Next Article in Journal
Visual Place Recognition Based on Dynamic Difference and Dual-Path Feature Enhancement
Previous Article in Journal
Effectiveness of Mixed Reality in Oral Surgery Training: A Systematic Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Polyaniline/Tungsten Disulfide Composite for Room-Temperature NH3 Detection with Rapid Response and Low-PPM Sensitivity

1
School of Measurement and Control Technology and Communication Engineering, Harbin University of Science and Technology, Harbin 150080, China
2
China Higher Educational Key Laboratory for Measuring & Control Technology and Instrumentations of Heilongjiang Province, Harbin 150080, China
*
Author to whom correspondence should be addressed.
Sensors 2025, 25(13), 3948; https://doi.org/10.3390/s25133948
Submission received: 21 May 2025 / Revised: 21 June 2025 / Accepted: 23 June 2025 / Published: 25 June 2025
(This article belongs to the Section Chemical Sensors)

Abstract

Polyaniline (PANI) is an important conductive-polymer gas-sensing material with working temperature and mechanical flexibilities superior to those of conventional metal oxide sensing materials. However, its applicability is limited by its low sensitivity, high detection limits, and long response/recovery times. In this study, we prepared PANI/WS2 composites via chemical oxidative polymerization and mechanical blending. A multilayer sensor structure—sequentially printed silver-paste heating electrodes, fluorene polyester insulating layer, silver interdigitated electrodes, and sensing material layer—was fabricated on a polyimide substrate via flexible microelectronic printing and systematically characterized using scanning electron microscopy, X-ray diffraction, and Fourier-transform infrared spectroscopy. The optimized 5 wt% WS2 composite showed enhanced gas-sensing performance, with 219.1% sensitivity to 100 ppm ammonia (2.4-fold higher than that of pure PANI) and reduced response and recovery times of 24 and 91 s, respectively (compared to 81 and 436 s for pure PANI, respectively). Notably, the PANI/WS2 sensor detected an ultralow ammonia concentration (100 ppb) with 0.104% sensitivity. The structural characterization and performance analysis results were used to deduce a mechanism for the enhanced sensing capability. These findings highlight the application potential of PANI/WS2 composites in flexible gas sensors and provide fundamental insights for PANI-based sensing materials research.

1. Introduction

Ammonia is a colorless, pungent, and lethal toxic gas. Its primary emission sources are industrial, agricultural, and vehicular exhaust. Through atmospheric circulation, ammonia generates ammonium salts that increase PM2.5 concentrations, posing a significant risk to health [1,2,3,4,5]. Acute exposure to high concentrations of ammonia causes severe physical harm and endangers human health. Prolonged exposure to ammonia gas concentrations exceeding 200 ppm can result in harmful respiratory irritation in adults and children, leading to headaches, nausea, and even life-threatening conditions. Furthermore, ammonia serves as a critical biomarker in medical diagnostics, particularly for liver diseases, Helicobacter pylori-related gastropathy, and kidney disorders. For instance, healthy individuals exhale ammonia at concentrations of approximately 0.3 ppm, while patients with renal dysfunction exhibit elevated levels averaging 0.8 ppm or higher [6,7]. Therefore, ammonia detection is vitally important to both environmental protection and medical diagnostics. Traditional gas analysis methods, such as gas chromatography and mass spectrometry, require bulky instruments, substantial labor, and time costs and suffer from result latency, making real-time monitoring challenging. Consequently, chemiresistive gas sensors with simple structures and real-time detection capabilities have attracted significant research attention. To date, numerous NH3 sensors have been developed, predominantly using metal oxides as the sensing materials. However, challenges persist: preparation processes are complex, costs are high, the sensors have limited flexibility, and power consumption is high owing to elevated operating temperatures. Thus, developing room-temperature ammonia sensors with high sensitivity, short response/recovery times, and flexible structures remains a crucial goal [8,9,10,11].
The selection of sensing materials is paramount in the development of high-performance gas sensors. Metal oxide semiconductors such as SnO2, WO3, and ZnO have been extensively applied to ammonia detection [12,13,14,15]. While these materials demonstrate good sensitivity, they typically require high operating temperatures (leading to high power consumption), exhibit poor selectivity, and lack the mechanical flexibility required for wearable applications. Recently, conductive polymers have gained increasing attention owing to their excellent physicochemical properties. Polyaniline (PANI) has emerged as a promising alternative for ammonia sensing [10,16,17]. PANI can be synthesized via simple chemical oxidative polymerization or electrochemical methods, and its conductivity can be tuned by adjusting oxidant types, reaction durations, and temperatures. Despite advantages like low operating temperatures and environmental stability, PANI-based sensors still suffer from inadequate sensitivity and slow response and recovery in ammonia detection. Modification is reportedly an effective method for enhancing the sensing performance of PANI. Representative modifiers include carbon-based materials and metal oxides. For example, Wan et al. [18] developed a flexible chemiresistive sensor using PANI/carbon nanotube (CNT) composites deposited on PET substrates, achieving responses to 30 to 100 ppm NH3 and a 1 ppm detection limit. Khuspe et al. [19] incorporated SnO2 nanoparticles into PANI via spin-coating, producing composite films with a 30% response to 100 ppm NH3 at room temperature (58% stability). These studies demonstrate that introducing carbon-based materials can provide additional conductive pathways, while the incorporation of n-type semiconductors, owing to the formation of p-n heterojunctions, significantly enhances the sensing performance of PANI.
With advancements in two-dimensional materials research, transition metal dichalcogenides (TMDs) have been shown to have significant potential in catalysis, lubrication, and optoelectronics applications. Compared to traditional metal oxides, TMDs exhibit unique 2D structures, larger specific surface areas, and higher charge carrier mobility, making them promising gas-sensing materials [20,21,22]. Zhao et al. [23] fabricated graphene/WS2 composite films with 2.42% and 1.73% responses to 100 ppm NH3 at 30 and 60 °C, respectively, and the same researchers validated the NH3 sensing capability of WS2 using first-principles calculations. Zhang et al. [24] synthesized MoS2/WS2 heterojunctions via hydrothermal methods, achieving a remarkable 236% response to 500 ppm NH3, a 20-ppm detection limit, and 2.6 s recovery time. Sharma et al. [25] developed WS2 nanosheet-based ammonia sensors with response and recovery times of 54 and 66 s, respectively, to 5 ppm NH3. These results highlight the advantages of WS2 in terms of its carrier mobility and adsorption capacity. However, the low intrinsic conductivity of this material and its poor mechanical flexibility limit its applicability in the design of flexible devices. Although conductive polymers like PANI exhibit excellent compatibility with flexible substrates, their sensitivity and response/recovery performance are unsatisfactory. These facts, as well as the fact that there are limited reports on PANI-TMD composites, necessitate further exploration in this field.
To address these challenges, in this study, we leveraged the complementary properties of PANI and WS2 to fabricate a flexible multilayer ammonia sensor using PANI/WS2 as the sensing material. PANI/WS2 composites were prepared using chemical oxidative polymerization and solution blending. A polyimide substrate was employed to construct the sensor architecture via layer-by-layer printing using a flexible microelectronic printer, incorporating fluorene polyester (FPE) insulation, conductive silver paste heating layers, and interdigitated electrodes. The morphology and structure of the composite were characterized using scanning electron microscopy (SEM), X-ray diffraction (XRD), and Fourier-transform infrared spectroscopy (FTIR). The sensing material was subsequently coated onto the interdigitated electrodes via dispensing printing to complete the flexible sensor. A gas-sensing test platform was used to evaluate the performance of the sensor across NH3 concentrations in the range of 0.1–100 ppm. Finally, the sensing mechanism was elucidated based on the characterization and sensing performance results, highlighting a simple, eco-friendly methodology for future wearable gas detection applications.

2. Experimental Section

2.1. Materials

All chemicals were used as received unless otherwise specified. Aniline (C6H5NH2, analytical reagent (AR) grade; Macklin Biochemical Co., Ltd., Shanghai, China) was distilled at 200 °C under reduced pressure prior to use and stored at low temperature. Ammonium persulfate ((NH4)2S2O8, AR grade), turpentine oil (C10H18O, AR grade), tungsten disulfide (WS2), hydrochloric acid (HCl), N-methylpyrrolidone (NMP, C5H9NO, CAS: 67-44-5, purity ≥99.9%), N,N-dimethylformamide (DMF, CAS: 68-12-2, purity ≥99.9%), and ethanol (C2H5OH, CAS: 64-17-5, purity ≥99.8%) were obtained from Fuchen Chemical Reagent Co., Ltd. (Tianjin, China) and Aladdin Reagent Co., Ltd. (Shanghai, China). Silver conductive paste was procured from Shenzhen Saiya Electronic Paste Co., Ltd. (Shenzhen, China).

2.2. Fabrication of PANI/WS2 Sensing Films

Figure 1 illustrates the synthesis process of PANI. The synthesis was initiated by dissolving distilled aniline (5 mL) in hydrochloric acid (HCl; 1 M, 500 mL) under magnetic stirring for 2 h to form a homogeneous aniline hydrochloride solution. In parallel, ammonium persulfate (11.4 g) was dissolved in HCl (1 M, 50 mL) with continuous agitation for 2 h to obtain the oxidant solution. The aniline solution was subsequently cooled in an ice bath, and the ammonium persulfate oxidant solution was introduced to it dropwise via a burette over a period of 1 h under vigorous stirring. The polymerization was allowed to proceed in the ice bath for 8 h, after which the dark green suspension was vacuum-filtered and the precipitate subjected to five consecutive washing cycles, with 0.1 M HCl and absolute ethanol alternately used as the solvent for washing, to remove residual impurities. Finally, after drying in a vacuum oven at 80 °C for 12 h, the purified PANI was yielded as dark green conductive particles.
Following the drying process, the PANI was combined with tungsten disulfide (WS2) nanosheets via a solution blending method (Figure 2). Precisely weighed PANI and WS2 (at a pre-determined mass ratio) were individually ground in an agate mortar for 1 h. The PANI powder was then dispersed in N,N-dimethylformamide (DMF; 5 mL) and subjected to sequential ultrasonication (40 kHz, 100 W) and magnetic stirring (500 rpm), each for 1 h. The ground WS2 nanosheets were then added to the PANI/DMF suspension, and this was followed by an additional cycle of ultrasonication and magnetic stirring (1 h each) to ensure homogeneous mixing. The mixture was dried in an oven at 80 °C for 24 h, after which the composite was re-ground for 1 h in a mortar with terpineol (2 mL) added dropwise. This final grinding step produced a uniform composite slurry with suitable rheological properties for direct ink writing using a microelectronic dispenser printer.

2.3. Characterization

The microstructural morphology of PANI and the PANI/WS2 composites was examined using field-emission scanning electron microscopy (FESEM; ZEISS Gemini SEM 300, Oberkochen, Germany) at accelerating voltages in the range of 3–15 kV. Crystal structure analysis was performed based on X-ray diffraction (XRD, Rigaku SmartLab 3 kW, Tokyo, Japan) with Cu-Kα radiation (λ = 1.5406 Å); the 2θ range was 20–80° and the scan rate 2° min−1. Functional group evolution was investigated using Fourier-transform infrared spectroscopy (FTIR, PerkinElmer Frontier 100, Waltham, MA, USA) in attenuated total reflectance (ATR) mode, with spectral acquisition spanning 400–4000 cm−1 at a resolution of 4 cm−1 (32 scans per measurement).

2.4. Design and Fabrication of Sensor Structure and Construction of Gas-Sensing Test Platform

We manufactured flexible gas sensors using a microelectronic distributor printer (Scientific 3, Zhongbin Technology Co., Ltd., Shanghai, China). The device structure consisted of four sequentially stacked layers on a polyimide substrate (Figure 3a): a flexible printed silver heating electrode, an FPE insulation layer, a silver forked electrode (IDE), and a composite sensing-material coating. To form the insulation layer, a homogeneous solution was prepared by dissolving FPE particles (0.4 g) in 4 mL of NMP, followed by sonicating the mixture for 2 h and magnetically stirring it for 3 h. Next, the coating module of the printer was used to coat the FPE solution blade onto the heating electrode (avoiding contact pads), and it was subsequently cured in a vacuum oven at 80 °C for 24 h. After patterning the IDE, the PANI/WS2 composite slurry (prepared using the method described in Section 2.2) was evenly distributed onto the IDE and dried at 80 °C for 12 h. The completed sensor was aged at 30 °C for 5 days to stabilize its performance.
The gas-sensing test platform comprised a sealed chamber connected to a DC power supply, data acquisition system, and computer (Figure 3b). Specific concentrations of ammonia gas were injected into the chamber using a gas-tight syringe. The heating electrode maintained the sensor at ambient temperature during testing. Resistance changes induced by the interaction of ammonia with the sensing layer were recorded via the IDE and analyzed using dedicated software to extract the response/recovery parameters. The physical diagram of the sensor structure is shown in Figure S1.

3. Results and Discussion

3.1. Characterization of Composite Materials

Figure 4(a1,a2) display scanning electron microscopy (SEM) images of pristine PANI at 1 μm and 500 nm magnifications. The as-synthesized PANI sample is characterized by irregular particle agglomeration and minimum particle dimensions of approximately 80 nm. Figure 4(b1–b3) show SEM images of the PANI/WS2 (5 wt%) composite at different scales. The solution blending method, coupled with ultrasonication and magnetic stirring, achieved enhanced dispersion homogeneity. Pure PANI exhibits an irregular aggregation state (Figure 4(a1,a2)), with a minimum particle size of approximately 80 nm, and pores with sizes of the order of micrometers on the surface. In the 5 wt% WS2 composite (Figure 4(b1–b3)), WS2 nanosheets are embedded into the PANI matrix, forming a micro/nano multilevel pore structure with an increased specific surface area. Notably, increasing the WS2 content to 7 wt% may lead to agglomeration (see Supplementary Materials, Figure S3), which may cause pore blockage and deteriorate the sensing performance. Figure 4(c1–c4) shows energy-dispersive spectroscopy (EDS) elemental maps of the 5 wt% WS2 composite, confirming the uniformity of the PANI distribution and successful WS2 incorporation within the composite, both of which are expected to lead to improved gas sensing capabilities. The cross-sectional image of the sensitive film is shown in Figure S2.
The XRD patterns of the different PANI/WS2 composites (Figure 5a) reveal a distinct structural evolution trend. The XRD pattern of pristine PANI features a broad amorphous halo at 20–30° (2θ) that is characteristic of the disordered polymer chains. With the incorporation of WS2 into the material, sharp diffraction peaks emerge at 14.4°, 29°, 32°, 33°, 39°, 44°, 49°, 58°, and 60° that are assignable to the (002), (004), (100), (101), (103), (006), (105), (110), and (112) crystallographic planes, respectively, of hexagonal WS2 (JCPDS 08-0237). In the XRD profiles, the position of the characteristic peak of WS2 (such as 14.4° (002)) in the composite material is consistent with that of pure WS2, indicating the absence of lattice distortion. The intensity of the amorphous dispersion peak of PANI decreases with increasing WS2 content, indicating that WS2 restricts the movement of PANI segments [26,27,28].
FTIR analysis (Figure 5b) allowed us to evaluate the molecular-level interactions in the composites. The broad N–H stretching vibration, at ~3400 cm−1 in pure PANI, gradually broadens with WS2 loading, suggesting that the hydrogen bonding was strengthened. The characteristic C=C stretching vibrations of benzenoid (1487 cm−1) and quinoid (1567 cm−1) rings undergo 5–8 cm−1 red shifts, which are indicative of enhanced π-π stacking between WS2 and the conjugated PANI backbone. The C=C bond redshift (benzene ring 1487 → 1482 cm−1, quinone ring 1567 → 1562 cm−1) indicates that the PANI aromatic ring forms π–π stacking with the WS2 sulfur plane in a face-to-face configuration, with a spacing of ~0.34 nm (typical π stacking distance) accommodated by the (002) interplanar spacing (0.62 nm) of layered WS2. The invariance of the in-plane C–H bending vibration at 1130 cm−1 confirms that the planarity of the polymer backbone is preserved. The FTIR spectra of samples, containing ≤3 wt% WS2, show only minor changes in peak intensities (<5%), and peak-shift or new peaks are not observed. This result indicates the absence of strong chemical bonds and suggests that van der Waals forces dominate the interactions. [29,30,31].

3.2. Sensor Performance Testing

The gas sensing performance of the prepared composite sensors with WS2 mass fractions of 0, 1, 3, 5, and 7 wt% was evaluated. Each sensor was placed in a gas chamber and tested under static conditions by sequentially exposing it to gaseous ammonia at concentrations of 1, 5, 10, 20, 50, and 100 ppm. The temperature was maintained at 25 ± 2 °C, and the relative humidity was 37 ± 3%. The sensitivity of the sensor was defined in terms of the sensor response as follows:
R e s p o n s e   % = R g R a R a × 100 %
where Rg is the resistance of the sensor in ammonia gas and Ra is the baseline resistance, i.e., the resistance in clean air.
The sensitivity curves of sensors with different compositions, shown in Figure 6a, indicate that the response to ammonia is concentration dependent. All the composites exhibit p-type semiconductor behavior when exposed to the reducing gas. The pure PANI sensor (0 wt% WS2) shows the lowest sensitivity (92.1% at 100 ppm), a prolonged response time (81 s), and a very prolonged recovery time (436 s), which significantly limit its practical applicability.
With increasing WS2 content over the range of 0–5 wt%, both the sensitivity and the response/recovery times progressively improve. The best performance was realized using the 5 wt% composite: 219.1% sensitivity at 100 ppm NH3, representing a 2.38-fold enhancement over that of pure PANI, along with a 3.3-fold shorter response time (24 s) and 4.8-fold shorter recovery time (91 s). However, at 7 wt% WS2, the sensitivity dropped to 196%, and the recovery was slower (95 s), which were attributed to a reduction in surface porosity owing to nanosheet agglomeration (Figure 6b).
To investigate the gas sensing capabilities at trace levels, the responses of the pure PANI and PANI/WS2 (5 wt%) sensors were compared at 1 ppm NH3 (Figure 7a,b). The pure PANI sensor exhibited a response value of 0.24% with long response and recovery times (69 and 415 s, respectively). In contrast, the PANI/WS2 (5 wt%) composite demonstrated significantly enhanced performance: a 5.37-fold higher response (1.29%), 3.28-fold shorter response time (21 s), and 7.54-fold shorter recovery time (55 s). This improvement is attributed to the hierarchical porosity of the composite material and WS2-induced heterojunction formation, which synergistically enhance gas adsorption kinetics and charge transfer efficiency.
The trace-level detection capability of the PANI/WS2 (5 wt% WS2) composite sensor was evaluated (Figure 8a). At 100 ppb ammonia, the sensor exhibited a detectable response of 0.104%, which increased to 0.871% at 800 ppb. Although a slight baseline drift was observed below 1 ppm, stable sensing, within an acceptable range, was maintained. Selectivity testing of the same sensor (Figure 8b) revealed distinct responses to 100 ppm of various gases, NH3 (218%), SO2 (9%), H2S (16.5%), acetic acid (5.54%), acetone (6%), and ethanol (7.5%), confirming the excellent ammonia selectivity of the sensor. Mechanical flexibility tests (Figure 8c) demonstrated consistent performance under bending stress. At bending angles of 30°, 60°, 90°, 120°, and 150°, the sensor showed average responses of 218.95%, 219.85%, 220.45%, 221.35%, and 223.15%, respectively, to 100 ppm NH3. The marginal response increase with larger angles may arise from tensile strain-induced surface area expansion in the composite. Repeatability testing (Figure 8d) over eight cycles at 100 ppm NH3 yielded an average response of 219% (RSD = 1.2%), confirming that the operational stability of the composite sensor was excellent.
The PANI/WS2 (5 wt%) composite sensor was tested at different temperatures (25, 35, 45, and 55 °C) in 1 ppm NH3 atmosphere (Figure 9a). The sensitivity decreased from 1.29% to 0.63%, while response times were shortened from 21 s to 12 s and recovery times from 55 s to 31 s as the temperature was increased over this range. This trend is attributed to the enhanced thermal motion of NH3 molecules at higher temperatures, which reduces adsorption stability but accelerates desorption kinetics. The sensor maintained reliable detection at 1 ppm NH3 over this temperature range from room temperature to 55 °C, suggesting that it could have the desired flexibility characteristics for applications requiring rapid responses when integrated with heating modules. Humidity variation in the range of 35–75% RH produced minimal changes in the response (218.83–220.11%) to 100 ppm NH3 (Figure S4). A 30-day long-term stability assessment result showed that the sensitivity of the composite sensor decreased by 3%, while that of the pure PANI sensor decreased by 11.4%. (Figure 9b). Comparing the results obtained in this study with those reported in the literature, the advantages of the composite in terms of its sensitivity, lower detection limit, and shorter response/recovery times are apparent (Table 1).

3.3. NH3 Sensing Mechanism of PANI/WS2

Although NH3 exhibits reducibility, its deprotonation dominates the response in the PANI system: NH3 binds to H+ to reduce the concentration of polarons, leading to an increase in resistance (p-type behavior). The n-type characteristic of WS2 amplifies this effect through the heterojunction. The intrinsic insulator PANI undergoes protonation in the presence of dilute hydrochloric acid during synthesis. The acidic environment facilitates protonation at the quinoid ring nitrogen atoms, converting quinoid segments into benzenoid structures. This structural rearrangement produces p-type semiconductor behavior under an external electric field owing to hole-dominated charge transport. When exposed to the alkali ammonia (NH3), the conductive PANI matrix is deprotonated, reducing the hole concentration and increasing the electrical resistance of the material. Conversely, upon NH3 removal, ammonium ions (NH4+) decompose into NH3 and protons (H+), and the conductivity of PANI is restored. The reversible reaction mechanism is expressed as follows [36,37,38]:
P A N I + N H P A N I + N H
As demonstrated by the results shown in Figure 8 and discussed above, the introduction of WS2 nanosheets into PANI significantly improves sensor performance. First, a p-n heterojunction is formed at the interface between p-type PANI and n-type WS2. According to previous studies, PANI has a bandgap of 2.8 eV (p-type semiconductor), while that of WS2 is 2.07 eV (n-type semiconductor). As shown in Figure 10, in the p-n heterojunction, electrons and holes move in opposite directions at the interface, forming a depletion layer. As carriers migrate, the Fermi level eventually equilibrates. The incorporation of WS2 introduces a higher potential barrier, resulting in increased resistance for the composite compared to that of pure PANI.
In the presence of ammonia, NH3 molecules neutralize protons in conductive PANI to form NH4+. Simultaneously, the depletion layer at the heterojunction interface provides additional holes for interaction with NH3 molecules, enhancing sensitivity. Furthermore, the SEM images of the composites reveal that the embedded WS2 nanosheets increase the PANI dispersion, leading to a more uniform distribution. The enhanced dispersion results in greater porosity, and the two-dimensional sheet structure provides a larger specific surface area. Compared to pure PANI, the composite offers more adsorption sites and therefore faster response and recovery. In summary, appropriate WS2 doping optimizes the gas sensing performance of PANI/WS2 composites.

4. Conclusions

In this study, a flexible ammonia sensor operable at room temperature based on a PANI/WS2 composite was realized. The composites were synthesized via chemical oxidative polymerization and mechanical blending, making use of microelectronic printing technology. The optimized composite with 5 wt% WS2 demonstrated significantly enhanced gas sensing performance: the sensitivity was 219.1% at 100 ppm NH3 (a 2.38-fold improvement over that of pure PANI), with response and recovery times reduced to 24 and 91 s (3.3- and 4.8-fold shorter, respectively, than the times for pure PANI). The sensor exhibited detectable responses to 100 ppb NH3 (0.104%), along with exceptional selectivity (an NH3 response 24–39 times higher than interferent gas responses), mechanical flexibility (stable sensitivity during 120° bending), and long-term stability (a 3% sensitivity drop over 30 days). Mechanistic studies revealed that p-n heterojunction formation owing to WS2 incorporation enhances carrier migration and surface adsorption, while the two-dimensional structure of WS2 improves PANI dispersion and porosity, synergistically boosting the sensitivity and response kinetics of the sensor. This work provides a novel general strategy for the development of highly sensitive, low-power flexible gas sensors for wearable healthcare monitoring and environmental detection applications based on heterojunction interface optimization or hybridization with 2D materials.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/s25133948/s1: Figure S1: Physical structure diagram of flexible gas sensor; Figure S2: Cross section of sensitive film; Figure S3: Scanning electron microscopy (SEM) image of PANI/WS2 composite with 7 wt% WS2 loading; Figure S4: Humidity-dependent sensitivity of 5 wt% PANI/WS2 composite sensor exposed to 100 ppm NH3.

Author Contributions

Conceptualization, Y.S.; methodology, B.T.; software, Q.C.; investigation, C.Z.; data curation, H.N. and Q.C.; writing—original draft preparation, H.N. and J.L.; writing—review and editing, K.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This work is supported by a grant from the National Natural Science Foundation of China (Grant no. 62271176) and China National Basic Enhancement Program (2022-JCJQ-JJ-0438).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The original contributions presented in this study are included in the article. Further inquiries can be directed to the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Zhang, X.; Davidson, E.A.; Mauzerall, D.L.; Searchinger, T.D.; Dumas, P.; Shen, Y. Managing nitrogen for sustainable development. Nature 2015, 528, 51–59. [Google Scholar] [CrossRef] [PubMed]
  2. Wu, M.; He, M.; Hu, Q.; Wu, Q.; Sun, G.; Xie, L.; Zhang, Z.; Zhu, Z.; Zhou, A. Ti3C2 MXene-based sensors with high selectivity for NH3 detection at room temperature. ACS Sens. 2019, 4, 2763–2770. [Google Scholar] [CrossRef] [PubMed]
  3. Hadano, F.S.; Gavim, A.E.X.; Stefanelo, J.C.; Gusso, S.L.; Macedo, A.G.; Rodrigues, P.C.; Yusoff, A.R.b.M.; Schneider, F.K.; de Deus, J.F.; da Silva, W.J. NH3 sensor based on rGO-PANI composite with improved sensitivity. Sensors 2021, 21, 4947. [Google Scholar] [CrossRef] [PubMed]
  4. Erisman, J.W.; Sutton, M.A.; Galloway, J.; Klimont, Z.; Winiwarter, W. How a century of ammonia synthesis changed the world. Nat. Geosci. 2008, 1, 636–639. [Google Scholar] [CrossRef]
  5. Balamurugan, C.; Lee, D.W. A selective NH3 gas sensor based on mesoporous p-type NiV2O6 semiconducting nanorods synthesized using solution method. Sens. Actuators B 2014, 192, 414–422. [Google Scholar] [CrossRef]
  6. Wu, G.; Du, H.; Cha, Y.L.; Lee, D.; Kim, W.; Feyzbar-Khalkhali-Nejad, F.; Oh, T.-S.; Zhang, X.; Kim, D.-J. A wearable mask sensor based on polyaniline/CNT nanocomposites for monitoring ammonia gas and human breathing. Sens. Actuators B 2023, 375, 132858. [Google Scholar] [CrossRef]
  7. Liu, L.; Fei, T.; Guan, X.; Lin, X.; Zhao, H.; Zhang, T. Room temperature ammonia gas sensor based on ionic conductive biomass hydrogels. Sens. Actuators B 2020, 320, 128318. [Google Scholar] [CrossRef]
  8. Milani Moghaddam, H.; Malkeshi, H. Self-assembly synthesis and ammonia gas-sensing properties of ZnO/Polythiophene nanofibers. J. Mater. Sci. Mater. Electron. 2016, 27, 8807–8815. [Google Scholar] [CrossRef]
  9. Jia, X.-S.; Tang, C.-C.; Yan, X.; Yu, G.-F.; Li, J.-T.; Zhang, H.-D.; Li, J.-J.; Gu, C.-Z.; Long, Y.-Z. Flexible polyaniline/poly(methyl methacrylate) composite fibers via electrospinning and in situ polymerization for ammonia gas sensing and strain sensing. J. Nanomater. 2016, 2016, 102828. [Google Scholar] [CrossRef]
  10. Bittencourt, J.C.; de Santana Gois, B.H.; Rodrigues de Oliveira, V.J.; da Silva Agostini, D.L.; de Almeida Olivati, C. Gas sensor for ammonia detection based on poly(vinyl alcohol) and polyaniline electrospun. J. Appl. Polym. Sci. 2019, 136, 47288. [Google Scholar] [CrossRef]
  11. Wu, Q.; Shen, W.; Lv, D.; Chen, W.; Song, W.; Tan, R. An enhanced flexible room temperature ammonia gas sensor based on GP-PANI/PVDF multi-hierarchical nanocomposite film. Sens. Actuators B 2021, 334, 129630. [Google Scholar] [CrossRef]
  12. Kwak, D.; Lei, Y.; Maric, R. Ammonia gas sensors: A comprehensive review. Talanta 2019, 204, 713–730. [Google Scholar] [CrossRef] [PubMed]
  13. Van Toan, N.; Hung, C.M.; Hoa, N.D.; Van Duy, N.; Thi Thanh Le, D.; Thi Thu Hoa, N.; Viet, N.N.; Phuoc, P.H.; Van Hieu, N. Enhanced NH3 and H2 gas sensing with H2S gas interference using multilayer SnO2/Pt/WO3 nanofilms. J. Hazard. Mater. 2021, 412, 125181. [Google Scholar] [CrossRef] [PubMed]
  14. Timmer, B.; Olthuis, W.; van den Berg, A. Ammonia sensors and their applications—A review. Sens. Actuators B 2005, 107, 666–677. [Google Scholar] [CrossRef]
  15. Li, Y.; Zhang, Y.; Zhou, Y.; Zhao, L.; Yan, X.; Liu, F.; Lu, G.; Sun, P. wearable gas sensor based on reticular antimony-doped SnO2/PANI nanocomposite realizing intelligent detection of ammonia within a wide range of humidity. ACS Sens. 2023, 8, 4132–4142. [Google Scholar] [CrossRef]
  16. Tanguy, N.R.; Thompson, M.; Yan, N. A review on advances in application of polyaniline for ammonia detection. Sens. Actuators B 2018, 257, 1044–1064. [Google Scholar] [CrossRef]
  17. Tian, X.; Cui, X.; Xiao, Y.; Chen, T.; Xiao, X.; Wang, Y. Pt/MoS2/polyaniline nanocomposite as a highly effective room temperature flexible gas sensor for ammonia detection. ACS Appl. Mater. Interfaces 2023, 15, 9604–9617. [Google Scholar] [CrossRef]
  18. Wan, P.; Wen, X.; Sun, C.; Chandran, B.K.; Zhang, H.; Sun, X.; Chen, X. Flexible transparent films based on nanocomposite networks of polyaniline and carbon nanotubes for high-performance gas sensing. Small 2015, 11, 5409–5415. [Google Scholar] [CrossRef]
  19. Khuspe, G.D.; Navale, S.T.; Bandgar, D.K.; Sakhare, R.D.; Chougule, M.A.; Patil, V.B. SnO2 nanoparticles-modified polyaniline films as highly selective, sensitive, reproducible and stable ammonia sensors. Electron. Mater. Lett. 2014, 10, 191–197. [Google Scholar] [CrossRef]
  20. Ouyang, C.; Chen, Y.; Qin, Z.; Zeng, D.; Zhang, J.; Wang, H.; Xie, C. Two-dimensional WS2-based nanosheets modified by Pt quantum dots for enhanced room-temperature NH3 sensing properties. Appl. Surf. Sci. 2018, 455, 45–52. [Google Scholar] [CrossRef]
  21. Qin, Z.; Song, X.; Wang, J.; Li, X.; Wu, C.; Wang, X.; Yin, X.; Zeng, D. Development of flexible paper substrate sensor based on 2D WS2 with S defects for room-temperature NH3 gas sensing. Appl. Surf. Sci. 2022, 573, 151535. [Google Scholar] [CrossRef]
  22. Liang, K.; Yin, M.; Ma, D.; Fan, Y.; Li, Z. Facile preparation and photocatalytic hydrogen production of WS2 and its composites. Int. J. Hydrogen Energy 2022, 47, 38622–38634. [Google Scholar] [CrossRef]
  23. Zhao, F.; Li, Z.; Fu, Y.; Wang, Q. Gas-sensitive characteristics of graphene composite tungsten disulfide to ammonia. Sensors 2022, 22, 8672. [Google Scholar] [CrossRef]
  24. Zhang, M.; Zhang, J. Highly selective NH3 sensor based on MoS2/WS2 heterojunction. Nanomaterials 2023, 13, 1835. [Google Scholar] [CrossRef] [PubMed]
  25. Sharma, S.; Saini, R.; Gupta, G.; Late, D.J. Room-temperature highly sensitive and selective NH3 gas sensor using vertically aligned WS2 nanosheets. Nanotechnology 2023, 34, 045704. [Google Scholar] [CrossRef]
  26. Ko, K.Y.; Song, J.G.; Kim, Y.; Choi, T.; Shin, S.; Lee, C.W.; Lee, K.; Koo, J.; Lee, H.; Kim, J.; et al. Improvement of gas-sensing performance of large-area tungsten disulfide nanosheets by surface functionalization. ACS Nano 2016, 10, 9287–9296. [Google Scholar] [CrossRef]
  27. Wang, X.; Gong, L.; Zhang, D.; Fan, X.; Jin, Y.; Guo, L. Room temperature ammonia gas sensor based on polyaniline/copper ferrite binary nanocomposites. Sens. Actuators B 2020, 322, 128615. [Google Scholar] [CrossRef]
  28. Xu, W.; Kozawa, D.; Zhou, Y.; Wang, Y.; Sheng, Y.; Jiang, T.; Strano, M.S.; Warner, J.H. Controlling photoluminescence enhancement and energy transfer in WS2:hBN:WS2 vertical stacks by precise interlayer distances. Small 2020, 16, e1905985. [Google Scholar] [CrossRef]
  29. Konwer, S.; Guha, A.K.; Dolui, S.K. Graphene oxide-filled conducting polyaniline composites as methanol-sensing materials. J. Mater. Sci. 2013, 48, 1729–1739. [Google Scholar] [CrossRef]
  30. Zhang, J.; Zhang, M.; Wang, S.; Wu, Z.; Zhang, Z. Moisture-resistant and highly selective NH3 sensor based on CdS/WS2 composite heterojunction. Langmuir 2023, 39, 10352–10366. [Google Scholar] [CrossRef]
  31. Li, S.; Liu, A.; Yang, Z.; Zhao, L.; Wang, J.; Liu, F.; You, R.; He, J.; Wang, C.; Yan, X.; et al. Design and preparation of the WO3 hollow spheres@PANI conducting films for room temperature flexible NH3 sensing device. Sens. Actuators B 2019, 289, 252–259. [Google Scholar] [CrossRef]
  32. Khuspe, G.D.; Bandgar, D.K.; Sen, S.; Patil, V.B. Fussy nanofibrous network of polyaniline (PANi) for NH3 detection. Synth. Met. 2012, 162, 1822–1827. [Google Scholar] [CrossRef]
  33. He, M.; Xie, L.; Luo, G.; Li, Z.; Wright, J.; Zhu, Z. Flexible fabric gas sensors based on PANI/WO3 p-n heterojunction for high performance NH3 detection at room temperature. Sci. China Mater. 2020, 63, 2028–2039. [Google Scholar] [CrossRef]
  34. Yang, R.; Zhang, J.; Liu, J.; Li, G.; Qiao, Y.; Zhang, X.; Gao, J.; Lu, H. PANI/Ti3C2Tx composite nanofiber-based flexible conductometric sensor for the detection of NH3 at room temperature. Sens. Actuators B 2023, 392, 134128. [Google Scholar] [CrossRef]
  35. Ma, J.; Fan, H.; Li, Z.; Jia, Y.; Yadav, A.K.; Dong, G.; Wang, W.; Dong, W.; Wang, S. Multi-walled carbon nanotubes/polyaniline on the ethylenediamine modified polyethylene terephthalate fibers for a flexible room temperature ammonia gas sensor with high responses. Sens. Actuators B 2021, 334, 129677. [Google Scholar] [CrossRef]
  36. Luo, G.; Xie, L.; He, M.; Jaisutti, R.; Zhu, Z. Fabric gas sensors based on reduced graphene-polyaniline nanocomposite for highly sensitive NH3 detection at room temperature. Nanotechnology 2021, 32, 305501. [Google Scholar] [CrossRef]
  37. Li, S.; Wang, T.; Yang, Z.; He, J.; Wang, J.; Zhao, L.; Lu, H.; Tian, T.; Liu, F.; Sun, P.; et al. Room temperature high performance NH3 sensor based on GO-rambutan-like polyaniline hollow nanosphere hybrid assembled to flexible PET substrate. Sens. Actuators B 2018, 273, 726–734. [Google Scholar] [CrossRef]
  38. Zhang, H.; Zhang, X.; Qiu, C.; Jia, P.; An, F.; Zhou, L.; Zhu, L.; Zhang, D. Polyaniline/ZnO heterostructure-based ammonia sensor self-powered by electrospinning of PTFE-PVDF/MXene piezo-tribo hybrid nanogenerator. Chem. Eng. J. 2024, 496, 154226. [Google Scholar] [CrossRef]
Figure 1. Schematic of polyaniline (PANI) preparation via chemical oxidative polymerization.
Figure 1. Schematic of polyaniline (PANI) preparation via chemical oxidative polymerization.
Sensors 25 03948 g001
Figure 2. Fabrication process of PANI/WS2 composite sensing material.
Figure 2. Fabrication process of PANI/WS2 composite sensing material.
Sensors 25 03948 g002
Figure 3. Sensor structure and gas-sensing test platform.
Figure 3. Sensor structure and gas-sensing test platform.
Sensors 25 03948 g003
Figure 4. Scanning electron microscopy (SEM) images of (a1,a2) PANI and (b1b3) of 5 wt% WS2 PANI/WS2 composite. (c1c4) Energy-dispersive spectroscopy (EDS) elemental maps of the 5 wt% WS2 PANI/WS2 composite.
Figure 4. Scanning electron microscopy (SEM) images of (a1,a2) PANI and (b1b3) of 5 wt% WS2 PANI/WS2 composite. (c1c4) Energy-dispersive spectroscopy (EDS) elemental maps of the 5 wt% WS2 PANI/WS2 composite.
Sensors 25 03948 g004
Figure 5. (a) X-ray diffraction (XRD) and (b) Fourier-transform infrared (FTIR) spectra of the composite material.
Figure 5. (a) X-ray diffraction (XRD) and (b) Fourier-transform infrared (FTIR) spectra of the composite material.
Sensors 25 03948 g005
Figure 6. (a) Response to different concentrations of ammonia of sensors made from PANI/WS2 composites with different WS2 contents. (b) Sensitivity, response time, and recovery time versus WS2 content of the composite in the sensor.
Figure 6. (a) Response to different concentrations of ammonia of sensors made from PANI/WS2 composites with different WS2 contents. (b) Sensitivity, response time, and recovery time versus WS2 content of the composite in the sensor.
Sensors 25 03948 g006
Figure 7. Response-recovery curves of (a) pure PANI and (b) 5% WS2 composite sensor at 1 ppm NH3.
Figure 7. Response-recovery curves of (a) pure PANI and (b) 5% WS2 composite sensor at 1 ppm NH3.
Sensors 25 03948 g007
Figure 8. Ammonia sensing performance characterization of PANI/WS2 (5 wt% WS2) composite sensor. (a) Low-concentration response curves. (b) Selectivity, (c) bending angle, and (d) repeatability test results at 100 ppm NH3.
Figure 8. Ammonia sensing performance characterization of PANI/WS2 (5 wt% WS2) composite sensor. (a) Low-concentration response curves. (b) Selectivity, (c) bending angle, and (d) repeatability test results at 100 ppm NH3.
Sensors 25 03948 g008
Figure 9. (a) Gas sensing performance at various temperatures. (b) Long-term stability results.
Figure 9. (a) Gas sensing performance at various temperatures. (b) Long-term stability results.
Sensors 25 03948 g009
Figure 10. Ammonia sensing mechanism of the sensor.
Figure 10. Ammonia sensing mechanism of the sensor.
Sensors 25 03948 g010
Table 1. Performance of several common types of NH3 gas sensor.
Table 1. Performance of several common types of NH3 gas sensor.
MaterialDetection LimitTemperatureConcentrationResponseResponse/Recovery TimeReference
PANI RT100 ppm96%122 s/1235 s[32]
PANI/WO33 ppmRT100 ppm150%122 s/165 s[33]
PANI/Ti3C2TX5 ppmRT20 ppm55.9%40 s/200 s[34]
PANI/MWCNTs33 ppmRT50 ppm117%47 s/~ s[35]
rGO-PANI1 ppmRT100 ppm6.2%219 s/541 s[36]
PANI/WS20.1 ppmRT100 ppm219.1%24 s/91 sThis work
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zhao, K.; Shi, Y.; Niu, H.; Chen, Q.; Liu, J.; Tang, B.; Zheng, C. Polyaniline/Tungsten Disulfide Composite for Room-Temperature NH3 Detection with Rapid Response and Low-PPM Sensitivity. Sensors 2025, 25, 3948. https://doi.org/10.3390/s25133948

AMA Style

Zhao K, Shi Y, Niu H, Chen Q, Liu J, Tang B, Zheng C. Polyaniline/Tungsten Disulfide Composite for Room-Temperature NH3 Detection with Rapid Response and Low-PPM Sensitivity. Sensors. 2025; 25(13):3948. https://doi.org/10.3390/s25133948

Chicago/Turabian Style

Zhao, Kuo, Yunbo Shi, Haodong Niu, Qinglong Chen, Jinzhou Liu, Bolun Tang, and Canda Zheng. 2025. "Polyaniline/Tungsten Disulfide Composite for Room-Temperature NH3 Detection with Rapid Response and Low-PPM Sensitivity" Sensors 25, no. 13: 3948. https://doi.org/10.3390/s25133948

APA Style

Zhao, K., Shi, Y., Niu, H., Chen, Q., Liu, J., Tang, B., & Zheng, C. (2025). Polyaniline/Tungsten Disulfide Composite for Room-Temperature NH3 Detection with Rapid Response and Low-PPM Sensitivity. Sensors, 25(13), 3948. https://doi.org/10.3390/s25133948

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