Status Solidi A, 2013, 210, 2341 CrossRef CAS PubMed.
- J. Vivekanandan, V. Ponnusamy, A. Mahudeswaran and P. S. Vijayanand, Arch. Appl. Sci. Res., 2011, 3, 147 CAS.
- A. Khan, A. S. Aldwayyan, M. Alhoshan and M. Alsalhi, Polym. Int., 2010, 59, 1690 CrossRef CAS PubMed.
- T. H. Hsieh, K. S. Ho, C. H. Huang, Y. Z. Wang and Z. L. Chen, Synth. Met., 2006, 156, 1355 CrossRef CAS PubMed.
- S. T. Navale, G. D. Khuspe, M. A. Chougule and V. B. Patil, J. Phys. Chem. Solids, 2014, 75, 236 CrossRef CAS PubMed.
- M. R. Saboktakin, A. Maharramov and M. A. Ramazanov, Nat. Sci., 2007, 5, 67 Search PubMed.
- B. E. J. Tabares, F. J. Isaza and S. I. Cordoba de Torresi, Mater. Chem. Phys., 2012, 132, 529 CrossRef PubMed.
- R. Mazeikiene, G. Niaura and A. Malinauskas, Electrochim. Acta, 2006, 51, 5761 CrossRef CAS PubMed.
- R. Mazeikiene, V. Tomkute, Z. Kuodis, G. Niaura and A. Malinauskas, Vib. Spectrosc., 2007, 44, 201 CrossRef CAS PubMed.
- M. G. Han and S. S. Im, Polymer, 2000, 41, 3253 CrossRef CAS.
- G. Wang, Y. Ling, D. A. Wheeler, K. E. N. George, K. Horsley, C. Heske, J. Z. Zhang and Y. Li, Nano Lett., 2011, 11, 3503 CrossRef CAS PubMed.
- P. Mills and J. L. Sullivan, J. Phys. D: Appl. Phys., 1983, 16, 723 CrossRef CAS.
- G. Wu, X. Tan, G. Li and C. Hu, J. Alloys Compd., 2010, 504, 371 CrossRef CAS PubMed.
- G. Qiu, Q. Wang and M. Nie, J. Appl. Polym. Sci., 2006, 102, 2107 CrossRef CAS PubMed.
- L. Konga, X. Lub and W. Zhang, J. Solid State Chem., 2008, 181, 628 CrossRef PubMed.
- S. Liu, D. Tao and L. Zhang, Powder Technol., 2012, 217, 502 CrossRef CAS PubMed.
- A. Z. Sadek, W. Wlodarski, K. Shin, R. B. Kaner and K. Kalantar-zadeh, Nanotechnology, 2006, 17, 4488 CrossRef CAS.
- S. K. Shukla, N. B. Singh and R. P. Rastogi, Indian J. Eng. Mater. Sci., 2013, 20, 319 CAS.
- R. Arora, A. Srivastav and U. K. Mandal, International Journal of Modern Engineering Research, 2012, 2, 2384 Search PubMed.
- S. T. Navale, D. K. Bandgar, S. R. Nalage, G. D. Khuspe, M. A. Chougule, Y. D. Kolekar, S. Sen and V. B. Patil, Ceram. Int., 2013, 39, 6453 CrossRef CAS PubMed.
- S. Srivastava, S. Kumar and Y. K. Vijay, Int. J. Hydrogen Energy, 2012, 37, 3825 CrossRef CAS PubMed.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: 10.1039/c5ra11512d |
|
This journal is © The Royal Society of Chemistry 2015 |
Click here to see how this site uses Cookies. View our privacy policy here.