DOI:
10.1039/C5RA13976G
(Paper)
RSC Adv., 2015,
5, 79729-79737
Copper(II) phthalocyanine supported on a three-dimensional nitrogen-doped graphene/PEDOT-PSS nanocomposite as a highly selective and sensitive sensor for ammonia detection at room temperature†
Received
15th July 2015
, Accepted 9th September 2015
First published on 11th September 2015
Abstract
Here we present a highly efficient ammonia (NH3) gas sensor made of copper(II) tetrasulfophthalocyanine supported on a three-dimensional nitrogen-doped graphene based framework (CuTSPc@3D-(N)GF)/poly(3,4-ethylenedioxythiophene)-poly(styrenesulfonate) (PEDOT-PSS) nanocomposite sensing film with high uniformity over a large surface area. The NH3 gas sensing performance of the nanocomposite was compared with those of sensors based on pure PEDOT-PSS and a pristine CuTSPc@3D-(N)GF. It was revealed that the synergistic behavior between both of these candidates allowed excellent sensitivity and selectivity for NH3 gas in a low concentration range of 1–1000 ppm at room temperature. The CuTSPc@3D-(N)GF/PEDOT-PSS nanocomposite gas sensor exhibited a much better response (∼5 and 53 times, with a concentration of NH3 gas at 200 ppm) to NH3 gas than those of the pure PEDOT-PSS and pristine CuTSPc@3D-(N)GF gas sensors, respectively. The combination of the CuTSPc@3D-(N)GF and PEDOT-PSS facilitated the enhancement in the sensing properties of the final nanocomposite and paved a new avenue for the application of CuTSPc@3D-(N)GF/PEDOT-PSS nanocomposites in the gas sensing field.
Introduction
Graphene, as a two-dimensional sp2 bonded carbon sheet, has attracted much attention in many diverse applications, especially chemical sensors,1–8 due to its excellent electronics, high mechanical stiffness and specific surface-to-volume ratio, as well as its superior conductivity.9–12 Three-dimensional (3D) graphene-based frameworks (3D-GFs) such as sponges, foams and aerogels are an important class of new-generation porous carbon materials, which exhibit high porosities, large surface areas and high electrical conductivities.13–17 These materials can serve as strong matrices for functionalizing metals, metal oxides and electrochemically active polymers for various applications in electrochemical capacitors,18–20 batteries21,22 and catalysis.23–25 Dong et al.26 have demonstrated that a 3D graphene electrode, as an electrochemical sensor for detection of dopamine, exhibits remarkable sensitivity (619.6 μA mM−1 cm−2) and a low detection limit (25 nM at a signal-to-noise ratio of 5.6), with a linear response of up to ∼25 μM. Therefore, 3D-GFs can provide a promising platform for the development of high performance electrochemical sensors for dangerous volatile organic compounds (VOCs).
Poly(3,4-ethylenedioxythiophene)-poly(styrenesulfonate) (PEDOT-PSS) as a conjugated polymer (a mixture of two ionomers) has been extensively studied as the active material in sensing applications because of its good electrical conductivity, high transparency, low redox potential and good processability.27–29 Nevertheless, its limited chemical and structural properties prevent its use in various practical applications, especially electrochemical sensing.30 Its composite with carbon nanostructures could be a promising solution to the shortcomings. Jian et al.31 have used a PEDOT-PSS composite film with O2 plasma-treated single-walled carbon nanotubes for the detection of ammonia (NH3) and trimethylamine gases. Seekaew et al.27 have reported the NH3 sensing behavior of a graphene–PEDOT-PSS composite film at room temperature. This evidence indicates that PEDOT-PSS composite films show potential as useful sensing materials, but their low sensitivities restrict their application in practical VOC sensors.
As a result of the above-mentioned reasons, and also our interest in the synthesis of graphene based materials for various applications,25,32–34 especially as new sensors,35–37 herein, we demonstrate the use of a copper(II) tetrasulfophthalocyanine supported on a three-dimensional nitrogen-doped graphene-based framework (CuTSPc@3D-(N)GF) and PEDOT-PSS nanocomposite as a novel gas sensor. This new architecture holds great appeal as a chemical sensor due to its large surface area, 3D multiplexed network of highly conductive pathways and continuously interconnected macroporous structures, as well as its modified active surface. To demonstrate its potential, we used it here for the detection of NH3, a highly toxic gas, which leads to irritation of the skin, eyes and respiratory tract of humans.
Experimental
Preparation of CuTSPc@3D-(N)GF/PEDOT-PSS nanocomposite
The CuTSPc@3D-(N)GF was synthesized based on our previous work.25 The obtained CuTSPc@3D-(N)GF was dispersed in DI water (∼1 mg ml−1) and mildly sonicated for 30 min in a bath sonicator (EUROSONIC® 4D, 50 kHz). The PEDOT-PSS aqueous solution (weight ratio = 1–6, Clevios™ P VP AI 4083, solid content 1.3–1.7%) was first dissolved in DI water with a weight concentration of 89.82%. The CuTSPc@3D-(N)GF dispersion with 6 wt% dimethyl sulfoxide (DMSO, Sigma-Aldrich Co) was added to the PEDOT-PSS solution and a homogeneous aqueous dispersion was obtained after 2 h of stirring and sonication for 30 min.
Fabrication of the CuTSPc@3D-(N)GF/PEDOT-PSS nanocomposite gas sensor
The nanocomposite gas sensor was prepared from the PEDOT-PSS and CuTSPc@3D-(N)GF materials with chemical structures schematically illustrated in Fig. 1. For the fabrication of the CuTSPc@3D-(N)GF/PEDOT-PSS nanocomposite gas sensor, interdigitated Au electrodes with 100 nm thickness were deposited on a SiO2/Si substrate (10 × 4 mm2) using a physical vapor deposition method. The prepared CuTSPc@3D-(N)GF/PEDOT-PSS nanocomposite solution was then drop casted over the interdigitated electrode (Fig. 2a). The width and inter-spacing of the electrodes were 200 μm and 400 μm, respectively. Then, the nanocomposite gas sensor was heated for 1 h in a furnace (Exciton, EX1200-4L) at 80 °C under a nitrogen atmosphere (Fig. 2b). The pristine CuTSPc@3D-(N)GF and PEDOT-PSS gas sensors were also fabricated and tested for comparison. The fabricated CuTSPc@3D-(N)GF/PEDOT-PSS nanocomposite gas sensor is displayed in Fig. 2c and d.
 |
| Fig. 1 Schematic structures of (a) PEDOT-PSS and (b) the CuTSPc@3D-(N)GF. | |
 |
| Fig. 2 Schematic steps of the gas sensor fabrication process. | |
Characterization methods
Transmission electron microscopy (TEM) was performed using a LEO 912AB electron microscope operated at an accelerating voltage of 120 kV. Scanning electron microscopy (SEM) was performed using a S-4160 electron microscope. Atomic force microscopy (AFM) images were obtained using a tapping mode with an ARA AFM (0201/A, Ara research Co, Iran). Brunauer–Emmett–Teller (BET) surface area measurements were obtained through nitrogen adsorption at 77 K using an ASAP2020 instrument. The conductivities of the sensing films were measured using a 4-point technique probe with 10 nA of applied current. The resistances of the sensors were measured in a closed steel chamber (lab-made) with a LCR meter (Pintek-LCR900) and vapor gas flows were injected into the closed steel chamber by a mass flow meter (Alicat scientific, Tucson, USA). The reference humidity and sensor temperature were monitored by PT100 and HIH4000, respectively. The responses and selectivities of the gas sensors towards NH3, methanol, ethanol, acetone, toluene, chlorobenzene, and water, with gas concentrations ranging from 1 ppm to 1000 ppm at room temperature, were then assessed using a standard flow-through method. A constant flux of synthetic air of about 50 cm3 min−1 was mixed with the NH3 gas source at different flow rate ratios to desired concentrations using mass flow controllers. All experiments were performed at room temperature (25 ± 2 °C) and a relative humidity of 10 ± 2%. The sensitivity was defined by the following equation: |
 | (1) |
withwhere R0 and Rgas are the resistances of the sensor in synthetic air and the test gas, respectively.
Results and discussion
The surface morphology of the CuTSPc@3D-(N)GF is presented in the SEM and TEM images of Fig. 3a and b, respectively. Fig. 3a shows the 3D morphology and the interconnected porous structure of the ultrathin graphene nanosheets. Moreover, the pore sizes range from a few hundred nanometers to several micrometers. Fig. 3b not only clearly shows the presence of the mesopores in the carbon walls, but also reveals the wrinkled paper-like texture of the sheets, consistent with previous reports.38 In addition, based on the BET method, the synthesized support (3D-(N)GFs) from our previous work has a high surface area (up to 266.0 m2 g−1) for the incorporation of CuTSPc as active sensing sites. The chemical compositions of the 3D-(N)GF and CuTSPc@3D-(N)GF were confirmed by several characterization methods, such as elemental analysis, Fourier transform infrared spectroscopy, thermogravimetric analysis, X-ray powder diffraction and X-ray photoelectron spectroscopy.25
 |
| Fig. 3 Microstructure of the as-prepared CuTSPc@3D-(N)GF: (a) SEM and (b) TEM images. | |
Fig. 4a and b show the surface morphologies of the drop-coated pure PEDOT-PSS and CuTSPc@3D-(N)GF/PEDOT-PSS nanocomposite film, respectively. As can be seen, the pure PEDOT-PSS has a relatively smooth surface (Fig. 4a). On the other hand, Fig. 4b shows that the CuTSPc@3D-(N)GF is uniformly dispersed through the PEDOT-PSS matrix without obvious agglomeration. This was ascribed to the hydrophilic nature of CuTSPc on the surface of the 3D-(N)GF which not only ensured the strong bonding with PEDOT-PSS but also improved the dispersity and stability of the CuTSPc@3D-(N)GF in aqueous solution (Fig. S1†). Further roughness evaluations by AFM show that the average surface roughness (Ra) of the pure PEDOT-PSS films is 1.98 nm, while the Ra of the CuTSPc@3D-(N)GF/PEDOT-PSS is 7.52 nm (Fig. S2†). The much larger Ra of the CuTSPc@3D-(N)GF/PEDOT-PSS nanocomposite film compared to the pure PEDOT-PSS film suggests a significant enhancement in the active surface-area for gas adsorption by the CuTSPc@3D-(N)GF.25,27
 |
| Fig. 4 SEM images of the pure PEDOT-PSS (a) and CuTSPc@3D-(N)GF/PEDOT-PSS nanocomposite (b) films. | |
Fig. 5a shows the dynamic response of the CuTSPc@3D-(N)GF/PEDOT-PSS nanocomposite gas sensor towards 1000 ppm NH3 at room temperature in air. It shows that the sensor has good repeatability in terms of its response towards repeated NH3-sensing cycles at room temperature. The sensitivity of the nanocomposite gas sensor increases upon exposure to NH3 and recovers to the initial value upon the removal of NH3 in air. As shown in Fig. 5b, the sensitivity of the CuTSPc@3D-(N)GF/PEDOT-PSS nanocomposite gas sensor increases dramatically when exposed to various concentrations of NH3 ranging from 200 ppm to 800 ppm, and recovers to the original values when the NH3 is replaced by air. The changing sensitivity behavior may be ascribed to the adsorption and desorption of NH3 molecules on the nanocomposite sensing film.27 The details of the sensing mechanism for the CuTSPc@3D-(N)GF/PEDOT:PSS nanocomposite gas sensor will be discussed in further. In addition, the conductivities of the pure PEDOT-PSS and CuTSPc@3D-(N)GF/PEDOT-PSS nanocomposite sensing films are 650 and 1430 S cm−1, respectively. This indicates that the conductivity of the pure PEDOT-PSS is increased by more than double, due to a significant increase in the charge carrier concentration as a result of CuTSPc@3D-(N)GF incorporation. Therefore, CuTSPc@3D-(N)GF exhibits a dominant effect in the charge transport through the PEDOT-PSS matrix.
 |
| Fig. 5 (a) Dynamic responses of the CuTSPc@3D-(N)GF/PEDOT-PSS nanocomposite gas sensor to 1000 ppm NH3 and (b) the sensitivity of the CuTSPc@3D-(N)GF/PEDOT-PSS nanocomposite gas sensor when exposed to different concentrations of NH3 at room temperature. | |
PEDOT-PSS based gas sensors usually operate at rather low temperatures with respect to gas sensors based on metal oxides.27,35,39 It can be seen that the CuTSPc@3D-(N)GF/PEDOT-PSS nanocomposite gas sensor shows lower sensitivity to NH3 with an increase of temperature (Fig. 6). Therefore, the optimal temperature of the CuTSPc@3D-(N)GF/PEDOT-PSS nanocomposite gas sensor for NH3 detection was found to be room temperature. Since the interactions between the CuTSPc@3D-(N)GF/PEDOT-PSS nanocomposite sensing film and gaseous volatile organic compounds (VOCs) are exothermic, the activation energy for the desorption of NH3 molecules from the nanocomposite sensing film is larger than that for the adsorption. This reveals that the decrease in sensitivity at higher temperatures is a result of the higher desorption rate for NH3 gas.
 |
| Fig. 6 The sensitivity of the CuTSPc@3D-(N)GF/PEDOT-PSS nanocomposite gas sensor as a function of temperature towards 200 ppm NH3. | |
The response times of the pure PEDOT-PSS and CuTSPc@3D-(N)GF/PEDOT-PSS nanocomposite gas sensors, when they were experiencing a 95% resistance change, were estimated to be ∼6 min and ∼2.5 min, respectively. Moreover, the recovery times of the pure PEDOT-PSS and CuTSPc@3D-(N)GF/PEDOT-PSS nanocomposite gas sensors were ∼2.5 min and ∼1 min, respectively. Therefore, the CuTSPc@3D-(N)GF/PEDOT-PSS nanocomposite gas sensor exhibits relatively short response and recovery times compared with the PEDOT-PSS one (see Fig. 7).
 |
| Fig. 7 The responses and recoveries of gas sensors based on (a) pure PEDOT-PSS and (b) the CuTSPc@3D-(N)GF/PEDOT-PSS nanocomposite with 200 ppm NH3 at room temperature. | |
Fig. 8 demonstrates the sensitivities of the pure PEDOT-PSS, pristine CuTSPc@3D-(N)GF and CuTSPc@3D-(N)GF/PEDOT-PSS nanocomposite gas sensors as a function of NH3 concentration at room temperature. At a 1000 ppm NH3 concentration, the sensitivities of the pure PEDOT-PSS, pristine CuTSPc@3D-(N)GF and CuTSPc@3D-(N)GF/PEDOT-PSS gas sensors were 14.8%, 9% and 91%, respectively. At low concentrations (50 ppm), the sensitivities for the mentioned gas sensors were 4%, 0.35% and 9%, respectively (inset of Fig. 8). As can be seen, the room temperature sensitivity of pure PEDOT-PSS was lower than that of the CuTSPc@3D-(N)GF, but its response was substantially increased after CuTSPc@3D-(N)GF incorporation. Thus, the CuTSPc@3D-(N)GF improves the interaction with NH3, leading to a higher charge reduction only when it is included in the PEDOT-PSS network. The detection limit of NH3 for the CuTSPc@3D-(N)GF/PEDOT-PSS nanocomposite gas sensor was thus estimated to be 10 ppm at the room temperature.
 |
| Fig. 8 Sensitivity of the pure PEDOT-PSS, pristine CuTSPc@3D-(N)GF and CuTSPc@3D-(N)GF/PEDOT-PSS nanocomposite gas sensors towards 1–1000 ppm of NH3 at room temperature. | |
Fig. 9 demonstrates the selectivity of the CuTSPc@3D-(N)GF/PEDOT-PSS nanocomposite gas sensor for various VOC vapors at a concentration of 200 ppm. The sensitivities of the CuTSPc@3D-(N)GF/PEDOT-PSS nanocomposite gas sensor to NH3, methanol, ethanol, acetone, toluene, chlorobenzene and water were 18.7%, 9.4%, 5.5%, 4.5%, 2.3%, 2.9% and 4.4%, respectively. The CuTSPc@3D-(N)GF/PEDOT-PSS nanocomposite gas sensor exhibited a remarkably high response to NH3 and was slightly sensitive to other VOC vapors. The performance of the as-prepared nanocomposite gas sensor in this work was better than the previous reports in the literature concerning NH3 detection (Table 1).
 |
| Fig. 9 The selectivity of the CuTSPc@3D-(N)GF/PEDOT-PSS nanocomposite gas sensor to various VOC vapors of 200 ppm concentration. | |
Table 1 The sensitivities (S), response times (R1), recovery times (R2), studied detection ranges (Dr), materials (M) and measured temperatures (Tm) of various NH3 gas sensorsa
Authors |
S (%) |
R1 (s) |
R2 (s) |
Dr (ppm) |
M |
Tm (°C) |
PANI: polyaniline; SWCNTs: single wall carbon nanotubes; MWCNT: multi wall carbon nanotube. |
Our work |
8 (50 ppm), 91 (1000 ppm) |
138 |
63 |
1–1000 |
CuTSPc@3D-(N)GF/PEDOT-PSS |
25 |
Xu et al.40 |
2 (10 ppm), 12 (70 ppm) |
10 |
10 |
10–70 |
PEDOT nanowire |
25 |
Kwon et al.41 |
2.1 (5 ppm), 24 (100 ppm) |
<1 |
30 |
5–100 |
PEDOT nanotube |
25 |
Seekaew et al.27 |
0.9 (5 ppm), 7 (1000 ppm) |
180 |
— |
5–1000 |
Graphene/PEDPT-PSS |
25 |
Jian et al.31 |
0.1 (2 ppm), 33 (300 ppm) |
12 |
18 |
2–300 |
SWCNTs/PEDOT-PSS |
25 |
Yoo et al.42 |
0.015 (20 ppm), 0.075 (100 ppm) |
100 |
700 |
0–100 |
pf-MWCNT/PANI |
25 |
Tai et al.43 |
1.67 (23 ppm), 5.55 (117 ppm) |
18 |
58 |
23–141 |
PANI/TiO2 |
25 |
Matsuguchi et al.44 |
1.16 (500 ppm) |
1500 |
— |
— |
PANI |
30 |
Kebiche et al.45 |
7.1 (92 ppm) |
834 |
600 |
92–4628 |
PANI |
25 |
Hong et al.46 |
0.14 (20 ppm), 0.2 (100 ppm) |
14 |
148 |
20–2000 |
Palladium/polypyrrole |
25 |
Crowley et al.47 |
0.24 (100 ppm) |
90 |
90 |
1–100 |
Inkjet-printed PANI |
80 |
Verma et al.48 |
0.9 (200 ppm) |
1 |
420 |
50–200 |
PANI |
25 |
Sengupta et al.49 |
2.3 (100 ppm) |
120 |
300 |
100 |
PANI |
25 |
The improved NH3 sensing properties of the CuTSPc@3D-(N)GF/PEDOT-PSS nanocomposite gas sensor are mainly ascribed to the (i) increased Ra in the CuTSPc@3D-(N)GF/PEDOT-PSS nanocomposite sensing film compared to the PEDOT-PSS one, (ii) inherent sensing properties of the CuTSPc@3D-(N)GF and (iii) π–π interactions resulting from CuTSPc@3D-(N)GF loading, including the 3D-(N)GF as a support and the Cu(II) complex as an active site in the sensing film. (i) Since the Ra of the sensing film is directly proportional to the gas sensitivity,27,50 the much larger Ra of the CuTSPc@3D-(N)GF/PEDOT-PSS nanocomposite sensing film improves the active surface area for gas adsorption. (ii) It is well-known that graphene under ambient conditions behaves as a p-type semiconductor, like conjugated polymers because their electronic properties can be reversibly controlled by doping/dedoping at room temperature.27,51–53 In addition, the electron-withdrawing sulfonic acid groups of the CuTSPc can be viewed as responsible for the charge delocalization of holes in the valence band. Therefore, when the CuTSPc@3D-(N)GF gas sensor is exposed to an electron donating gas like NH3, proton transfer from the sulfonic acid groups of CuTSPc to the nitrogen atoms of NH3 directly gives the self-doped zwitterionic form. Therefore, the depletion of holes from the valence band of the CuTSPc@3D-(N)GF occurs, leading to a significant increase in the resistance. (iii) NH3 molecules may interact with not only the CuTSPc@3D-(N)GF and PEDOT-PSS but also with the π–π bonds formed between the CuTSPc@3D-(N)GF and PEDOT-PSS.9 On exposure to polar molecules like NH3, the interactions can not only induce charge-transfer across delocalized π-electrons but can also lead to the formation of a neutral polymer backbone and a decrease in charge carries, resulting in improved sensing performances.
The NH3 sensing mechanism of the CuTSPc@3D-(N)GF/PEDOT-PSS nanocomposite gas sensor may be explained based on three possible mechanisms: (1) according to the reversible reaction:27,45,54 CuTSPc@3D-(N)GF/PEDOT-PSS–H+ + NH3 ⇄ CuTSPc@3D-(N)GF/PEDOT-PSS + NH4+, proton transfer from the sulfonic acid groups of CuTSPc to the nitrogen atoms of the NH3 molecules directly gives ammonium ions (NH4+). This process is reversible, and in fact, when the NH3 atmosphere is removed, the NH3 molecules decrease the doping level of the CuTSPc@3D-(N)GF/PEDOT-PSS nanocomposite sensing film by partially compensating for the influence of the initial dopants,55 which may change the resistance. (2) The NH3 molecules are absorbed on the surface of the CuTSPc@3D-(N)GF/PEDOT-PSS nanocomposite sensing film by physisorption and the holes of the conductive CuTSPc@3D-(N)GF/PEDOT-PSS nanocomposite sensing film will interact with the electron-donating NH3 analyte.27,54 The charge transfer from adsorbed NH3 molecules not only increases the delocalization degree of the conjugated π-electrons of the nanocomposite sensing film, but also decreases the electrical conductivity of the nanocomposite sensing film.31,41 This mechanism is widely adopted for explaining the change in conductivity of conductive polymers to acidic/basic analytes (doping/dedoping processes). (3) The swelling of the CuTSPc@3D-(N)GF/PEDOT-PSS nanocomposite sensing film can increase the PEDOT distance and decrease the CuTSPc@3D-(N)GF’s conductive pathways, leading to a significant increase in the resistance of the CuTSPc@3D-(N)GF/PEDOT-PSS nanocomposite gas sensor upon NH3 exposure and therefore an enhanced NH3 response.
Conclusion
In this work, a novel NH3 gas sensor based on the CuTSPc@3D-(N)GF/PEDOT-PSS nanocomposite has been successfully fabricated and studied for the first time. The resultant CuTSPc@3D-(N)GF/PEDOT-PSS nanocomposite exhibited excellent sensitivity, dynamic behavior and selectivity to NH3 due possibly to the increase in the specific surface area, intrinsic sensing properties of the CuTSPc@3D-(N)GF and π–π interactions in CuTSPc@3D-(N)GF/PEDOT-PSS. The nanocomposite gas sensor exhibited a much better response to NH3 gas than those of sensors based on pure PEDOT-PSS and a pristine CuTSPc@3D-(N)GF (∼5 and 53 times respectively, with the concentration of NH3 gas at 200 ppm). The response and recovery times of the nanocomposite gas sensor (2.5 min and 1 min, respectively) were much lower than that of pure PEDOT-PSS (6 min and 2.5 min, respectively) towards 200 ppm NH3. This suggests that the CuTSPc@3D-(N)GF/PEDOT-PSS nanocomposite gas sensor, which offers several distinct advantages for NH3 detection over other fabricated sensors including high sensitivity, high productivity, low temperature processing and low cost, is expected to hold great promise for real-world applications.
References
- Y. Dan, et al., Intrinsic response of graphene vapor sensors, Nano Lett., 2009, 9(4), 1472–1475 CrossRef CAS PubMed.
- G. Lu, L. E. Ocola and J. Chen, Gas detection using low-temperature reduced graphene oxide sheets, Appl. Phys. Lett., 2009, 94(8), 083111 CrossRef PubMed.
- F. Schedin, et al., Detection of individual gas molecules adsorbed on graphene, Nat. Mater., 2007, 6(9), 652–655 CrossRef CAS PubMed.
- B. H. Chu, et al., Effect of coated platinum thickness on hydrogen detection sensitivity of graphene-based sensors, Electrochem. Solid-State Lett., 2011, 14(7), K43–K45 CrossRef CAS PubMed.
- H. Vedala, et al., Chemical sensitivity of graphene edges decorated with metal nanoparticles, Nano Lett., 2011, 11(6), 2342–2347 CrossRef CAS PubMed.
- D. Zhang, et al., Room-temperature high-performance acetone gas sensor based on hydrothermal synthesized SnO2-reduced graphene oxide hybrid composite, RSC Adv., 2015, 5(4), 3016–3022 RSC.
- H. Zhang, et al., SnO2 nanoparticles-reduced graphene oxide nanocomposites for NO2 sensing at low operating temperature, Sens. Actuators, B, 2014, 190, 472–478 CrossRef CAS PubMed.
- S. Liu, et al., High performance room temperature NO2 sensors based on reduced graphene oxide-multiwalled carbon nanotubes–tin oxide nanoparticles hybrids, Sens. Actuators, B, 2015, 211, 318–324 CrossRef CAS PubMed.
- X. Huang, et al., Reduced graphene oxide–polyaniline hybrid: preparation, characterization and its applications for ammonia gas sensing, J. Mater. Chem., 2012, 22(42), 22488–22495 RSC.
- C. Lee, et al., Measurement of the elastic properties and intrinsic strength of monolayer graphene, Science, 2008, 321(5887), 385–388 CrossRef CAS PubMed.
- M. D. Stoller, et al., Graphene-based ultracapacitors, Nano Lett., 2008, 8(10), 3498–3502 CrossRef CAS PubMed.
- A. A. Balandin, et al., Superior thermal conductivity of single-layer graphene, Nano Lett., 2008, 8(3), 902–907 CrossRef CAS PubMed.
- M. A. Worsley, et al., Synthesis of graphene aerogel with high electrical conductivity, J. Am. Chem. Soc., 2010, 132(40), 14067–14069 CrossRef CAS PubMed.
- Z. Tang, et al., Noble-metal-promoted three-dimensional macroassembly of single-layered graphene oxide, Angew. Chem., 2010, 122(27), 4707–4711 CrossRef PubMed.
- S. H. Lee, et al., Three-Dimensional Self-Assembly of Graphene Oxide Platelets into Mechanically Flexible Macroporous Carbon Films, Angew. Chem., Int. Ed., 2010, 49(52), 10084–10088 CrossRef CAS PubMed.
- Z. Chen, et al., Three-dimensional flexible and conductive interconnected graphene networks grown by chemical vapour deposition, Nat. Mater., 2011, 10(6), 424–428 CrossRef CAS PubMed.
- Z.-S. Wu, et al., Three-dimensional graphene-based macro- and mesoporous frameworks for high-performance electrochemical capacitive energy storage, J. Am. Chem. Soc., 2012, 134(48), 19532–19535 CrossRef CAS PubMed.
- Z. S. Wu, et al., Three-Dimensional Nitrogen and Boron Co-doped Graphene for High-Performance All-Solid-State Supercapacitors, Adv. Mater., 2012, 24(37), 5130–5135 CrossRef CAS PubMed.
- B. G. Choi, et al., 3D macroporous graphene frameworks for supercapacitors with high energy and power densities, ACS Nano, 2012, 6(5), 4020–4028 CrossRef CAS PubMed.
- X. Cao, et al., Preparation of novel 3D graphene networks for supercapacitor applications, Small, 2011, 7(22), 3163–3168 CrossRef CAS PubMed.
- J. Xiao, et al., Hierarchically porous graphene as a lithium–air battery electrode, Nano Lett., 2011, 11(11), 5071–5078 CrossRef CAS PubMed.
- W. Chen, et al., Self-Assembly and Embedding of Nanoparticles by In Situ Reduced Graphene for Preparation of a 3D Graphene/Nanoparticle Aerogel, Adv. Mater., 2011, 23(47), 5679–5683 CrossRef CAS PubMed.
- Z.-S. Wu, et al., 3D nitrogen-doped graphene aerogel-supported Fe3O4 nanoparticles as efficient electrocatalysts for the oxygen reduction reaction, J. Am. Chem. Soc., 2012, 134(22), 9082–9085 CrossRef CAS PubMed.
- Y.-C. Yong, et al., Macroporous and monolithic anode based on polyaniline hybridized three-dimensional graphene for high-performance microbial fuel cells, ACS Nano, 2012, 6(3), 2394–2400 CrossRef CAS PubMed.
- M. SadegháLaeini, Aqueous aerobic oxidation of alkyl arenes and alcohols catalyzed by copper(II) phthalocyanine supported on three-dimensional nitrogen-doped graphene at room temperature, Chem. Commun., 2014, 50(58), 7855–7857 RSC.
- X. Dong, et al., 3D graphene foam as a monolithic and macroporous carbon electrode for electrochemical sensing, ACS Appl. Mater. Interfaces, 2012, 4(6), 3129–3133 CAS.
- Y. Seekaew, et al., Low-cost and flexible printed graphene–PEDOT:PSS gas sensor for ammonia detection, Org. Electron., 2014, 15(11), 2971–2981 CrossRef CAS PubMed.
- S. Timpanaro, et al., Morphology and conductivity of PEDOT/PSS films studied by scanning–tunneling microscopy, Chem. Phys. Lett., 2004, 394(4), 339–343 CrossRef CAS PubMed.
- A. M. Nardes, et al., Conductivity, work function, and environmental stability of PEDOT: PSS thin films treated with sorbitol, Org. Electron., 2008, 9(5), 727–734 CrossRef CAS PubMed.
- X. Wu, et al., Flexible organic light emitting diodes based on double-layered graphene/PEDOT:PSS conductive film formed by spray-coating, Vacuum, 2014, 101, 53–56 CrossRef CAS PubMed.
- J. Jian, et al., Gas-sensing characteristics of dielectrophoretically assembled composite film of oxygen plasma-treated SWCNTs and PEDOT/PSS polymer, Sens. Actuators, B, 2013, 178, 279–288 CrossRef CAS PubMed.
- H. S. Dehsari, et al., Efficient preparation of ultralarge graphene oxide using a PEDOT: PSS/GO composite layer as hole transport layer in polymer-based optoelectronic devices, RSC Adv., 2014, 4(98), 55067–55076 RSC.
- M. Mahyari, A. Shaabani and Y. Bide, Gold nanoparticles supported on supramolecular ionic liquid grafted graphene: a bifunctional catalyst for the selective aerobic oxidation of alcohols, RSC Adv., 2013, 3(44), 22509–22517 RSC.
- M. Mahyari and A. Shaabani, Nickel nanoparticles immobilized on three-dimensional nitrogen-doped graphene as a superb catalyst for the generation of hydrogen from the hydrolysis of ammonia borane, J. Mater. Chem. A, 2014, 2(39), 16652–16659 CAS.
- A. Hasani, et al., Sensor for volatile organic compounds using an interdigitated gold electrode modified with a nanocomposite made from poly(3,4-ethylenedioxythiophene)-poly(styrenesulfonate) and ultra-large graphene oxide, Microchim. Acta, 2015, 182(7–8), 1551–1559 CrossRef CAS.
- J. N. Gavgani, et al., A room temperature volatile organic compound sensor with enhanced performance, fast response and recovery based on N-doped graphene quantum dots and poly(3,4-ethylenedioxythiophene)-poly(styrenesulfonate) nanocomposite, RSC Adv., 2015, 5(71), 57559–57567 RSC.
- H. Hosseini, et al., Ordered carbohydrate-derived porous carbons immobilized gold nanoparticles as a new electrode material for electrocatalytical oxidation and determination of nicotinamide adenine dinucleotide, Biosens. Bioelectron., 2014, 59, 412–417 CrossRef CAS PubMed.
- M. J. McAllister, et al., Single sheet functionalized graphene by oxidation and thermal expansion of graphite, Chem. Mater., 2007, 19(18), 4396–4404 CrossRef CAS.
- C.-Y. Lin, et al., Using a PEDOT:PSS modified electrode for detecting nitric oxide gas, Sens. Actuators, B, 2009, 140(2), 402–406 CrossRef CAS PubMed.
- J. Xu, et al., Self-assembly of conducting polymer nanowires at air–water interface and its application for gas sensors, Mater. Sci. Eng., B, 2009, 157(1), 87–92 CrossRef CAS PubMed.
- O. S. Kwon, et al., Novel flexible chemical gas sensor based on poly(3,4-ethylenedioxythiophene) nanotube membrane, Talanta, 2010, 82(4), 1338–1343 CrossRef CAS PubMed.
- K.-P. Yoo, et al., Effects of O2 plasma treatment on NH3 sensing characteristics of multiwall carbon nanotube/polyaniline composite films, Sens. Actuators, B, 2009, 143(1), 333–340 CrossRef PubMed.
- H. Tai, et al., Fabrication and gas sensitivity of polyaniline–titanium dioxide nanocomposite thin film, Sens. Actuators, B, 2007, 125(2), 644–650 CrossRef CAS PubMed.
- M. Matsuguchi, et al., Effect of NH3 gas on the electrical conductivity of polyaniline blend films, Synth. Met., 2002, 128(1), 15–19 CrossRef CAS.
- H. Kebiche, et al., Relationship between ammonia sensing properties of polyaniline nanostructures and their deposition and synthesis methods, Anal. Chim. Acta, 2012, 737, 64–71 CrossRef CAS PubMed.
- L. Hong, Y. Li and M. Yang, Fabrication and ammonia gas sensing of palladium/polypyrrole nanocomposite, Sens. Actuators, B, 2010, 145(1), 25–31 CrossRef CAS PubMed.
- K. Crowley, et al., Fabrication of an ammonia gas sensor using inkjet-printed polyaniline nanoparticles, Talanta, 2008, 77(2), 710–717 CrossRef CAS PubMed.
- D. Verma and V. Dutta, Role of novel microstructure of polyaniline-CSA thin film in ammonia sensing at room temperature, Sens. Actuators, B, 2008, 134(2), 373–376 CrossRef CAS PubMed.
- P. P. Sengupta, P. Kar and B. Adhikari, Influence of dopant in the synthesis, characteristics and ammonia sensing behavior of processable polyaniline, Thin Solid Films, 2009, 517(13), 3770–3775 CrossRef CAS PubMed.
- M. Suchea, et al., Low temperature indium oxide gas sensors, Sens. Actuators, B, 2006, 118(1), 135–141 CrossRef CAS PubMed.
- E. Bekyarova, et al., Chemically functionalized single-walled carbon nanotubes as ammonia sensors, J. Phys. Chem. B, 2004, 108(51), 19717–19720 CrossRef CAS.
- J. Huang, et al., Polyaniline nanofibers: facile synthesis and chemical sensors, J. Am. Chem. Soc., 2003, 125(2), 314–315 CrossRef CAS PubMed.
- S. Virji, et al., Polyaniline nanofiber gas sensors: examination of response mechanisms, Nano Lett., 2004, 4(3), 491–496 CrossRef CAS.
- Z. Wu, et al., Enhanced sensitivity of ammonia sensor using graphene/polyaniline nanocomposite, Sens. Actuators, B, 2013, 178, 485–493 CrossRef CAS PubMed.
- M. Matsuguchi, A. Okamoto and Y. Sakai, Effect of humidity on NH3 gas sensitivity of polyaniline blend films, Sens. Actuators, B, 2003, 94(1), 46–52 CrossRef CAS.
Footnotes |
† Electronic supplementary information (ESI) available. See DOI: 10.1039/c5ra13976g |
‡ The first two authors contributed equally to this work. |
|
This journal is © The Royal Society of Chemistry 2015 |
Click here to see how this site uses Cookies. View our privacy policy here.