Next Article in Journal
Influence of Optical Brightening Agent Concentration on Properties of Cotton Fabric Coated with Photochromic Microcapsules Using a Pad-Dry-Cure Process
Next Article in Special Issue
Electrochemical MIP Sensor for Butyrylcholinesterase
Previous Article in Journal
Cellulose in Ionic Liquids and Alkaline Solutions: Advances in the Mechanisms of Biopolymer Dissolution and Regeneration
Previous Article in Special Issue
A Highly Selective Turn-on and Reversible Fluorescent Chemosensor for Al3+ Detection Based on Novel Salicylidene Schiff Base-Terminated PEG in Pure Aqueous Solution
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Novel Application of Electroactive Polyimide Doped with Gold Nanoparticles: As a Chemiresistor Sensor for Hydrogen Sulfide Gas

by
Lee Marvin G. Padua
1,
Jui-Ming Yeh
2,* and
Karen S. Santiago
3,*
1
Department of Math and Physics, College of Science, University of Santo Tomas, Manila 1008, Philippines
2
Department of Chemistry, Research and Development Center for membrane Technology, Center for Nanotechnology, Chung Yuan Christian University, Zhongli, Taoyuan 32023, Taiwan
3
Department of Chemistry, College of Science; Research Center for Natural and Applied Sciences, University of Santo Tomas, Manila 1008, Philippines
*
Authors to whom correspondence should be addressed.
Polymers 2019, 11(12), 1918; https://doi.org/10.3390/polym11121918
Submission received: 4 November 2019 / Revised: 20 November 2019 / Accepted: 20 November 2019 / Published: 21 November 2019
(This article belongs to the Special Issue Polymers for Chemosensing II)

Abstract

:
This research paper presents a new application of electroactive polyimide doped with gold nanoparticles (PI/AuNPs) as a chemiresistor sensor for detecting hydrogen sulfide gas. The synthesis of PI/AuNPs was done in a simple 3-step process of polymerization using the as prepared amine-capped aniline trimer (ACAT), followed by imidization, and doping. Spectral analyses via FTIR, LC-MS and 1H-NMR confirmed the formation of amine-capped aniline trimer with a MW of 288 g mol−1. Comparison of ACAT, BSAA, and PI FTIR spectra showed successful polymerization of the last, while XRD validated the incorporation of metal nanoparticles onto the polymer matrix, showing characteristic diffraction peaks corresponding to gold. Furthermore, TEM, and FE-SEM revealed the presence of well-dispersed Au nanoparticles with an average diameter of about 60 nm. The electroactive PI/AuNPs-based sensor showed a sensitivity of 0.29% ppm−1 H2S at a linear concentration range of 50 to 300 ppm H2S (r = 0.9777). The theoretical limit of detection was found at 0.142 ppm or 142 ppb H2S gas. The sensor provided a stable response reading at an average response time of 43 ± 5 s, which was easily recovered after an average time of 99 ± 5 s. The sensor response was highly repeatable and reversible, with RSD values of 8.88%, and 8.60%, respectively. Compared with the performance of the conventional conducting polyaniline also doped with gold nanoparticles (PANI/AuNPs), the fabricated electroactive PI/AuNPs exhibited improved sensing performance making it a potential candidate in monitoring H2S in the environment and for work-related safety.

Graphical Abstract

1. Introduction

Characterized by a pungent odor similar to the smell of rotten eggs, this colorless and flammable gas known as hydrogen sulfide (H2S) is considered poisonous at certain concentrations. As reported in Occupational and Safety and Health Administration [1,2], inhaling 20 ppm H2S may cause possible fatigue, loss of appetite, headache, irritability, poor memory, and dizziness. Depending on the length of exposure to H2S gas, severe conditions may occur when exposed to higher concentrations. For example, prolonged exposure to 150 ppm H2S gas could quickly paralyze the olfactory nerve disabling it from its capacity to recognize its presence. An hour of exposure to 700 ppm H2S gas could seriously damage the eyes, cause rapid loss of consciousness, or worst lead to possible sudden death [3,4,5,6].
Commonly, hydrogen sulfide gas is released in high quantities during excavation of landfills or swamps [7,8], and in higher amounts during a volcanic activity [9,10,11]. Furthermore, H2S gas is present in petrochemical reservoirs—in oil and natural gas wells, pipelines where unrefined petroleum is transported, and in refinery stations where H2S is removed [12,13,14]. In the paper industry [15,16], and iron smelters [17], H2S is a usual by-product. As H2S gas is denser than air, it can easily accumulate in areas like mine tunnels [18], sewers [19], and manure tanks [20,21]. Due to the ubiquity of H2S gas, and the possible danger this gas may cause, its detection is deemed highly important. However, the gas should not be directly perceived by the olfactory nose, nor should the level be closely assessed without the use of external devices.
There have been reports on the use of gas chromatography (GC) [22,23,24], infrared spectroscopy (IR) [25], ultraviolet spectroscopy (UV) [26,27,28,29], and fluorometry [30] in H2S determination, but these techniques involved the use of relatively expensive bulky instrumentation, and required tedious sampling preparations. Alternatively, the use of sensors was also a subject of heightened interest due to the many modes of detection that could be explored, such as electrochemical [30,31,32], optical [33,34], piezoelectric [35,36], and chemiresistive [37,38]. Sensors based on the chemiresistive mode was the simplest, operating based on the change in electrical resistance as the sensing material absorbs the gas analyte [39,40,41,42,43]. Metal oxide sensors (MOS) have been widely used in the construction of chemiresistive sensors of noxious gases due to their inexpensiveness and simplicity in both preparation and operation. These sensors also offer a wide array of detectable gases, hence expanding possible fields of applications. However, gas sensors fabricated using inorganic materials like metal oxides require operation at higher temperatures, as the high sensitivity of the said sensors is based on the high working temperature. This is often possible through the use of heated filament, and this entails the requirement of a higher amount of electricity to operate [44,45]. On the other hand, conducting polymers (CPs) also appear attractive as active elements in gas sensors. Compared with MOs, CPs present many advantages, such as high sensitivity, short response time, and room temperature operation. However, CPs offer low processability, and poor mechanical strength and chemical stability [46,47]. There have also been reports on the use of metal nanoparticles such as silver nanoparticles [26,48], and gold nanoclusters [30,49], but the detection is coupled to costly optical measurement devices.
This study, therefore, explored the use of combined electroactive polymers and gold nanoparticles to address issues encountered when these materials are used individually. Based on the literature survey conducted, this is the first time that electroactive polyimide decorated with gold nanoparticles has been used and developed for detecting trace levels of hydrogen sulfide gas at ambient room temperature. In the past, electroactive polyimide doped with gold nanoparticles was investigated for other applications such as electrochemical sensing of ascorbic acid [50]. Other studies of these combined materials focused on the effects of gold nanoparticles on polyimide in improving fluxes of the polymer for its possible nanofiltration application [51]. Most studies involving various uses dealt with EPI in different forms, while maximizing its inherent properties. Based on its redox properties, EPI was used as a recognition element in the fabrication of electrochemical sensors [52,53,54,55] and as an anticorrosion material [56,57,58,59,60,61,62,63,64]. Based on its doping and dedoping capability, EPI was also explored as a gas separation membrane [65], and as a smart material characterized for its switchable superhydrophobicity [66]. There have also been papers that investigated means to improve EPI’s mechanical strength [67,68] and thermal properties [69]. It is worthy of mentioning that the results obtained in this research showed that compared with the performance of the traditional conducting PANI/AuNPs, the fabricated electroactive PI/AuNPs exhibited improved sensing performance, making it a potential candidate in monitoring H2S in the environment and for work-related safety.

2. Experimental

2.1. Chemicals and Equipment

Aniline (99.0%, Fluka) was distilled under reduced pressure prior to use. Chemicals such as 4, 4’-diaminodiphenylamine sulfate hydrate (>97.0%, TCI), 4,4′-(4,4′-isopropylidene diphenoxy)-bis(phthalic anhydride) (BSAA, 97.0%, Aldrich), gold(III)chloride trihydrate (HAuCl4·3H2O, 99.99%, Alfa Aesar) 1-methyl-2-pyrrolidone (NMP, 99.7%, MACRON), N,N-dimethylacetamide (DMAc, 99.0%, MACRON, and 99%, Sigma-Aldrich, St. Louis, MO, USA), acetone (Ac, 99.5%, J.T. Baker, New Jersey, USA), methanol (MeOH, 99.8%, J.T. Baker), tetrahydrofuran (THF, 99.8%, RCI Labscan, Bangkok, Thailand), cyclohexanone (CH, 100%, J.T. Baker), dimethylformamide (DMF, 99.9%, J.T. Baker), acetonitrile (ACN, 100%, J.T. Baker) were used without further purification.
Infrared spectra were recorded using JASCO FT/IR-4100. To confirm presence of the Au nanoparticles, the samples were cut using a diamond knife to a 60–90 nm thick sections, and observed using a JEOL-200FX TEM (JEOL, Tokyo, Japan). The morphology of samples was recorded using scanning electron microscopy (SEM) (Hitachi S-2300). X-ray diffraction (XRD) analyses were recorded using a Philips X′pert Pro X-ray diffractometer. Electrochemical properties were investigated using a conventional three-electrode system connected to a VoltaLab 40 analytical voltammeter. 1H NMR investigations were carried out using Bruker 300 spectrometer, referenced to internal standard of tetramethylsilane (TMS). DMSO was used as solvent. LC-MS further facilitated analysis of samples using Bruker Daltonics ion trap mass spectrometer (Model: Esquire 2000 with an Agilent ESI source (Model: G1607-6001, Leipzig, Germany).
A UNI-T UT71A multimeter with UT71A_B V3.00 software/Agilent 34410A desktop multimeter with Keysight BenchVue V 3.0 software (MatLab, CA, USA), was used in testing the gas sensing capability of the polymeric membranes.

2.2. Synthesis and Characterization of PI and PI/AuNPs

Amino-capped aniline trimer (ACAT) was prepared by following the procedure reported by Wei et al. Briefly, the synthesis of ACAT involved the use of ammonium persulfate to oxidize 2 equivalents of aniline and 1,4-phenylenediamine [70]. The success of ACAT preparation was confirmed by 1H NMR mass, LC-MS, and FT-IR spectroscopy.
The electroactive polyimide (PI) was synthesized using ACAT and BSAA. Solution A was prepared by dissolving BSAA (0.52 g, 1 mmol) in 8.0 g of DMAc with continuous stirring for 30 min at room temperature. Solution B was prepared by dissolving ACAT (0.29 g, 2 mmol) in 8.0 g of DMAc with continuous stirring for 30 min at room temperature. After stirring for 30 min, solution B was added to solution A, and mechanically stirred for 24 h to prepare a poly(amic acid) solution. The poly(amic acid)(PAA) was converted to polyimide(PI) by chemical imidization reaction. This reaction was done by adding acetic anhydride (0.102 g, 1 mmol) and pyridine (0.079 g, 1 mmol) to the solution, and refluxing for 3 h under nitrogen. Methanol (250 mL) was used to wash the PI and dried at 60 °C under vacuum for 1 day. The PI was characterized by 1H NMR, LC-MS, FT-IR spectroscopy, X-ray diffraction.
The gold nanoparticle-doped electroactive polyimide (PI/AuNPs) samples were prepared by dissolving PI in DMAc, and adding aqueous HAuCl4 solution (1 mM) at room temperature, then stirring for 6 h. After the reaction, centrifugal filtration was done to collect the precipitate. Large volumes of distilled water were used to wash the residue. The PI/AuNPs were dried in a vacuum oven at 60 °C. The PI-AuNP was characterized using TEM, FT-SEM, and X-ray diffraction.

2.3. Evaluation of Fabricated Sensors Toward H2S

H2S headspace generation. Hydrogen sulfide is generated in a reaction chamber by mixing 7.1 mg ferrous sulfide (FeS) and 10 mL 5 M hydrocholoric acid (HCl). This reaction produced 50 ppm concentration of H2S in a 55 mL sample chamber. The procedure was repeated to achieve 100, 150, 200, 250, and 300 ppm H2S.
As illustrated in Figure 1a, the gas sensing set-up was composed of a nitrogen tank connected to an enclosed reaction chamber (left, V = 55 mL), where generation of hydrogen sulfide gas occurred. The generated H2S gas was flushed into the sample chamber (middle, V = 55 mL) for determination using the fabricated sensor attached to a UNI-T multimeter. The resistance data obtained was recorded continuously using a computer. The gas exhausted from the sample chamber was bubbled in the third chamber containing water.
Substrate preparation. The substrate (Figure 1b) is composed of gold wires (18 karats, Ø = 0.5 mm, l = 0.5 cm) soldered to polished copper wires (Ø = 0.5 mm, l = 2.5 cm). A pair of soldered Au-Cu wires, separated by a 200 µm gap, was mounted into a custom made Teflon bar (10 mm × 10 mm × 4 mm). The wires were cleaned via ultrasonication in detergent solution for 15 min. This step was immediately followed with washing of wires using water, acetone, methanol, water, and then dried. The wires were held in a sturdy and upright position by applying a cyanoacrylate adhesive (Mighty Bond™ by Pioneer ® adhesives). The exposed Au ends, which served as the substrate, were subsequently polished with high grade of abrasive paper (1200 grit) and pad with aluminum slurry (Ø = 3 µm). The polished electrodes were then washed with methanol, sonicated in ultrapure water for 5 minutes, and dried.
Sensor performance evaluation. The sensor performance was evaluated based on sensitivity, linearity, linear concentration range, limit of detection, dynamic response and recovery characteristics, repeatability and reproducibility, and selectivity.

3. Results and Discussion

3.1. Synthesis and Characterization of PI-Based Materials

3.1.1. Synthesis of electroactive polyimide-based materials

The ACAT was synthesized by the oxidative coupling of 4-4′-diaminodiphenylamine and aniline monomer by the addition of ammonium persulfate as the oxidant (Scheme 1). The synthesized ACAT was then subjected to chemical polymerization with the aid of BSAA, resulting in the formation of poly(amic acid). Finally, the poly(amic acid) underwent chemical imidization by acetic anhydride and pyridine to form the final product, electroactive polyimide.
Scheme 2, adapted from the work of Ji et al. [50], shows the proposed mechanism of the formation of electroactive PI/AuNPs by doping gold nanoparticles onto the synthesized PI, with the aid of dimethyl acetamide and ethyl acetate solvents.

3.1.2. Structural and Morphological Characterization

Figure 2a,b shows the LC-MS spectra of the synthesized ACAT. It can be seen that the maximum intensity in the MS+ spectrum occurred at 289.1 m/z, while the maximum in the MS- spectrum occurred at 286.9 m/z. These results suggest that the synthesized powder is ACAT. In theory, the molecular weight of ACAT is 288 g mol−1. The peak shifted by +/−1 which could be attributed to H.
Further confirmation of the successful synthesis of ACAT using 1H-NMR spectroscopy was done. The sample used was prepared by dissolving a small amount of the synthesized ACAT in dimethylsulfoxide-d6((CD3)2SO). Figure 2c shows the spectrum obtained from the analysis.
The results show that PI (Figure 3c) has the same characteristic peaks with ACAT (Figure 3a) at 1504 and 1596 cm−1. The characteristic peaks at 1504 and 1596 cm−1 may be due to the vibration bands of benzenoid rings and quinoid rings of ACAT, respectively. At 1750 cm−1, PI (Figure 3c) and BSAA (Figure 3b) have characteristic peaks referring to the vibration bands of BSAA′s C=O groups. Based on the overall spectra, the synthesis of PI was successfully achieved.
TEM and FE-SEM micrographs reveal the presence of well-dispersed Au nanoparticles with an average diameter of about 60 nm, as shown in Figure 4.
The diffractogram in Figure 5a shows a broad peak between 15–20° (2θ), which can be attributed to the semi-crystalline nature of the PI sample, having periodic parallel polymer chains. The diffractogram of the PI/AuNPs (Figure 5b) show the characteristic peaks of gold occurring at 38.01°, 43.96°, 64.50°, and 77.42° (2θ), corresponding to reflections from the planes, (111), (200), (220), and (311) [71]. The prominent diffraction at 38.01° revealed that zero valent gold grew and mostly fixed in the direction (111). This behavior also signifies the formation of pure gold nanocrystals [72].

3.2. Evaluation of the Sensor′s Performance

Sensitivity, linearity, and limit of detection. Shown in Figure 6a is the calibration curve constructed based in the performance of electroactive polyimide doped with gold nanoparticles (PI/AuNPs) towards increasing concentration of hydrogen sulfide gas. Responses were reported as (R/R0) × 100%; R=/RgR0/, where Rg is the resistance of the chemiresistor gas sensor and R0 refers to baseline resistance. The sensor presented a sensitivity of 0.29% ppm−1 H2S gas and a linearity of 0.9777 at a dynamic linear concentration range of 50 to 300 ppm H2S gas. The theoretical detection limit was found to be 0.142 ppm or 142 ppb H2S, estimated according to the least-square method of fitting in the linear regime.
As indicated in Figure 6b, the decrease in resistance in the electroactive PI/AuNPs film could possibly be attributed to further doping of the polymer due to the sulfurous acid formed. Hydrogen sulfide gas (Ka = 1.1 × 10−7) in humid condition may generate sulfurous acid (Ka = 1.3 × 10−2), which consequently dissociates more effectively into hydrogen and sulfite ions. At room temperature and pressure, the hydrogen ions, then, protonate the polymer’s imine nitrogen sites of the ACAT chains. This phenomenon results to a corresponding increase in conductivity or decrease in resistance. Experimental data show that higher concentrations of H2S gas lead to higher decrease in resistance. It is speculated that the resulting positive charges in the imine nitrogen are counteracted by the negatively charged sulfite ions. Furthermore, the gold nanoparticles could possibly be contributing to the stabilization of the interaction by attracting the sulfur-containing counter ions toward the ACAT segments.
Repeatability of the sensor′s response.Figure 7a shows a comparison of calibration plots of electroactive PI/AuNPs and PANI/AuNPs chemiresistor sensors towards increasing concentration of hydrogen sulfide gas. The latter has a sensitivity of 0.11% ppm−1 H2S, a theoretical detection limit of 0.45 ppm H2S, and a linearity of 0.9551 at the same linear concentration range of 50 to 300 ppm H2S gas. Based on sensitivity values, there was a 62% improvement in the performance with electroactive PI/AuNPs sensor. This may be attributed to the formation of additional conducting paths due to further doping of PI′s active sites. This, consequently, improved hopping of electrons throughout the polymeric chain.
The repeatability studies depicted in Figure 7b,c prove the better precision of all the trials, with respect to sensitivity values, in electroactive PI/AuNPs (% RSD = 8.88%) than PANI/AuNPs (% RSD = 28.36%).
Response and Recovery Characteristics of the Sensor. Response time refers to the duration it takes the sensor to change from its initial state to its stable value, while recovery time is the period it takes the sensor to return to its initial state. As shown in Figure 8, a response time of 43 s and recovery time of 99 s was exhibited by PI/AuNPs while a response time of 55 s and recovery time of 103 s was shown by the PANI/AuNPs. The faster response time and recovery in PI/AuNPs could be attributed to the polymer′s three-dimensional (3D) surface, i.e., the porous sensing membrane contributing to higher surface area (see FE-SEM presented in Figure 4d). This 3D structure, in both PI/AuNPs and PANI/AuNPs, allows easier diffusion of gas analyte into and out of the polymer matrix. Morphological studies revealed that PI/AuNPs are more porous than PANI/AuNPs. This increase in surface area could possibly be the cause of the higher response at a faster gas diffusion rate in PI/AuNPs.
Reversibility of the Sensor′s response. As shown in Figure 9a, the PI/AuNPs sensor exhibited a good reversibility within 3 cycles of exposure to alternating N2 blank and 200 ppm H2S (RSD = 8.60%). The reading returned to its baseline value once the gas analyte was removed.
The case is different compared with the performance of PANI/AuNPs sensor. Figure 9b shows that the sensor did not exhibit a good reversible process. The sensor experienced difficulty in reverting to the baseline reading after the first cycle of exposure to H2S gas. An RSD of 41.79% validated that PANI/AuNPs showed a poor reversibility.
Selectivity. The selectivity studies in Figure 10 revealed that the electroactive PI/AuNPs-based chemiresistor sensor resulted in the highest resistance change towards 300 ppm hydrogen sulfide gas compared to minimal to almost negligible response when the sensor was exposed to pure forms of solvents of varying polarity such as hexane, ethyl acetate, and methanol vapors. The same sensor, when exposed and tested towards the same concentration of ammonia gas, resulted in a signal that was opposite in direction, implying a different sensing mechanism.

4. Conclusions

An electroactive polyimide/gold nanoparticles-based chemiresistor sensor that operates at ordinary room temperature was developed. The change in the electrical property, particularly the resistance of the sensing material, was used as a means to monitor the change in H2S gas concentration. Spectral studies confirmed successful synthesis of its sensing material. Morphological studies, on the other hand, confirmed the increase in surface area of electroactive polyimide after decorating it with gold nanoparticles. Compared with the conventional conducting PANI, likewise doped with AuNPs to ensure similar matrix, the electroactive PI/AuNPs exhibited an improved sensing performance. The proposed sensor offers an alternative means of monitoring hydrogen sulfide gas in a simple and inexpensive way.

Author Contributions

J.-M.Y. and K.S.S. conceptualized the project and overall design of the experiment. L.M.G.P. performed the experimental part under the supervision of K.S.S., and both wrote the manuscript. J.-M.Y. and K.S.S. edited and finalized the paper.

Funding

The authors extend their gratitude for the financial assistance provided by the Department of Science and Technology-Philippine Council for Industry, Energy, and Emerging Technology Research and Development (DOST-PCIEERD, through the Manila Economic and Cultural Office-MECO), and University of Santo Tomas-Research Center for the Natural and Applied Sciences; and the Ministry of Science and Technology, Taiwan, R.OSC (MOST 104-2113-M-033-001-MY3 and MOST 104-2923-M-033-001-MY2) as well as the Department of Chemistry at Chung Yuan Christian University.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. OSHA Fact Sheet; U.S. Department of Labor: Washington, DC, USA. Available online: https://www.osha.gov/SLTC/hydrogensulfide/hazards.html (accessed on 3 November 2019).
  2. Occupational Safety and Health Administration Technical Center. Hydrogen Sulfide Backup Data Report (ID-141); Occupational Safety and Health Administration Technical Center: Salt Lake City, UT, USA, 1989.
  3. Rubright, S.; Pearce, L.; Peterson, J. Environmental toxicology of hydrogen sulfide. Nitric Oxide 2017, 71, 1–13. [Google Scholar] [CrossRef]
  4. Guidotti, T.L. Hydrogen Sulfide Intoxication. In Handbook of Clinical Neurology, 3rd ed.; Lotti, M., Bleecker, M.L., Eds.; Elsevier: Amsterdam, The Netherlands, 2015; Chapter 8; Volume 131, pp. 111–133. [Google Scholar]
  5. Roth, S.H. Handbook of Hazardous Materials; Academic Press: Cambridge, UK, 1993; pp. 367–376. [Google Scholar]
  6. Terrence, S. Hydrogen Sulphide in Agricultural Biogas Systems; Ministry of Agriculture Food and Rural Affairs: Centre Wellington, ON, Canada, 2011; pp. 1–7.
  7. Plaza, C.; Xu, Q.; Townsend, T.; Bitton, G.; Booth, M. Evaluation of alternative landfill cover soils for attenuating hydrogen sulfide from construction and demolition (C&D) debris landfills. J. Environ. Manag. 2007, 84, 314–322. [Google Scholar]
  8. Shen, D.S.; Du, Y.; Fang, Y.; Hu, Li.; Fang, Ch.; Long, Yu. Characteristics of H2S emission from aged refuse after excavation exposure. J. Environ. Manag. 2015, 154, 159–165. [Google Scholar] [CrossRef]
  9. Textor, C.; Graf, H.F.; Herzog, M.; Oberhuber, J.M. Injection of gases into the stratosphere by explosive volcanic eruptions. J. Geophys. Res. 2003, 108, 4606. [Google Scholar] [CrossRef]
  10. Oppenheimer, C.; Scaillet, B.; Martin, R.S. Sulfur Degassing from Volcanoes: Source Conditions, Surveillance, Plume Chemistry and Earth System Impacts. Sulfur Magmas Melts 2011, 73, 363–422. [Google Scholar]
  11. Edmonds, M.; Grattan, J.; Michnowicz, S. Volcanic Gases: Silent Killers. In Observing the Volcano World. Advances in Volcanology; Fearnley, C.J., Bird, D.K., Haynes, K., McGuire, W.J., Jolly, G., Eds.; Springer: Cham, Switzerland, 2015. [Google Scholar]
  12. Jafarinejad, S. Odours emission and control in the petroleum refinery: A review. Curr. Sci. Perspect. 2016, 2, 78–82. [Google Scholar]
  13. Jafarinejad, S. Pollutions and Wastes from the Petroleum Industry. In Petroleum Waste Treatment and Pollution Control; Butterworth-Heinemann: Oxford, UK, 2017; pp. 19–83. [Google Scholar]
  14. Shivanthan, M.C.; Perera, H.; Jayasinghe, S.; Karunanayake, P.; Chang, T.; Ruwanpathirana, S.; Jayasinghe, N.; De Silva, Y.; Jayaweerabandara, D. Hydrogen sulphide inhalational toxicity at a petroleum refinery in Sri Lanka: A case series of seven survivors following an industrial accident and a brief review of medical literature. J. Occup. Med. Toxicol. (Lond. UK) 2013, 8, 9. [Google Scholar] [CrossRef]
  15. Kangas, J.; Jappinen, P.; Savolainen, H. Exposure to Hydrogen Sulfide, Mercaptans and Sulfur Dioxide in Pulp Industry. Am. Ind. Hyg. Assoc. J. 1984, 45, 787–790. [Google Scholar] [CrossRef]
  16. Veluchamy, C.; Kalamdhad, A.S. Enhancement of hydrolysis of lignocellulose waste pulp and paper mill sludge through different heating processes on thermal pretreatment. J. Clean. Prod. 2017, 168, 219–226. [Google Scholar] [CrossRef]
  17. Crundwell, F.; Moats, M.; Ramachandran, V.; Robinson, T.; Davenport, W.G. Extractive Metallurgy of Nickel, Cobalt and Platinum Group Metals; Elsevier Ltd.: Oxford, UK, 2011; pp. 443–455. ISBN 9780080968094. [Google Scholar]
  18. Chaulya, S.K.; Prasad, G.M. Chapter 3—Gas Sensors for Underground Mines and Hazardous Areas. In Sensing and Monitoring Technologies for Mines and Hazardous Areas Monitoring and Prediction Technologies; Elsevier: Amsterdam, The Netherlands, 2016; pp. 161–212. [Google Scholar]
  19. Jiang, G.; Keller, J.; Bond, P.L. Determining the long-term effects of H2S concentration, relative humidity and air temperature on concrete sewer corrosion. Water Res. 2014, 65, 157–169. [Google Scholar] [CrossRef]
  20. Andriamanohiarisoamanana, F.J.; Sakamoto, Y.; Yamashiro, T.; Yasui, S.; Iwasaki, M.; Ihara, I.; Tsuji, O.; Umetsu, K. Effects of handling parameters on hydrogen sulfide emission from stored dairy manure. J. Environ. Manag. 2015, 154, 110–116. [Google Scholar] [CrossRef] [PubMed]
  21. Maurer, D.L.; Koziel, J.A.; Bruning, K.; Parker, D.B. Farm-scale testing of soybean peroxidase and calcium peroxide for surficial swine manure treatment and mitigation of odorous VOCs, ammonia and hydrogen sulfide emissions. Atmos. Environ. 2017, 166, 467–478. [Google Scholar] [CrossRef]
  22. Kim, K.H. Performance characterization of the GC/PFPD for H2S, CH3SH, CH3SCH3, and CH3SSCH3 in air. Atmos. Environ. 2005, 39, 2235–2242. [Google Scholar] [CrossRef]
  23. Vitvitsky, V.; Banerjee, R. H2S analysis in biological samples using gas chromatography with sulfur chemiluminescence detection. Methods Enzym. 2015, 554, 111–123. [Google Scholar]
  24. Varlet, V.; Giuliani, N.; Palmiere, C.; Maujean, G.; Augsburger, M. Hydrogen Sulfide Measurement by Headspace-Gas Chromatography-Mass Spectrometry (HS-GC-MS): Application to Gaseous Samples and Gas Dissolved in Muscle. J. Anal. Toxicol. 2015, 39, 52–57. [Google Scholar] [CrossRef] [PubMed]
  25. Larsen, E.S.; Hong, W.W.; Spartz, M.L. Hydrogen Sulfide Detection by UV-Assisted Infrared Spectrometry. Appl. Spectrosc. 1997, 51, 1656–1667. [Google Scholar] [CrossRef]
  26. Chen, R.; Morris, H.R.; Whitmore, P.M. Fast detection of hydrogen sulfide gas in the ppmv range with silver nanoparticle films at ambient conditions. Sens. Actuators B Chem. 2013, 186, 431–438. [Google Scholar] [CrossRef]
  27. Gersen, S.; Van Essen, M.; Visser, P.; Ahmad, M.; Mokhov, A.; Sepman, A.; Alberts, R.; Douma, A.; Levinsky, H. Detection of H2S, SO2 and NO2 in CO2 at pressures ranging from 1-40 bar by using broadband absorption spectroscopy in the UV/VIS range. Energy Procedia 2014, 63, 2570–2582. [Google Scholar] [CrossRef]
  28. Shariati-Rad, M.; Irandoust, M.; Jalilvand, F. Spectrophotometric determination of hydrogen sulfide in environmental samples using sodium 1,2-naphthoquinone-4-sulfonate and response surface methodology. Int. J. Environ. Sci. Technol. 2016, 13, 1347–1356. [Google Scholar] [CrossRef]
  29. Wallace, K.J.; Cordero, S.R.; Tan, C.P.; Lynch, V.M.; Anslyn, E.V. A colorimetric response to hydrogen sulfide. Sens. Actuators B Chem. 2007, 120, 362–367. [Google Scholar] [CrossRef]
  30. Zhang, Y.; Li, M.; Niu, Q.; Gao, P.; Zhang, G.; Dong, C.; Shuang, S. Gold nanoclusters as fluorescent sensors for selective and sensitive hydrogen sulfide detection. Talanta 2017, 171, 143–151. [Google Scholar] [CrossRef] [PubMed]
  31. Lawrence, N.S.; Deo, R.P.; Wang, J. Electrochemical determination of hydrogen sulfide at carbon nanotube modified electrodes. Anal. Chim. Acta 2004, 517, 131–137. [Google Scholar] [CrossRef]
  32. Zeng, L.; He, M.; Yu, H.; Li, D. An H2S Sensor Based on Electrochemistry for Chicken Coops. Sensors 2016, 16, 1398. [Google Scholar] [CrossRef]
  33. Ke, Z.J.; Tang, D.L.; Lai, X.; Dai, Z.Y.; Zhang, Q. Optical fiber evanescent-wave sensing technology of hydrogen sulfide gas concentration in oil and gas fields. Optik 2018, 157, 1094–1100. [Google Scholar] [CrossRef]
  34. Zhou, H.; Wen, J.Q.; Zhang, X.Z.; Wang, W.; Feng, D.Q.; Wang, Q.; Jia, F. A Study on Fiber-optic Hydrogen Sulfide Gas Sensor. Phys. Procedia 2014, 56, 1102–1106. [Google Scholar] [CrossRef]
  35. He, H.; Dong, C.; Fu, Y.; Han, W.; Zhao, T.; Xing, L.; Xue, X. Self-powered smelling electronic-skin based on the piezo-gas-sensor matrix for real-time monitoring the mining environment. Sens. Actuators B Chem. 2018, 267, 392–402. [Google Scholar] [CrossRef]
  36. Kuchmenko, T.A.; Kochetova, Z.Y.; Silina, Y.E.; Korenman, Y.I.; Kulin, L.A.; Lapitski, I.V. Determination of Trace Amounts of Hydrogen Sulfide in a Gas Flow Using a Piezoelectric Detector. J. Anal. Chem. 2007, 62, 781–787. [Google Scholar] [CrossRef]
  37. Berahman, M.; Sheikhi, M. Hydrogen sulfide gas sensor based on decorated zigzag graphene nanoribbon with copper. Sens. Actuators B Chem. 2015, 219, 338–345. [Google Scholar] [CrossRef]
  38. Tomchenko, A.A.; Harmer, G.P.; Marquis, B.T.; Allen, J.W. Semiconducting metal oxide sensor array for the selective detection of combustion gases. Sens. Actuators B Chem. 2003, 93, 126–134. [Google Scholar] [CrossRef]
  39. MalekAlaie, M.; Jahangiri, M.; Rashidi, A.; HaghighiAsl, A.; Izadi, N.; Rashidi, A. Selective hydrogen sulfide (H2S) sensors based on molybdenum trioxide (MoO3) nanoparticle decorated reduced graphene oxide. Mater. Sci. Semicond. Process. 2015, 38, 93–100. [Google Scholar] [CrossRef]
  40. Kaur, M.; Jain, N.; Sharma, K.; Bhattacharya, S.; Roy, M.; Tyagi, A.; Gupta, S.; Yakhmi, J.V. Room-temperature H2S gas sensing at ppb level by single crystal In2O3 whiskers. Sens. Actuators B Chem. 2008, 133, 456–461. [Google Scholar] [CrossRef]
  41. Alaie, M.M.; Jahangiri, M.; Rashidi, A.; Asl, A.H.; Izadi, N.; Rashidi, A. A novel selective H2S sensor using dodecylamine and ethylenediamine functionalized graphene oxide. J. Ind. Eng. Chem. 2015, 29, 97–103. [Google Scholar] [CrossRef]
  42. Sarfraz, J.; Ihalainen, P.; Määttänen, A.; Peltonen, J.; Linden, M. Printed hydrogen sulfide gas sensor on paper substrate based on polyaniline composite. Thin Solid Film. 2013, 534, 621–628. [Google Scholar] [CrossRef]
  43. Li, M.; Zhou, D.; Zhao, J.; Zheng, Z.; He, J.; Hu, L.; Xia, Z.; Tang, J.; Liu, H. Resistive gas sensors based on colloidal quantum dot (CQD) solids for hydrogen sulfide detection. Sens. Actuators B Chem. 2015, 217, 198–201. [Google Scholar] [CrossRef]
  44. Kumar, V.; Sunny; Rawal, I.; Mishra, V.; Dwivedi, R.; Das, R. Fabrication and characterization of gridded Pt/SiO2/Si MOS structure for hydrogen and hydrogen sulphide sensing. Mater. Chem. Phys. 2014, 146, 418–424. [Google Scholar] [CrossRef]
  45. Emelin, E.V.; Nikolaev, I.N. Sensitivity of mos sensors to hydrogen, hydrogen sulfide, and nitrogen dioxide in different gas atmospheres. Meas. Tech. 2006, 49, 524–528. [Google Scholar] [CrossRef]
  46. Crowley, K.; Morrin, A.; Sheperd, R.; Panhuis, M.; Wallace, G.; Smyth, M.; Killard, A. Fabrication of Polyaniline-Based Gas Sensors using Piezoelectric Inkjet and Silkscreen Printing for the Detection of Hydrogen Sulfide. IEEE Sens. J. 2010, 10, 1419–1426. [Google Scholar] [CrossRef] [Green Version]
  47. Liu, C.; Hayashi, K.; Toko, K. Au nanoparticles decorated polyaniline nanofiber sensor for detecting volatile sulfur compounds in expired breath. Sens. Actuators B Chem. 2012, 161, 504–509. [Google Scholar] [CrossRef]
  48. Chen, R.; Whitmore, P.M. Chapter 6: Silver Nanoparticle Films as Hydrogen Sulfide Gas Sensors with Applications in Art Conservation. In The Science and Function of Nanomaterials: From Synthesis to Application; Harper-Leatherman, A.S., Solbrig, C.M., Eds.; ACS Symposium Series; American Chemical Society: Washington, DC, USA, 2014. [Google Scholar]
  49. Yuan, Z.; Lu, F.; Peng, M.; Wang, C.-W.; Tseng, Y.T.; Du, Y.; Cai, N.; Lien, C.W.; Chang, H.T.; He, Y.; et al. Selective Colorimetric Detection of Hydrogen Sulfide Based on Primary Amine-Active Ester Cross-Linking of Gold Nanoparticles. Anal. Chem. 2015, 87, 7267–7273. [Google Scholar] [CrossRef]
  50. Ji, W.F.; Chu, C.M.; Hsu, S.C.; Lu, Y.D.; Yu, Y.C.; Santiago, K.S.; Yeh, J.M. Synthesis and characterization of organo-soluble aniline oligomer-based electroactive doped with gold nanoparticles, and application to electrochemical sensing of ascorbic acid. Polymers 2017, 128, 218–228. [Google Scholar] [CrossRef]
  51. Vanherck, K.; Vankelecom, I.; Verbiest, T. Improving fluxes of polyimide membranes containing gold nanoparticles by photothermal heating. J. Membr. Sci. 2011, 373, 5–13. [Google Scholar] [CrossRef]
  52. Weng, C.J.; Jhuo, Y.S.; Chen, Y.L.; Feng, C.F.; Chang, C.H.; Chen, S.W.; Yeh, J.M.; Wei, Y. Intrinsically electroactive polyimide microspheres fabricated by electrospraying technology for ascorbic acid detection. J. Mater. Chem. 2011, 21, 15666–15672. [Google Scholar] [CrossRef]
  53. Weng, C.J.; Huang, K.Y.; Jhuo, Y.S.; Chen, Y.L.; Feng, C.F.; Cho-Ming, C.; Yeh, J.M. Controllable Electroactive Polyimide Particles Size Generated by Electrospraying. Polym. Int. 2012, 61, 205–212. [Google Scholar] [CrossRef]
  54. Huang, T.C.; Lin, S.T.; Yeh, L.C.; Chen, C.A.; Huang, H.Y.; Nian, Z.Y.; Chen, H.H.; Yeh, J.M. Aniline pentamer-based electroactive polyimide prepared from oxidation coupling polymerization for electrochemical sensing application. Polymers 2012, 53, 4373–4379. [Google Scholar] [CrossRef]
  55. Huang, T.C.; Yeh, L.C.; Huang, H.Y.; Nian, Z.Y.; Yeh, Y.C.; Chou, Y.C.; Yeh, J.M.; Tsai, M.H. The use of a carbon paste electrode mixed with multiwalled carbon nanotube/electroactive polyimide composites as an electrode for sensing ascorbic acid. Polym. Chem. 2014, 5, 630–637. [Google Scholar] [CrossRef]
  56. Huang, K.Y.; Jhuo, Y.S.; Wu, P.S.; Lin, C.H.; Yu, Y.H.; Yeh, J.M. Electrochemical Studies for the Content of Amine-Capped Aniline Trimers on the Corrosion Protection Effect of as-prepared Electroactive Polyimide Coatings. Eur. Polym. J. 2009, 45, 485–493. [Google Scholar] [CrossRef]
  57. Weng, C.J.; Huang, J.Y.; Huang, K.Y.; Jhuo, Y.S.; Tsai, M.H.; Yeh, J.M. Advanced anticorrosive coatings prepared from electroactive polyimide–TiO2 hybrid nanocomposite materials. Electrochim. Acta 2010, 55, 8430–8438. [Google Scholar] [CrossRef]
  58. Huang, H.Y.; Huang, T.C.; Yeh, T.C.; Tsai, C.Y.; Lai, C.L.; Tsai, M.H.; Yeh, J.M.; Chou, Y.C. Advanced anticorrosive materials prepared from amine-capped aniline trimer-based electroactive polyimide-clay nanocomposite materials with synergistic effects of redox catalytic capability and gas barrier properties. Polymers 2011, 52, 2391–2400. [Google Scholar] [CrossRef]
  59. Huang, T.C.; Yeh, T.C.; Huang, H.Y.; Ji, W.F.; Chou, Y.C.; Hung, W.I.; Yeh, J.M.; Tsai, M.H. Electrochemical Studies on aniline-pentamer-Based Electroactive Polyimide Coating: Corrosion Protection and Electrochromic Properties. Electrochim. Acta 2011, 56, 10151–10158. [Google Scholar] [CrossRef]
  60. Chang, K.C.; Lu, H.I.; Peng, C.W.; Lai, M.C.; Hsu, S.C.; Hsu, M.H.; Tsai, Y.K.; Chang, C.H.; Hung, W.I.; Wei, Y.; et al. Nanocasting Technique to Prepare Lotus-leaf-like Superhydrophobic Electroactive Polyimide as Advanced Anticorrosive Coatings. ACS Appl. Mater. Interfaces 2013, 5, 1460–1467. [Google Scholar] [CrossRef]
  61. Chou, Y.C.; Lee, P.C.; Hsu, T.F.; Huang, W.Y.; Zi-Han, L.; Chuang, C.Y.; Yang, T.I.; Yeh, J.M. Synthesis and anticorrosive properties of electroactive polyimide/SiO2 composites. Polym. Compos. 2014, 35, 617–625. [Google Scholar] [CrossRef]
  62. Chang, K.C. Advanced anticorrosive coatings prepared from electroactive polyimide/graphene nanocomposites with synergistic effects of redox catalytic capability and gas barrier properties. Express Polym. Lett. 2014, 8, 243–255. [Google Scholar] [CrossRef]
  63. Huang, T.C.; Yeh, L.C.; Lai, G.H.; Huang, B.S.; Yang, T.I.; Hsu, S.C.; Lo, A.Y.; Yeh, J.M. Advanced Superhydrophobic Electroactive Fluorinated Polyimide and Its Application in Anticorrosion Coating. Int. J. Green Energy 2016, 14, 113–120. [Google Scholar] [CrossRef]
  64. Ji, W.F.; Li, C.W.; Yu, S.K.; Chen, P.J.; Chen, H.L.; Chen, R.D.; Chen, B.H.; Hsu, C.L.; Yeh, J.M. Biomimetic electroactive polyimide with rose petal-like surface structure prepared from nanocasting technique for anticorrosive coating application. Express Polym. Lett. 2017, 11, 635–644. [Google Scholar] [CrossRef]
  65. Weng, C.J.; Jhuo, Y.S.; Huang, K.Y.; Feng, C.F.; Yeh, J.M.; Wei, Y.; Tsai, M.H. Mechanically and Thermally Enhanced Intrinsically Dopable Polyimide Membrane with Advanced Gas Separation Capabilities. Macromolecules 2011, 44, 6067–6076. [Google Scholar] [CrossRef]
  66. Weng, C.J.; Jhuo, Y.S.; Chang, C.H.; Feng, C.F.; Peng, C.W.; Dai, C.F.; Yeh, J.M.; Wei, Y. A smart surface prepared using the switchable superhydrophobicity of neat electrospun intrinsically electroactive polyimide fiber mats. Soft Matter 2011, 7, 10313–10318. [Google Scholar] [CrossRef]
  67. Chang, K.C.; Huang, K.Y.; Hsu, C.H.; Ji, W.F.; Lai, M.C.; Hung, W.I.; Chuang, T.L.; Yeh, J.M. Synthesis of ultra-high-strength electroactive polyimide membranes containing oligoaniline in the main chain by thermal imidization reaction. Eur. Polym. J. 2014, 56, 26–32. [Google Scholar] [CrossRef]
  68. Chang, K.C.; Lu, H.I.; Lai, M.C.; Hsu, C.H.; Hsiao, Y.R.; Huang, K.Y.; Chuang, T.L.; Yeh, J.M.; Liu, W.R. Enhanced Physical Properties of Electroactive Polyimide Nanocomposites by Addition of Graphene Nanosheets. Polym. Int. 2014, 63, 1011–1017. [Google Scholar] [CrossRef]
  69. Jeon, H.; Lee, K. Effect of gold nanoparticle morphology on thermal properties of polyimide nanocomposite films. Colloids Surf. A Physicochem. Eng. Asp. 2019, 579, 123651. [Google Scholar] [CrossRef]
  70. Wei, Y.; Yang, C.; Ding, T. A one-step method to synthesize n,n’-bis(4’-aminopiienyl)- 1,4-quinonenediimine and its derivatives. Tetrahedron Lett. 1996, 37, 731–734. [Google Scholar] [CrossRef]
  71. Krishnamurthy, S.; Esterle, A.; Sharma, N.C.; Sahi, S.V. Yucca-derived synthesis of gold nanomaterial and their catalytic potential. Nanoscale Res. Lett. 2014, 9, 627. [Google Scholar] [CrossRef] [Green Version]
  72. Sneha, K.; Sathishkumar, M.; Kim, S.; Yun, Y.S. Counter ions and temperature incorporated tailoring of biogenic gold nanoparticles. Process. Biochem. 2010, 45, 1450–1458. [Google Scholar] [CrossRef]
Figure 1. An illustration of the (a) gas sensing set-up: (1) N2 tank, (2) reaction chamber, (3) sample chamber, (4) PI/AuNPs-based H2S gas chemiresistor sensor, (5) reservoir for exhausted gas, (6) multimeter, and (7) computer; and (b) sensor substrate configuration: left-side view, and right-top view (1′) Au discs (end of Au wires), (2′) Teflon block, and (3′) insulated Cu wires.
Figure 1. An illustration of the (a) gas sensing set-up: (1) N2 tank, (2) reaction chamber, (3) sample chamber, (4) PI/AuNPs-based H2S gas chemiresistor sensor, (5) reservoir for exhausted gas, (6) multimeter, and (7) computer; and (b) sensor substrate configuration: left-side view, and right-top view (1′) Au discs (end of Au wires), (2′) Teflon block, and (3′) insulated Cu wires.
Polymers 11 01918 g001
Scheme 1. Proposed mechanism for the synthesis of electroactive polyimide.
Scheme 1. Proposed mechanism for the synthesis of electroactive polyimide.
Polymers 11 01918 sch001
Scheme 2. Proposed mechanism for the synthesis of gold nanoparticle-doped electroactive polyimide.
Scheme 2. Proposed mechanism for the synthesis of gold nanoparticle-doped electroactive polyimide.
Polymers 11 01918 sch002
Figure 2. LC-MS spectra (a) MS+ mode and (b) MS mode, and (c) 1H-NMR spectrum of the synthesized ACAT powder.
Figure 2. LC-MS spectra (a) MS+ mode and (b) MS mode, and (c) 1H-NMR spectrum of the synthesized ACAT powder.
Polymers 11 01918 g002
Figure 3. FTIR spectra of (a) synthesized ACAT, (b) BSAA, (c) synthesized electroactive PI.
Figure 3. FTIR spectra of (a) synthesized ACAT, (b) BSAA, (c) synthesized electroactive PI.
Polymers 11 01918 g003
Figure 4. (a,b) TEM, and (c,d) FE-SEM images of the synthesized electroactive PI/AuNPs.
Figure 4. (a,b) TEM, and (c,d) FE-SEM images of the synthesized electroactive PI/AuNPs.
Polymers 11 01918 g004
Figure 5. XRD pattern of electroactive (a) PI and (b) PI/AuNPs.
Figure 5. XRD pattern of electroactive (a) PI and (b) PI/AuNPs.
Polymers 11 01918 g005
Figure 6. (a) Calibration plot obtained after the electroactive PI/AuNPs-based chemiresistor sensor was exposed towards increasing concentrations of H2S gas at RT (n = 3); and (b) the proposed sensing mechanism.
Figure 6. (a) Calibration plot obtained after the electroactive PI/AuNPs-based chemiresistor sensor was exposed towards increasing concentrations of H2S gas at RT (n = 3); and (b) the proposed sensing mechanism.
Polymers 11 01918 g006aPolymers 11 01918 g006b
Figure 7. (a) A comparison of the calibration plots of electroactive PI/AuNPs-, and PANI/AuNPs-based chemiresistor gas sensors towards increasing concentrations of H2S gas (50 to 300 ppm, n = 3, RT). Repeatability studies based on the calibration plots of (b) PI/AuNPs, and (c) PANI/AuNPs.
Figure 7. (a) A comparison of the calibration plots of electroactive PI/AuNPs-, and PANI/AuNPs-based chemiresistor gas sensors towards increasing concentrations of H2S gas (50 to 300 ppm, n = 3, RT). Repeatability studies based on the calibration plots of (b) PI/AuNPs, and (c) PANI/AuNPs.
Polymers 11 01918 g007aPolymers 11 01918 g007b
Figure 8. Dynamic response and recovery characteristics of (a) PI/AuNPs, and (b) PANI/AuNPs-based chemiresistor sensors towards 150 ppm H2S at RT (cycle = 1).
Figure 8. Dynamic response and recovery characteristics of (a) PI/AuNPs, and (b) PANI/AuNPs-based chemiresistor sensors towards 150 ppm H2S at RT (cycle = 1).
Polymers 11 01918 g008aPolymers 11 01918 g008b
Figure 9. A comparison of the reversibility of the response of (a) PI/AuNPs, and (b) PANI/AuNPs-based chemiresistor sensors towards 200 ppm H2S at RT (cycles = 3).
Figure 9. A comparison of the reversibility of the response of (a) PI/AuNPs, and (b) PANI/AuNPs-based chemiresistor sensors towards 200 ppm H2S at RT (cycles = 3).
Polymers 11 01918 g009
Figure 10. Selectivity test of electroactive PI/AuNPs with 300 ppm concentration of H2S, ethyl acetate (pure), methanol (pure), hexane (pure); (n = 3, RT).
Figure 10. Selectivity test of electroactive PI/AuNPs with 300 ppm concentration of H2S, ethyl acetate (pure), methanol (pure), hexane (pure); (n = 3, RT).
Polymers 11 01918 g010

Share and Cite

MDPI and ACS Style

Padua, L.M.G.; Yeh, J.-M.; Santiago, K.S. A Novel Application of Electroactive Polyimide Doped with Gold Nanoparticles: As a Chemiresistor Sensor for Hydrogen Sulfide Gas. Polymers 2019, 11, 1918. https://doi.org/10.3390/polym11121918

AMA Style

Padua LMG, Yeh J-M, Santiago KS. A Novel Application of Electroactive Polyimide Doped with Gold Nanoparticles: As a Chemiresistor Sensor for Hydrogen Sulfide Gas. Polymers. 2019; 11(12):1918. https://doi.org/10.3390/polym11121918

Chicago/Turabian Style

Padua, Lee Marvin G., Jui-Ming Yeh, and Karen S. Santiago. 2019. "A Novel Application of Electroactive Polyimide Doped with Gold Nanoparticles: As a Chemiresistor Sensor for Hydrogen Sulfide Gas" Polymers 11, no. 12: 1918. https://doi.org/10.3390/polym11121918

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop