Next Article in Journal
Synchronously Strengthen and Toughen Polypropylene Using Tartaric Acid-Modified Nano-CaCO3
Next Article in Special Issue
A Study on Machine Learning Methods’ Application for Dye Adsorption Prediction onto Agricultural Waste Activated Carbon
Previous Article in Journal
pNiPAM-Nanoparticle-Based Antiapoptotic Approach for Pro-Regenerative Capacity of Skeletal Myogenic Cells
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Emerging MXene–Polymer Hybrid Nanocomposites for High-Performance Ammonia Sensing and Monitoring

1
Research Cell and Department of Physics, Bhagini Nivedita College, University of Delhi, New Delhi 110045, India
2
Centre for Neural and Cognitive Sciences, University of Hyderabad, Hyderabad 500046, India
3
Mads Clausen Institute, NanoSYD, University of Southern Denmark, Alsion 2, 6400 Sønderborg, Denmark
4
NanoBioTech Laboratory, Health System Engineering, Department of Environmental Engineering, Florida Polytechnic University, Lakeland, FL 33805, USA
*
Authors to whom correspondence should be addressed.
Nanomaterials 2021, 11(10), 2496; https://doi.org/10.3390/nano11102496
Submission received: 3 August 2021 / Revised: 12 September 2021 / Accepted: 18 September 2021 / Published: 24 September 2021
(This article belongs to the Special Issue Applications of Nanomaterials in Environmental Science)

Abstract

:
Ammonia (NH3) is a vital compound in diversified fields, including agriculture, automotive, chemical, food processing, hydrogen production and storage, and biomedical applications. Its extensive industrial use and emission have emerged hazardous to the ecosystem and have raised global public health concerns for monitoring NH3 emissions and implementing proper safety strategies. These facts created emergent demand for translational and sustainable approaches to design efficient, affordable, and high-performance compact NH3 sensors. Commercially available NH3 sensors possess three major bottlenecks: poor selectivity, low concentration detection, and room-temperature operation. State-of-the-art NH3 sensors are scaling up using advanced nano-systems possessing rapid, selective, efficient, and enhanced detection to overcome these challenges. MXene–polymer nanocomposites (MXP-NCs) are emerging as advanced nanomaterials of choice for NH3 sensing owing to their affordability, excellent conductivity, mechanical flexibility, scalable production, rich surface functionalities, and tunable morphology. The MXP-NCs have demonstrated high performance to develop next-generation intelligent NH3 sensors in agricultural, industrial, and biomedical applications. However, their excellent NH3-sensing features are not articulated in the form of a review. This comprehensive review summarizes state-of-the-art MXP-NCs fabrication techniques, optimization of desired properties, enhanced sensing characteristics, and applications to detect airborne NH3. Furthermore, an overview of challenges, possible solutions, and prospects associated with MXP-NCs is discussed.

1. Exploring Needs of Ammonia Detection and Monitoring: Introduction

Ammonia (NH3) is extensively used in various commercial products and manufacturing industries, including refrigerants, fertilizers, explosives, hydrogen storage and production, pharmaceuticals, packaging, and chemicals [1,2]. As per statistics, NH3 was the second-highest produced chemical commodity, with 235 million tons in 2019 and is predicted to reach around 290 million tons in 2030 [1]. However, NH3 is highly toxic to human health and a major environmental contaminant. As per the Occupational Safety and Health Administration (OSHA), the maximum exposure limit of NH3 at 25 ppm is 8 h, and when reaching 35 ppm, is 10 min at the workplace [3]. The NH3 transforms into ammonium salt on inhalation or absorption through the eyes, nose, skin, or respiratory tract. It can cause illness, lung edema, renal dysfunction, nervous disorder, and mortality exceeding 200–500 mg/kg of body weight [3].
Moreover, NH3 is majorly produced from the agricultural sector due to fertilizers at large scale, livestock production, and low nitrogen use efficiency [4]. Airborne NH3 possesses a relationship with spatiotemporal disparity of particulate matter (PM 2.5) levels, which is dangerous for public health on penetrating the human body, especially during coronavirus disease (COVID-19) [5,6]. Furthermore, NH3 is found in exhaled human breath and acts as a crucial biomarker for diagnosing gastric ulcers and renal disorder caused by Helicobacter pylori infection, and recently COVID-19 [7,8]. Renal failure causes the accumulation of urea in the form of ammonium ions in the blood, which penetrate through the lung membrane. As a result, the exhaled breath of patients with peptic ulcer and renal disorder carries NH3 in a concentration range of 0.82 to 14.7 ppm [9]. Hence, real-time monitoring of low-level NH3 emission has become a priority area for lifesaving, environmental well-being, and biomedical applications. The need of NH3 sensing and its state-of-the-art sensing approaches are illustrated in Figure 1.
Various NH3 detecting techniques (Figure 2) include calorimetric, gravimetric, electrochemical, optical, chromatographic, laser-based, catalytic, resistive, acoustic, and quartz crystal microbalance-based methods [10,11,12,13,14,15,16,17,18]. Among them, chemiresistors have become critical NH3 sensing technologies due to their miniaturized design, low-cost, user-friendly, portable, and energy efficiency to detect NH3 (at ~1ppb) in human breath, workplace leak, and agricultural fields [10,13]. Various NH3 sensing approaches along with advantages, disadvantages, and potential application are discussed in Table 1.
Chemiresistor is a sensor that consists of a sensing material in the form of a layer deposited over a substrate containing electrodes followed by detecting circuitry [10]. It detects alterations in the conductance of sensing material in the presence of NH3. The commercialization of NH3 chemiresistors for personalized health, industrial, environmental, and agricultural monitoring is proliferating. NH3 monitoring remains an area of active research owing to urgent requirements for reliable, efficient, miniature, portable, and intelligent sensing systems for security and healthcare. The current state-of-the-art in gas sensors for NH3 monitoring has already been discussed in literature [2,3,4,5,6,7,8,9,10].
Currently, 2D sensing nanomaterials owing to their larger specific surface area, tunable chemistries, and their fabrication techniques, have been extensively studied in terms of three S’s (sensitivity, selectivity, and stability) and five essential R’s (room temperature operation, range of detection, repeatability, response and recovery, and reproducibility) [11,14,15,16,17,18,19,20,21]. A 2D material (such as graphene and its derivatives, 2D organic frameworks, and polymers, black phosphorus, borophene, metal dichalcogenides, phosphorene, molybdenum disulfide, metal carbides, or nitrides (MXenes)) is generally stated to an atomically thin layer or few layers of flakes with thickness varying from 1–5 nm [14,21]. Although various conventional 2D materials, including graphene or metal dichalcogenides, have shown promising NH3 sensing properties due to their high specific surface area and tunable chemistries, they are lacking due to complex synthesis techniques, difficult surface functionalization, and poor selectivity [18,19,20]. MXenes are one of the recent and most popular NH3 sensing materials owing to their excellent hydrophilicity, excellent conductivity, large surface to volume ratio, biocompatibility, high mechanical strength, and enriched surface chemistry (due to the presence of surface terminated groups such as –OH, –O or –F) [22,23,24]. Due to excellent metallic type conductivity, MXenes are promising to detect reducing NH3 with a higher probability of its oxidation [25]. However, due to high adsorption energy toward NH3, they lack complete and rapid recovery [25]. MXene-based NH3 sensors must be produced for real-world applications in large quantities with high mechanical flexibility for machine processability. Thus, the conductivity and flexibility must be optimized simultaneously, which is still challenging [26,27]. In spite of remarkable features, the use of MXene based sensors for practical applications is somehow limited to the best of our literature knowledge.
Conversely, chemiresistors based on nanostructured polymers (e.g., polypyrrole (PPy), polyvinyl alcohol (PVA), polyaniline (PAN), poly (aniline co-pyrrole) (PAP), polystyrene (PS), poly 3,4-ethylene dioxythiophene (PEDOT), and cationic polyacrylamide (CPAM)) of desired functionality and conductivity continue to enhance NH3 sensing performance [10,28,29,30,31]. Polymer-based NH3 sensors are popular for miniaturized configuration, user, and environmentally friendly, tunable chemistries, low-cost, flexibility, and energy efficiency. However, owing to the high affinity of polymers toward humidity and volatile organic compounds present in the environment, it exhibits poor stability and cross-sensitivity [10,28,29,30].
Important issues related to cross-sensitivity and poor stability of polymers, and the poor recovery and difficult processability of MXene, etc., hinder their versatile utilizations in NH3 sensor fabrication. The use of hybrid nanocomposites of these two materials may result in NH3 sensors with efficient and enhanced NH3 sensing characteristics. Hybrid nanocomposites are a multi-component material in which at least one of the parent components are of nanoscale dimensions and can be classified as intercalated systems, phase-separated systems, and exfoliated systems [32,33]. It has been predicted that host−guest chemistry and the use of polymer and MXene counterparts in hybrid nanocomposites collectively help to eliminate their shortcomings owing to complementary or synergetic effects, leading to the development of improved NH3 sensors [27,34]. MXene–polymer nanocomposites have recently aroused inordinate interest in NH3 sensing applications. However, there are reports on nanocomposites of polymers with other 2D materials such as graphene and its derivatives, metal dichalcogenides and molybdenum disulfide [11,14,17,18,19,20]. However, the fabrication and functionalization of such nanocomposites is tedious, which limits their application in ammonia monitoring [11,14,17,18,19,20]. As well, the presence of surface functional groups over MXene makes them more prone to polymers for formation of enhanced hybrids through various multi-interactions such as hydrogen bonding, van der Waals forces, covalent bonding or electrostatic interactions [35,36].
Additionally, MXene–polymer nanocomposites (MXP-NCs) have already been explored for outstanding energy storage applications [35,36,37]. As well, various theoretical studies based on density functional theory predict MXP-NCs to possess better affinity toward ammonia [38,39,40]. This proposes MXP-NCs as a promising sensing material for ammonia monitoring as compared to other 2D materials –polymer nanocomposites. There is significant literature dedicated to MXene for gas detection applications; however, the reports on MXene–polymer nanocomposite-based chemiresistors are scarce [25,27]. To this date, there is no dedicated systematic review on MXene–polymer nanocomposites-based NH3 chemiresistors in the literature. With this motivation, this review focuses on MXene–polymer nanocomposites-based NH3 sensors and acts as a fundamental work structure to lead future research. This review emphasizes various prospects related to MXP-NCs types, their fabrication and structures, monitoring of NH3 in the environment, agricultural fields, and biomedical applications. Additionally, efforts have been made to discuss the NH3 sensing mechanism of MXene–polymer nanocomposites. However, better optimization is required for the commercialization of nanocomposite-based NH3 sensors.

2. Exploring MXP-NCs for Efficient NH3 Sensing

The MXP-NCs are rapidly growing in the research of advanced functional materials (Figure 3). MXP-NCs are combinations of nanoscale MXene and polymer counterparts, wherein multi-interactions at the molecular level generate exceptional characteristics at the interface.
Since 2014, the research group of Gogotsi has been exploring the origin, fundamentals, developments, prospects, and potential applications of MXP-NCs [23]. This group has explored MXene and MXene–polymer hybrid composites and anticipated this area of research as a multidisciplinary field. Their first report was on the fabrication of a hybrid nanocomposite of Ti3C2Tz, the most widely studied MXene to date, with PVA and PDDA [37]. Although there is extensive research dedicated to MXP-NCs after this first report, the scale is smaller than the studies on MXene itself [38]. Various polymers, including hydrophilic (such as PVA, PS, PA, PI, silicones) and conducting (PAN, PVP, PPy, PEDOT, PSS), have been used for making MXP hybrids [27,38,39,40]. Efforts have been dedicated to studying, optimization and improving of properties generated at the interfaces through molecular and supramolecular dynamics of MXP hybrid nanocomposites [27,38,39,40]. Experimental studies have revealed that incorporation of polymers in the MXene matrix or vice-versa leads to MXP-NCs with enhanced and improved properties [27,38,39,40]. The MXP-NCs offer diverse characteristics absent in their precursors, such as excellent conductivity and flexible mechanical properties due to interfacial multi-interactions among the two phases. Therefore, MXP-NCs have been predicted for various applications such as energy generation and storage, electromagnetic shielding, sensors, actuators, and optical limiters.
A promising and demanding field of research to investigate the applications is based on utilizing MXP-NCs as gas-sensing materials. Many reports based on density functional theory (DFT) have predicted the specific affinity of MXP-NCs toward NH3, which has led to dedicated research toward MXP-NCs-based NH3 sensors [38,39,40]. These MXP-NCs are suspected to demonstrate exceptional hetero-interfacial effects, enhancing interfacial charge transfer and contact with NH3 molecules due to abundant surface functionalities and synergistic effects. Other than interfacial considerations, the effect of various NC parameters such as NC composition and fabrication technique is still a matter of dedicated interest to achieve sensitive and selective detection. Hence, in developing MXP-NCs, a long-term goal is to advance new fabrication strategies to control morphology and composition. This control allows the optimization of MXP-NCs properties (including electrical, mechanical, and optical), which secondarily impacts their NH3 sensing performances. The classification of MXP-NCs based on their composition and adopted techniques to synthesize MXP-NCs for gas sensing applications is discussed below.

3. Classification of MXP-NCs Functional Structures

As a new family to 2D inorganic materials, MXene is at the infancy stage, obtained from selective elimination of “A” from its precursor MAX phase [23]. The MAX phases are hexagonal layered early transition metal carbides and nitrides with a composition represented by Mn+1AXn (n = 1, 2 or 3), where “M” is an early transition material, A is 13 or 14 group element, and “X” can be carbon, nitrogen, or both [23]. It has been revealed that there are around 70 MAX phases with ordered double early transition metal structures and more than 30 MXenes. Generally, MXenes are represented by the universal formula Mn+1XnTx with (n + 1) layers of “M” are interleaved with “n” layers of “X”. Various surface functionalities (including hydroxyl, oxygen, fluorine, and chlorine) are represented by “T” in the formula, resulting from different fabrication approaches [23]. Among all, M3 × 2Tx and M2X2Tx have been evaluated for NH3 sensing performance. As revealed by Lee et al. [41], titanium carbide (Ti3C2Tx) was the first MXene (M3X2Tx) reported for NH3 sensing, followed by specific reports [42]. Further, it has been revealed that M2X2Tx-MXenes (such as niobium carbide: Nb2CTx) possess higher specific surface areas owing to their fewer atomic layers compared to M3X2Tx-MXenes [43] and are a potential candidate for NH3 sensing. However, the practical applications of pristine MXene sensors are inhibited due to few drawbacks such as easy restacking, low flexibility, and poor stability in an oxygen atmosphere [44,45,46].
Taking advantage of easy processability, low cost, mechanical flexibility, and low toxicity of macromolecules, the combination of MXenes and polymers can further enhance the NH3 sensing performance of their nanocomposites, including mechanical flexibility, electrical conductivity, and solution solubility. Depending upon the nature and structure of precursor MXene (variation in ‘n’), MXP-NCs can be generally represented by formula Mn+1XnTx-P with “P” representing precursor polymer or polymer combination. To date, M3X2Tx-P and M2X2Tx-P classes of MXP-NCs have only been studied for NH3 sensing due to ease of synthesis and stability compared to other MXenes [25,27].
As revealed, polymer addition leads to exfoliation of MXene due to a surge in the interlayer distance. It inhibits the stacking of MXene layers and leads to the formation of hierarchal lamellar structures. Depending upon the nature of the polymer, interfacial multi-interactions (Figure 4) including covalent interaction (poly(2-(dimethylamino)ethyl methacrylate: (PDMAEMA) [47], electrostatic interaction (PAN, PEDOT) [48,49] and hydrogen bonds (PPY, PAM, PSS) [50,51] results in the formation of MXP-NCs with MXene layers stacked by polymer chains. Li et al. [52] reported forming a hierarchy of core-shell-type PAN wrapped Ti3C2Tx nanosheets accompanied by PAN dendrites. The dendritic structure over PAN nanoparticles increases the interlayer distance in lamellar structure, enlarging its specific surface area. Hence, the state-of-the-art fabrication of MXP-NCs is dedicated to developing new fabrication techniques to optimize morphology and properties. The adopted approaches to design Mn+1XnTx-P NCs with desired properties are discussed below.

4. Fabrication of MXP-NCs-Based High-Performance NH3 Sensors

The exceptional properties of MXP-NCs for NH3 sensing result from constant advancements in its fabrication technique. These developments are planned to efficiently optimize the required properties of target sensing material such as morphology, topology, crystallinity, and electrical, mechanical, and thermal properties. Several ways to develop state-of-the-art MXP-NCs based NH3 sensors generally consist of three stage procedures: synthesis of MXene, synthesis of MXP-NCs, and fabrication of chemiresistor.

4.1. Stage-1: Synthesis of MXene

The synthesis of MXene is the primary stage to develop a state-of-the-art MXP-NCs based NH3 sensing device [38,39]. Recently, many techniques to synthesize MXene have been developed, which can be briefed into three main strategies: selective etching techniques (Wet etching from precursors, molten-salt etching), bottom-up techniques (chemical vapor deposition: CVD, salt templated growth) and chemical transformations (ammoniation, deoxygenation and carburization). Numerous advancements in these techniques are already extensively reviewed [23,42].
A typical top-down etching-based approach consists of three basic steps: etching, washing, and delamination [23,53] (Figure 5). Etching involves removal of “A” element from MAX phase through the action of etchants such as hydrofluoric acid (HF), lithium fluoride (LiF), hydrochloric acid (HCl), or their combinations. It is followed by washing through centrifugation, decantation, or dispersing into freshwater, in which the pH of a remnant is adjusted to neutral. Lastly, single-layer MXene or few layers are obtained through manual shaking, ultrasonication, or by adding intercalants under delamination process [23,42]. The chemical mechanism associated with the synthesis of MXene using preferential HF etching process can be elaborately demonstrated as follows [23,39,54,55]:
Mn+1AXn + 3HF → AF3 + 3/2 H2 + Mn+1Xn
Mn+1Xn + 2H2O → Mn+1Xn(OH)2 + H2
Mn+1Xn + 2HF → Mn+1XnF2 + H2
In a typical etching based synthesis, Wang et al. [56,57] used HF etching followed by tetrapropylammonium hydroxide (TPOH) intercalation process to achieve ultrathin 2D niobium carbide MXene nanosheets (Nb2CTx) from niobium aluminum carbide MAX phase (Nb2AlC). A similar procedure was adopted by Li et al. [52] in the synthesis of 2D titanium carbide (Ti3C2Tx) nanosheets from titanium aluminum carbide (Ti3AlC2) using HF etching and TPOH intercalation.
There are other reports on HF-free synthesis of MXenes by substituting it with ammonium fluoride (NH4F) aqueous etching method to avoid the toxicity and corrosive nature of HF [23,58]. Xie et al. [59] substituted HF with HCl to fabricate Ti3C2Tx, but the obtained MXene was not found to be ideal. Bottom-up approaches include chemical vapor deposition (CVD) and high-temperature molten salt method, but those are complex, time-consuming, and costly techniques [23,39,60]. As well, there are no reports on the synthesis of MXene using a bottom-up approach to develop state-of-the-art MXP-NCs based NH3 sensors.

4.2. Stage-2: Synthesis of MXP-NCs

This stage is composed of synthesizing MXP-NCs using various strategies. The strategies reported for the synthesis of MXP-NCs can be summarized into two major approaches, including ex situ and in situ routes [38,39,40] (Figure 6). Ex situ routes include blending of separately synthesized precursor materials (MXenes and polymers). In contrast, in situ routes involve either the synthesis of both precursors and synthesizing one precursor in the presence of others. However, other reports on novel routes include RIR-MAPLE deposition, roller mill assisted synthesis and emulsion-based synthesis for stage-2 [27,39] NH3 sensor chemiresistor fabrication. Hence, this review will majorly emphasize in situ and ex situ routes.

4.2.1. Ex situ Routes

The ex situ synthesis routes are more commonly applied since it is easy to control the composition and properties of MXP-NCs. Taking advantage of weak interaction (electrostatic interaction, van der Waals attraction, and hydrogen bonding force) between polymer and MXene precursors, MXP-NCs are mainly synthesized by ex situ approaches. In ex situ routes, presynthesized precursors (MXene and polymer) are mixed through various techniques such as blending (solvent processing, wet spinning methods), alternate deposition, and melt processing to design MXP-NCs [27,38,39].
Majorly, blending techniques are applied for hydrophilic precursors, in which presynthesized precursors are blended in a common solvent [25,38,39]. In a typical synthesis, aqueous or colloidal MXene is added to the polymer solution. The solvent is removed by a suitable technique such as vacuum filtration, evaporation, and precipitation [39]. During wet spinning, an aqueous solution of precursors is blended through centrifugation to obtain MXP-NCs. Zhao et al. [61] fabricated Ti3C2Tx/CPAM NC by stirring the aqueous solution of Ti3C2Tx and CPAM through centrifugation. However, in solution blending, two different precursor solutions (which are mutually soluble) are used [39]. Naguib et al. [62] reported Ti3C2Tx/PAM NC fabrication by blending two distinct dispersions of MXene in dimethylsulfoxide (DMSO) and aqueous PAM. When both the precursors are blended, the polymer chains push the MXene sheets to the interstitial space among the chains, leading to the formation of a 3D continuous hierarchal network of conductive sheets. This continuous linked conductive network and lamellar structure are essential to design a state-of-the-art sensor. Hence, blending routes are advantageous due to the exclusion of toxic solvents, ease of processing, and ability to modulate the physical and chemical properties by controlling the concentration of precursors [38,39]. However, these approaches are limited due to the nature of solvent and polymer. Polar solvents (including protic and aprotic) can only be used in blending approaches due to their intrinsic compatibility with MXenes [38].
Similarly, only soluble or solution-dispersible polymers can be processed using these routes. In contrast, nonpolar or weekly polar polymers require surface modifications to increase solution dispersibility before solvent blending. Yu et al. [63] reported the use of cetyltrimethylammonium bromide (CTAB) to increase the dispersion of Ti3C2Tx in thermoplastic polyurethane (TPU). Recently, Wang et al. [56] reported an alternate layer deposition technique for Nb2CTx/PAN-based NC-based NH3 sensor fabrication. The Nb2CTx/PAN NC was obtained by sequential deposition of the PAN layer over polyimide surface through in situ polymerization and Nb2CTx layer using spray coating. Since neither of the precursors was deposited in the presence of others, the approach belongs to the ex situ route. The route is advantageous to control the thickness and composition of precursors in NC and possess the ease of processing. However, it lacks in the formation of intrinsic hetero-interfacial junctions among NC.
MXene is incorporated into the polymer matrix in melt processing through thermal techniques, including hot pressing, extrusion, and injection molding conducted above the melting point of precursor polymer [38,39]. Sheng et al. [64] reported the fabrication of Ti3C2Tx/polyethylene glycol (PEG) NC using the melt-processing technique. This technique is advantageous for the large-scale processing of MXP-NCs. However, this technique can only be used for polymers that can be melt-processed [38]. In addition, high-temperature processing may degrade precursors, which impacts the intrinsic properties of fabricated MXP-NC [39].
Though ex situ routes are majorly used for the fabrication of MXP-NCs, their commercial applications are limited due to various drawbacks such as struggle in evaporation of solvent, poor flatness, lack of the intrinsic linkage between precursors, substantial waste, less mechanical strength, non-uniform inclusion of precursors, and environmental contamination.

4.2.2. In situ Routes

In in situ approaches, both precursors are either synthesized together or one in the presence of another [38]. It has been reported that for the fabrication of MXP-NCs, a specific monomer has been polymerized on presynthesized MXene using in situ polymerization [27]. These approaches can be summarized into two broad categories: solvent-assisted methods (spin coating, dip coating, drop-casting, inkjet printing, spray coating) and heat-assisted ring-opening polymerization [39]. For solvent-based processing, the desired amount of monomer and MXene is mixed in an appropriate solvent and further added with a suitable amount of polymerization agents, including dopant, oxidant, and surfactant, for a definite amount of time. Typically, the synthesis of NC is confirmed by markers indicating completion of polymerization, such as a change in color (pale yellow or green for PAN). There are reports on the fabrication of PAN/MXene by chemically polymerizing aniline via oxidant APS (ammonium persulfate) and HCl (dopant) over Ti3C2Tx nanosheets. Jin et al. [65] reported the fabrication of PEDOT: PSS/Ti3C2Tx NC through in situ polymerization of EDOT on Ti3C2Tx sheets in the presence of PSS. Furthermore, Qin et al. [66] utilized in situ electrodeposition to obtain MXene/PPy nanocomposite with a 3D porous structure. Carey et al. [67] obtained polyamide (PA)-6-based MXene NC utilizing thermally assisted ring-opening polymerization. Typically, a mixture of monomer and MXene is heated at elevated temperatures (above the melting point of polymer) under ambient conditions.
The authors have further demonstrated in their respective reports that in situ routes improve dispersity, electrical, optical, and chemical properties [25,38,39]. The advantage of the in situ route is to obtain well-dispersed precursors, which results in the formation of hetero-interfacial junctions within the NC structure. The unique advantage of this approach is the formation of polymer chains on the MXene surface, which leads to exfoliation and expansion of layers of MXene-layered structures [36,38,39]. It helps in achieving a larger specific surface area with more significant porosity. The formation of hetero-interfacial junctions and increase in specific surface area is essential to design a state-of-the-art sensor. However, only a small amount of monomers can be polymerized over the MXene surface due to insufficient polymerization energy. As revealed by Chen et al. [39], there are two primary types of electron transfer during charge transfer induced polymerization of EDOT, including either shift of electrons from highest occupied molecular orbitals (HOMO) of monomer to lowest unoccupied molecular orbitals (LUMO) of MXene or HOMO of MXene to LUMO of monomers. In addition, the monomer must be close to the MXene layer to complete charge transfer during polymerization. Hence, both the processing routes own merits and demerits in designing state-of-the-art sensors, and advances in them are currently researched and developed.

4.3. Stage-3: Fabrication of Chemiresistor

Chemiresistors are processed in the final stage by fabricating a sensing film over a suitable substrate utilizing various techniques (Figure 7) such as drop-casting, electrospinning, dip coating, and inkjet printing [17,25,36,45,68]. Various flexible and nonflexible substrates used in the processing of chemiresistors include a glass substrate, fluorine-doped tin oxide (FTO) substrate, indium-tin-oxide (ITO) substrate, polyindole (PI), and polyethylene terephthalate (PET) [10,11,27,42]. Furthermore, suitable conducting electrode configurations such as parallel and interdigitated configured electrodes are deposited on the sensing layer/substrate utilizing suitable techniques such as thermal evaporation, inkjet printing, mask coating, and dip coating [11]. Typically, electrodes of excellent conductors (such as gold (Au)) are used to avoid the presence of noise due to the Schottky effect [10,17].
There are reports on using a dip-coating approach to obtain MXP-NCs over PI substrates with Au interdigitated electrodes [52,65]. Wang et al. [56,57] fabricated a Nb2CTx/PAN chemiresistor by spray coating prepared MXP-NCs over PI substrate (with Au interdigitated electrodes) in two different reports. However, Zhao et al. [61] used PET substrate with interdigitated silver electrodes to fabricate the CPAM/Ti3C2Tx based NH3 sensor. Few reports on the fabrication of NH3 sensors suggest dispersing polymeric systems such as PAP or its composites such as polyaniline–silver nanocomposite in m-cresol on glass substrates [3,69]. However, those are not flexible and are not reported for MXP-NCs to date. Hence, the processing of state-of-the-art NH3 sensors is accomplished in these three stages, and the electrical, molecular, mechanical, structural, and thermal properties of MXP-NCs can be optimized using appropriate synthesis routes and varying operational parameters or precursors.

5. Unique Properties of MXP-NCs for NH3 Sensing

Generally, the key to any application lies in tuning the physical and chemical properties of materials used. For instance, controlling conductivity and morphology of sensing material is a crucial parameter in state-of-the-art sensors. It has been revealed that materials with superior conductivity are utilized to detect reducing analytes such as NH3 [11,42,69] and that lower conductivities are suitable for monitoring oxidizing analytes such as sulfur dioxide [70,71,72]. Hence, the desired properties of sensing materials are optimized during the synthesis process by varying operational parameters. Mostly, the critical advantage of NCs is the specificity of their properties since specific materials can be utilized to detect specific analytes selectively. MXP-NCs have been proven to be excellent sensing materials with improved properties such as superior conductivity, porous morphology, improved thermal property, enhanced mechanical strength, and flexibility. In addition, oxidation of MXenes is significantly alleviated due to the casing of MXene through polymer [33,38,40]. Advancements in various properties of MXP-NCs prone to improved NH3 sensing are discussed below.

5.1. Morphological Properties and Molecular Interactions

Morphology and molecular properties are the most critical parameters to determining the sensing performance of MXP-NCs NH3 sensors. The specific surface area, porosity, and roughness of sensing materials are essential for sensing performance since sensing is a surface phenomenon. It has been revealed that the inclusion of polymer exfoliates MXene layers, resulting in higher specific areas and improved porosity [39]. The unique morphologies of NCs with optimum porosity have rendered them attractive as the NH3 sensing material.
Li et al. [52] reported the hierarchy of core-shell type PAN wrapped Ti3C2Tx nanosheets accompanied by PAN dendrites in Ti3C2Tx/PAN NC (Figure 8). It was revealed that the presence of dendrites on PAN nanoparticles has significantly increased the interlayer distance between MXenes, which enlarges its specific surface area and porosity. Jin et al. [65] observed significantly enhanced interlayer separation between Ti3C2Tx nanosheets by introducing PEDOT: PSS. Zhao et al. [61] observed the formation of a multilayer Ti3C2Tx structure with different layers glued by CPAM. Similar results were obtained for M2X2Tx-P NCs by Wang et al. [56], utilizing in situ and ex situ synthesis approaches in distinct reports. As a product of the ex situ route, the formation of porous PAN nanofibers coated by ultrathin Nb2CTx nanosheets was observed. In addition, these conductive NC fibers were reported to form heterogeneous interfaces owing to multi-interactions among precursors.
In contrast, 3D porous NC film formation due to the growth of PAN nanofibers over 2D Nb2CTx nanosheets was fabricated by an in situ route [57]. The formation of 3D spatial structured NC was ascribed to the multi-interaction phenomenon (hydrogen bonding, electrostatic interactions, and van der Waals forces) due to surface functionalities. Numerous irregular holes over the NC surface were reported to significantly facilitate adsorption and desorption of NH3 during sensing phenomena.
Generally, surface functionalities and molecular interactions are explored experimentally by various spectroscopic studies. Wang et al. [57] revealed hydrogen bonding among PAN-Nb2CTx NC precursors through the blue shift in Fourier transform infrared spectra of pristine PAN. Zhao et al. [56] revealed the presence of surface defects through Raman spectra with an increase in I(D)/I(G) (the D-peak to G-peak intensity ratio) for Ti3C2Tx/CPAM NC compared to that of pristine Ti3C2Tx. Similar observations revealing the presence of hydrogen bonding and electrostatic interactions between precursors were observed by Jin et al. [65] in terms of enhanced I(D)/I(G) with a relative shift in D-band and G-band. These interactions and specified surface functionalities are observed to promote NH3 adsorption over that of other analyte molecules.
The presence of surface defects with enhanced surface functionalities and multi-interactions significantly promotes the adsorption of analyte molecules. The advancements in NH3 sensing materials are concerned with attaining lamellar porous structures with enlarged interlayer distances with significant molecular interactions/defects. These hierarchal layered structures linked by macromolecules promote rapid and extensive adsorption of NH3 molecules due to higher specific surface area, large porosity, and enhanced surface functionalities. The current research is dedicated to designing different architectures of MXP-NCs to facilitate molecular interactions.

5.2. Electrical Properties of MXP-NCs

Generally, MXenes possess excellent conductivity (around 9880 S/cm) [23,58], and their incorporation in polymer improves the electrical properties of NCs such as electrical conductivity, ionic conductivity, charge transport, and conducting pathways [39]. In addition, the intercalation of polymer into MXenes averts restacking and endorses molecular-level coupling among them. The electrical properties of NCs are generally optimized by tuning the composition of precursors [36,39]. Jin et al. [65] systematically investigated the variation in conductivity of pristine PEDOT: PSS (~0.01 S/cm) with an increase in Ti3C2Tx weight percentage (wt%) in NC. The conductivity first increased (~0.02 S/cm for 8%, 0.02 S/cm for 8%, 0.07 S/cm for 15%) and then decreased (~0.03 for 20% and ~0.02 for 25%) as the addition of Ti3C2Tx. The highest conductivity was observed with 15 wt% Ti3C2Tx. It can be attributed to the balance between the confinement and percolation threshold of Ti3C2Tx, where low Ti3C2Tx addition accelerates charge transport, while the high Ti3C2Tx addition retards it to rigid confinement network. Similar results were obtained from PAN/Nb2CTx NCs by Wang et al. [56] by varying the Nb2CTx spray volume (0.05 mL to 0.2 mL). They revealed that the larger concentration of Nb2CTx (0.2 mL) is not conductive and not suitable for NH3 sensing. Hence, the movement of charge carriers is restricted after the percolation threshold of NCs and reveals the importance of optimizing the composition of precursors in NCs.
Li et al. [52] observed the significantly enhanced degree of protonation in PAN due to the addition of Ti3C2Tx, which facilitates electron transfer in fabricated NC, resulting in its enhanced sensing characteristics. As revealed from X-ray photoelectron spectra (XPS) (Figure 8), the degree of protonation for PAN/Ti3C2Tx corresponding to the percentage of =NH+- and -NH2+- (58.2%) was higher than that of pristine PAN (41.4%). Similar observations were made for Ti3C2Tx/perfluorosulfonic acid (Nafion) composite membrane by Liu et al. [73]. Hence, incorporating MXene into polymer facilitates charge transport by providing better conductive pathways [39,74]. The enhanced charge carrier transport facilitates the potential applications of MXP-NCs in the field of state-of-the-art NH3 sensors.

5.3. Thermal Properties of MXP-NCs

Generally, thermal properties such as thermal conductivity, thermal expansion, and thermal stability are closely related to material processing and practical applications. The thermal stability of sensing material is the main parameter for its commercial development. According to the first principle DFT calculations, the predicted thermal stability of MXenes is higher than those of most conducting and semiconducting low dimensional materials, indicating their potential to increase the stability of MXP-NCs based sensing devices.
Jin et al. [65] investigated the thermal stability of PEDOT: PSS/Ti3C2Tx NCs by thermogravimetric analysis (TGA). It is revealed that the rate of weight loss with temperature (between room temperature to 300 °C) for NC is slower compared to that of pristine PEDOT: PSS (Figure 9). The enhanced thermal stability of MXP-NC is attributed to multi-electron interactions among its precursors.
Furthermore, many studies [75,76,77] revealed the improvement of thermal conductivity of NC due to the addition of MXenes. It is ascribed to the excellent thermal conductivities of MXenes, which improve the thermal stability of NCs. The increase in thermal conductivity of MXP-NC results in a higher heat dissipation capacity of NC, improving its thermal stability [39]. In addition, it was observed that the incorporation of MXene reduces the random arrangement of a polymer chain, which deteriorates the scattering of phonons [76]. Hence, the enhanced thermal stability of MXP-NCs contributes to extend its sensing application under different temperatures, particularly in harsh environments. However, a direct role of thermal conductivity during NH3-sensing phenomena in MXP-NC has not been reported yet.

5.4. Other Advanced Properties

In addition, other properties of MXP-NCs have been revealed to be relevant to determining its NH3 sensing performances. Mechanical flexibility and mechanical strength are highly desirable for the commercial development of NH3 sensor [11,13,78]. MXP-NCs are an example of engineered material with mechanical properties, which possess significant mechanical flexibility due to polymer and enhanced mechanical strength in terms of tensile strength, Young’s modulus, and less crack formation due to the presence of MXene [38]. Zhao et al. [61] reported Ti3C2Tx/CPAM NC flexibility for NH3 sensing with a bending test. It is revealed that the NCs are strong and flexible due to the binding action of CPAM among Ti3C2Tx layers. It is ascribed to hydrogen bond forces between surface functionalities (amide group) of Ti3C2Tx surface and CPAM chains. However, Li et al. [52] and Wang et al. [57] reported using flexible PI substrates to achieve mechanical flexibility of designed MXP-NC-based NH3 sensors. The present reports provide promising strategies for the commercial fabrication of MXP-NCs based sensors with significantly reinforced mechanical properties. However, several unresolved tribological concerns remain due to the distribution of precursors, stress transfer mechanism, size-effects due to precursors, etc.
Recently, it has been revealed that the ordering of molecules/chains in polymer and its composites play a vital role in determining their sensing performances [79]. Generally, polymers are amorphous in nature [3,28,33]. The incorporation of MXene in polymer matrix results in the ordering of its polymer chains up to a certain extent, which turns MXP-NCs into semi-crystalline nature [52,56,61,65]. In addition, Li et al. [52] revealed the reduction in peak intensity and peak width at half-height (FWHM) for PAN peak in Ti3C2Tx/PAN NC, which may facilitate charge carrier transport. Chaudhary [79] revealed that the interchain spacing between polymer chains increases with a reduction in FWHM value, which facilitates intrachain hopping of charge carrier and restricts interchain hopping. In amorphous macromolecules, the direct current charge carrier transport is governed by Mott’s variable model, in which charge carriers hop from one site to another in variable range (in 3D) [70,79,80].
However, incorporating foreign material in NCs such as metal or metal oxide may restrict the hop of charge carriers to one dimension, which facilitates the charge carrier transport during sensing phenomena and enhances its sensing performance [79] (Figure 10). However, similar observations have not been reported for MXP-NC to date. Nevertheless, based on the same theory, it can be depicted that the charge carrier transport in polymer precursors of MXP-NCs may be enhanced due to prominent intrachain hopping (owing to reduced FWHM). Hence, MXP-NCs provide better charge transport pathways to conduct charge carriers constituting sensing signals during sensing phenomena. Many efforts have been dedicated to exploring the superior properties of MXP-NCs, which further extend their utility in NH3 sensing. However, the precise relationship between physiochemical properties and microstructure of MXP-NCs still requires to be thoroughly investigated, which is essential to guide the architecture of MXP-NCs with outstanding NH3 sensing performance.

6. Approaches for NH3 Gas Sensing

The NH3 sensing behavior of MXP-NC sensors is obtained by estimating time-dependent variation in the magnitude of conductance (in terms of resistance, current, or voltage) as a function of NH3 concentration at room temperature [53,56,57,61,65]. Typically, an NH3 sensing setup consists of a test chamber with provisions for controlling temperature and humidity along with NH3 flow (Figure 11). It has been revealed that the exposure to NH3 molecules onto an NC-based chemiresistor results in a variation in resistance, which is recorded by the use of a measuring device such as a digital multimeter (DMM) or source meter unit (SMU) interfaced to a computer through software such as LAB View [34,57,69]. Simultaneously, NH3 flow is cut off, and nitrogen/air is introduced into the test chamber, which restores the original resistance magnitude. NH3 sensing measurements are performed by two strategies, including either static mode or dynamic mode. A definite NH3 concentration is introduced into a test chamber during static mode measurements, and time-dependent change in resistance is recorded until saturation [65]. Jin et al. [65] used static mode to record NH3 sensing performance of Ti3C2Tx/PEDOT: PSS NC, in which NH3 is added to an air-tight test chamber through a microsyringe. However, in dynamic mode, a continuous flow of definite NH3 concentration is maintained by a mass flow controller (MFC), and the corresponding change in resistance is monitored [57]. Li et al. [52] maintained a constant flow of NH3 through the test chamber during dynamic measurement of the time-dependent variation of resistance in Ti3C2Tx/PAN NC. These two standard strategies are majorly reported to analyze NH3 performance in current state-of-the-art sensors.

7. NH3 Sensing Mechanism Supported by MXP-NCs

The resistance of MXP-NCs varies time-dependently as a function of NH3 concentration. The observed variation in resistance of MXP-NCs in NH3 is primarily ascribed to synergistic effect, charge transfer amid NC and NH3 molecules, and redox reactions among NC and NH3 [52,65]. It has been revealed that the electrical conductivity of MXenes varies from metallic to semiconducting regime depending upon the nature of “M”, “X” and surface functionalities [58,81]. In addition, the incorporation of polymers into MXenes results in the semiconducting behavior of MXP-NCs [38,39]. However, the shift in electrical conducting behavior of MXP-NCs depends on its precursors’ nature and composition. Hence, optimizing surface functionalities and precursor concentration to achieve desired conductivity are current challenges in state-of-the-art sensors.

7.1. Ammonia Sensing Mechanism in Pristine Precursors of MXP-NCs

To develop a better understanding of an ammonia-sensing mechanism in MXP-NCs, it is essential to understand the interaction of ammonia with pristine precursor materials. Hence, by unifying the interaction mechanism of ammonia with individual precursors, a theory for ammonia sensing through NC can be developed.
Ti3C2Tx and Nb2CTx are MXenes precursors, and p-type conducting polymers (such as PAN, PEDOT, etc.) are polymer precursors majorly utilized to design MXP-NCs-based NH3 sensors. Hence, the NH3 sensing mechanism exclusively depends on heterojunctions/interfaces between the precursors of MXP-NCs. However, understanding the NH3 sensing mechanism of precursors of NCs plays a crucial role in exploring that of NC. The p-type conducting polymers (such as PAN, PPy) undergo protonation/deprotonation on adsorption/desorption of NH3 molecules [3,82].
The interaction of p-polymer with NH3 results in the compensation of charge carriers by transfer of protons or electrons from polymer ring to NH3 molecule [83,84] and can be represented as adsorption–desorption phenomenon given by
P + X + NH 3 δ Adsorption Desorption P 0 + NH 3 + δ X & P + X + NH 3 Adsorption Desorption   [ P 1 H ] 0 + NH 4 + X
where   X may be any dopant species such as Cl , NO 3 , ClO 4 SO 4 . In such an interaction, the number density of charge carriers decreases, resulting in a decrease of conductivity and an increase of electrical resistance of the polymer.
For instance, a change in the redox state of PAN occurs from emeraldine salt to emeraldine based on exposure to NH3 molecules. Hence, on exposure to NH3, the resistance of p-type polymers increases due to a reduction in its majority charge carrier (hole) density [3,82]. Moreover, Kim et al. [44] revealed that the surface functionalization transforms metallic Ti3C2Tx into a semiconductor of a narrow bandgap. Li et al. [52] revealed that Ti3C2Tx demonstrates p-type sensing behavior toward reducing analytes such as NH3. In contrast, Wang et al. [57] revealed the n-type sensing behavior of Nb2CTx sheets toward reducing NH3. Hence, the NH3 sensing mechanism in MXP-NCs is primarily governed by the nature of heterojunctions/interface (p–p-type or p–n-type) between its precursors and secondarily depends upon chemisorption, hydrogen bonding, and formation of unique hetero-interface functions, which are common in both the types of MXP-NCs. These four significant phenomena together govern the NH3 sensing mechanism of MXP-NCs and are discussed below.

7.2. Ammonia Sensing Mechanism in MXP-NCs

7.2.1. Chemisorption Based Ammonia Sensing

In ambient conditions, the p-type MXP-NCs layer absorbs oxygen from surroundings and converts them into oxygen anions by trapping electrons from the conduction band of NC [85,86]. Hence, it forms an electron depletion layer of high resistance among p-type polymer and p-MXene. The interaction between oxygen anion on the NC surface and adsorbed NH3 molecules results in the release of electrons back to conduction band p-MXP-NC [87]. The space charge region varies with an increase of electrons in the conduction band of NC, which decreases its electrical resistance. The original resistance magnitude is restored by cutting off the NH3 supply and introducing fresh air due to the replacement of NH3 molecules by oxygen anions. Similar observations were confirmed for Ti3C2Tx/PEDOT: PSS NC by Jin et al. [65] (Figure 12), where they summarized entire chemisorption phenomena through chemical reactions, including:
O2 + e → O2
4NH3 + 5O2 → 4NO + 5e + 6H2O
However, this phenomenon is limited at room temperature due to the minute conversion of oxygen into oxygen anions [65]. Hence, the contribution of chemisorption phenomena to the net sensing mechanism is infinitesimal at room temperature.

7.2.2. Formation of Hydrogen Bonds

Various theoretical DFT studies have predicted the formation of a large number of intermolecular hydrogen bonds between precursors of MXP-NCs [38,39,57,87]. It has been revealed that the large number of hydrophilic surface functionalities of MXP-NC is occupied by these intermolecular hydrogen bonds, which decreases the water adsorption sites, and in turn, promotes adsorption of NH3 molecules [57]. Hence, the formation of hydrogen bonds on MXP-NCs surface contributes to the net-sensing mechanism by facilitating adsorption of NH3 molecules.

7.2.3. Formation of Unique Hetero-Interfacial Functional Groups

Various theoretical and experimental studies [36,39,43] have revealed the formation of unique hetero-interfacial functional groups among precursors (MXene and polymer), which transforms the electric behavior of MXP-NCs into semiconducting range. It has been revealed that the presence of unique hetero-interface functionalities facilitates interfacial charge transfer, which in turn enhances NH3 sensing performance of MXP-NCs. Similar observations have been reported for Nb2CTx/PAN and Nb2CTx/PAN-TENG-based NH3 sensors by Wang et al. [56,57] in two separate reports. Hence, hetero-interface functionalities between precursors contribute to the net NH3 sensing mechanism by enhancing the interfacial charge carrier transfer.

7.2.4. Physisorption-Based Gas Sensing

Physisorption is the predominant phenomenon contributing to the net NH3 sensing mechanism in MXP-NCs. The resistance of MXP-NCs either decreases or increases on exposure to NH3 molecules. It is ascribed to the dominance of either precursor over other during sensing phenomena. Jin et al. [65] revealed the decrease of resistance of Ti3C2Tx/PEDOT: PSS NC in the presence of NH3 due to increased charge carrier concentration. It is attributed to the enhancement of π–π interaction, specific adsorption area, and the number of majority charge carriers of PEDOT: PSS by addition of Ti3C2Tx in NC [65,88]. There is direct charge carrier transfer between NH3 molecules and NC surface, which leads to increased charge carriers, decreasing the resistance of NC [65]. However, the explanation is not supported by evidence from the experimental analysis.
In contrast, Li et al. [52] reported an increase in Ti3C2Tx/PAN NC resistance on interaction with NH3 molecules. It is ascribed to the predominance of PAN and the formation of Schottky junctions in NC [52]. It is revealed that the p-type PAN has a higher work function compared to that of p-type Ti3C2Tx, which results in the formation of Schottky junctions at their interfaces with hole depletion layer lying in PAN. The interaction of electron-donating NH3 molecules with the NC surface causes a decrease in the width of the depletion layer, which narrows the conductive pathways for charge carriers in PAN during sensing phenomena. Hence, the modulation in the space charge region of NC on interaction with NH3 molecules results in increased NC resistance (Figure 13). Similar observations were made for Ti3C2Tx/CPAM-based NH3 sensor by Zhao et al. [61]. Hence, in p–p-type semiconducting MXP-NCs, the sensing mechanism is governed by the predominance of either precursor or synergistic effects due to the formation of the interface between precursors.
However, in p–n-type MXP-NCs (Nb2CTx/PAN), Wang et al. [57] revealed the predominance of p-type PAN over n-type Nb2CTx during NH3 sensing (Figure 14). It was speculated that there is the formation of the p–n junction at the interface between the precursors, which facilitates NH3 sensing performance of NC. It results in the formation of a barrier layer between n-Nb2CTx and p–PAN. The interaction of NC surface with NH3 molecules causes a decrease in hole concentration (majority carriers) of PAN, which results in broadening the width of the depletion layer toward the PAN side. The narrowed conducting pathways due to broad depletion width at the PAN side increase the resistance of NC [57]. Hence, the net resistance of NC increases on interaction with NH3 molecules and restores its original value on flushing through air molecules [56]. A similar mechanism has already been revealed for various organic-inorganic NCs [20,34,46]. Hence, in MXP-NCs with PAN as a precursor, the predominance of PAN over MXene is observed, which can be ascribed to the unique redox nature of PAN and higher affinity toward NH3. However, the predominance of any precursor is also a function of the composition and type of NC. The following section focuses on the application of MXP-NCs to detect NH3 for environmental monitoring.

8. MXP-NCs-Based NH3 Sensing Performance

The NH3 sensing potential of MXP-NC sensors in terms of various sensing characteristics, including 3S’s, five essential R’s, flexibility, the effect of humidity and temperature, self-driven capability, and real-time monitoring, is discussed in this section. Advancement in sensing layer material and improvement in NH3 sensing characteristics are also critically discussed. Sensor fabrication using MXP-NCs for commercial prospects in agriculture, environmental monitoring, bio-medical fields, and workplace safety has also been highlighted.

8.1. Sensitivity, Room Temperature Operation, and Low Detection Limit

Sensitivity/sensing response is the crucial sensing characteristic to evaluate the NH3 sensing performance of a sensor. Generally, sensitivity toward NH3 is represented in terms of the degree of change/percentage change in magnitude of conductance (in terms of resistance or conductivity or current or voltage) on interaction with NH3 molecules [34,65,71] and described in terms of:
Degree change (S) of sensing parameter (R) is given by:
S = ΔR/Ro
where ΔR is the change in resistance of chemiresistor on exposure to NH3 (Ra−Ro) with Ro is the stable value of its sensing parameter in ambient conditions, and Ra is the value of sensing parameter in the presence of NH3.
Percentage change in sensing parameter: S(%) = (ΔR/Ro) × 100.
The magnitude of sensitivity majorly depends on NC properties such as morphology, topology, dimension, the surface to volume ratio, effective surface area, porosity, chemical properties, defects, functionalization, composition, bandgap, conductivity, charge transport, charge carrier pathway of sensing material [10,11,18,20,70,89]. Hence, the research is dedicated to improving these properties in state-of-the-art sensors. As mentioned before, the presence of both the precursors (Polymer and MXene) in MXP-NCs contributes to its improved sensing performance, which predicts MXP-NCs as a potential candidate for NH3 monitoring. Various reports on MXP-NCs, including M3X2Tx-P and M2XTx-P types, have revealed the enhanced sensitivity of fabricated NCs compared to their precursors (Table 2). A comparative analysis of MXP-NCs, pristine precursors, and other prominent sensing materials are listed in Table 2.
The authors credited the improved sensing behavior of MXP-NCs compared to other materials owing to their higher specific surface area, large adsorption sites, unique surface functionalities, and optimum porosity.
The lowest detection limit (LDL) is defined as the lowest level of concentration of analyte, which can bring significant recordable change in sensing parameter of sensing material. It is essential to specify the application field of designed sensors. For instance, a sensor detecting NH3 at less than 0.8 ppm is suitable to act as a biomarker to detect renal failures/gastric ulcers. Among all reported MXP-NC NH3 sensors, Wang et al. [57] observed maximum NH3 sensitivity with the lowest detection limit (1.19% for 20 ppb NH3) Nb2CTx/PAN NC sensor, which was further evaluated for human breath analysis. They reported Nb2CTx/PAN NC sensor fabrication through the in situ route, which is suspected of forming hetero-interfacial junctions between its precursors [34]. Improved sensitivity is ascribed to 3D morphology, enhanced interlayer distance, and high specific surface area of NC, along with the presence of hetero-interfacial junctions. Wang et al. [56] measured NH3 (100 ppm) sensitivity of Nb2CTx/PAN NC (197.20%, 301.31%, 108.36%, and 48.90%) as function of Nb2CTx concentration (0.05, 0.1, 0.15, and 0.2 mL, respectively). It is revealed that the NH3 sensitivity first increased and then decreased with the addition of Nb2CTx. The highest was observed for 0.1 mL of Nb2CTx. It can be ascribed to the balance between the adsorption sites and Nb2CTx concentration, where after reaching the threshold, Nb2CTx hinders the NH3 adsorption sites, resulting in a decrease of sensitivity. After reaching the percolation threshold, the conductivity of MXP-NCs decreases, leading to a reduction in the probability of NC oxidation by NH3. Similar observations were made for PEDOT: PSS/Ti3C2Tx NC by Jin et al. [65]. The percolation threshold was observed at 15 wt.% of Ti3C2Tx with the highest conductivity of 0.07 S/cm and NH3 sensitivity (36.6%).
Further increasing the wt% of Ti3C2Tx in NC resulted in a decrease in its conductivity, decreasing its sensitivity. It is ascribed to optimal interlayer spacing at 15 wt% of Ti3C2Tx, which provides a higher specific surface area with large adsorption sites. Hence, the precursor concentration determines NH3 sensitivity and LDL of MXP-NCs by modulating its conductivity, interlayer distance, and the number of adsorption sites.
Room temperature operation is essential for commercial advancements of sensors since no micro-heating assemblages are required for their operation [10,11]. It reduces the cost and complexity of fabrication and energy for operation. Additionally, operating the sensor at room temperature prevents sensing material from degradation, increasing its working lifetime [72]. All MXP-NC sensors summarized in Table 2 work at room temperature owing to their enhanced properties, as previously discussed. Hence, MXP-NCs NH3 sensors are sensitive, cost-effective, energy-efficient with simple configuration, and possess a high life with a lower detection limit.

8.2. Response-Recovery Times, Repeatability, and Detection Range

Response/recovery times are defined as the time taken for sensing response (elevation curve/descent curve) to reach 90% of the saturated change in the sensing response when the sensor is subjected to a step-change in the original value of the sensing parameter [3,17]. They depend on the degree of adsorption/desorption of NH3 molecules over sensing material [71]. A suitable sensor is characterized in short response time, indicating rapid detection and short recovery time, indicating faster recovery to its original state. They are essential for repetitive utilization and confirm sensor long service life.
Jin et al. [65] revealed the dependence of response and recovery time of a Ti3C2Tx/PEDOT: PSS NC sensor on Ti3C2Tx concentration in terms of conductivity, adsorption sites, and specific surface area. A comparative analysis of reported response and recovery time for various NH3 chemiresistors is listed in Table 3. Among all reported MXP-NCs sensors, the lowest response and recovery time (12–14 s) were observed for Ti3C2Tx/CPAM by Zhao et al. [61]. It is ascribed to the choice of sensing parameter (current), which illustrates the importance of detecting signals in state-of-the-art sensors. Generally, MXP-NCs containing PAN as polymer precursors show slow response and recovery despite high sensitivity and the lowest detection limit. It can be ascribed to the affinity of PAN toward NH3, owing to its redox properties, which results in slower adsorption/desorption of NH3 molecules from the NC surface. However, Wang et al. [56,90,91,92,93,94] revealed that the coupling of triboelectric nanogenerators (TENG) with NC sensors results in improved response and recovery speed.
The repeatability of a sensor is characterized by recording the sensing behavior for many consecutive cycles [17]. It governs the stable sensing performance of a sensor without any influence on environmental conditions. A comparative analysis of repeatability results for various NH3 sensors is listed in Table 3. The reported MXP-NCs sensors are observed to be repeatable for many consecutive cycles, showing the stability of their sensing response. For instance, Zhao et al. [61] revealed the stability of the CPAM/Ti3C2Tx NC sensor for ten consecutive cycles of NH3 exposure.
Furthermore, a sensor is characterized in terms of linear detection range described by its linear fitting regression value (1 for perfect linear detection range) [11,17]. Linear detection range is a crucial parameter for commercial prospects of the sensor, which helps optimize and control sensing performance. Linearity for a sensor is determined by estimating the linear fit from a plot of sensing response as a function of NH3 concentration. Jin et al. [65] reported two linear ranges for a fabricated PEDOT: PSS/Ti3C2Tx sensor, including 10–100 ppm (with R2 = 0.957) and 100–100 ppm (with R2 = 0.983). Two linear regimes are ascribed to an increase in the interaction of NH3 molecules with sensor surface with an increase in NH3 concentration. The interaction between NH3 molecules and sensor is rapid at high concentrations and becomes moderate at low concentrations. Hence, two linear regimes are generally reported for NH3 sensors corresponding to low and high NH3 concentrations, respectively. However, the sensing response of highly porous sensing materials with large specific surface areas increases at a moderate rate, which results in a single linear regime throughout the detection range [45,56]. Hence, generally, M3 × 2Tx-P NH3 sensors consist of two linear detection ranges, and M2XTx-P NH3 sensors show only a single linear detection range. Consequently, improving response and recovery speed while designing NC with a porous architect is the current challenge in state-of-the-art NH3 sensors for their commercial prospects.

8.3. Selectivity Demonstrated by MXP-NCs-Based NH3 Sensors

Cross-sensitivity is the major challenge in the commercial development of gas sensors. A designed sensor/array of sensors is expected to detect specific gas/to distinguish among various gases for future intelligent devices such as the electric nose. The selectivity of the gas sensor is a crucial characteristic for its practical applications, which is observed by comparing its sensing response toward a particular analyte with other interfering analytes [11]. Alternatively, it is measured by recording the sensing response of the sensor toward a particular analyte in the presence of other interfering gases [10,17,95]. However, the selectivity of MXP-NCs NH3 sensors has been reported only through comparative analysis [52,57,61,65]. Various reports on MXP-NC sensors reported the selectivity test of designed sensors by exposing them to a specific concentration of different analytes such as ethanol, acetone, methanol, and sulfur dioxide. A comparative analysis of the selectivity test reported for various MXP-NCs is listed in Table 4.
Various MXP-NC sensors are highly selective toward NH3 compared to other interfering analytes (Table 4). It is ascribed to the adsorption energy of various analytes at different surfaces [3,28]. Li et al. [52] reported the sensitivity of Ti3C2Tx/PAN toward 10 ppm of NH3 to be at least one magnitude larger than that of other interfering gases. It is ascribed to unique characteristics of MXP-NCs such as surface functionalities, larger specific surface area, and exclusive affinity of precursors toward NH3. Yu et al. [96] revealed the specific selectivity of Ti2C-type MXene by theoretical DFT calculations and ascribed it to high adsorption energy of NH3 compared to other analytes. Lee et al. [41] described the enhanced selectivity due to functional groups and defects over the MXene surface, which facilitates its bonding and interaction with NH3. Wang et al. [56] also showed similar specific selectivity of a designed Nb2CTx/PAN sensor toward NH3. The fabricated sensor exhibits substantial sensitivity (around 9%) toward SO2, which is significant considering the literature [34,70,71,72]. However, the sensing signal for SO2 can be easily distinguished from that of NH3 since the sensing mechanism for both are different due to the opposite nature of gases (NH3: reducing, SO2: oxidizing) [3,70,71]. Hence, in addition to highly selective MXP-NCs NH3 sensor, these reports also open a prospect for detecting SO2 through MXP-NCs.

8.4. Mechanical Flexibility Demonstrated by MXP-NCs-Based NH3 Sensors

Mechanical flexibility is one of the essential characteristics of NH3 sensors for its commercial development, which makes them portable, foldable, bendable, and stable in every wear and tear situation [11,13,17]. Mechanically flexible sensors are easy to install at every emission site for monitoring NH3 emission/leakage. Generally, flexible sensors are built by utilizing two routes: either by designing self-standing flexible films or using flexible polymer substrates such as PI or PET [13]. The flexibility of NH3 sensors is tested by comparing its sensitivity under different bending angles and multiple folds to that of under unbent condition [11,13]. Zhao et al. [61] reported that incorporating CPAM into Ti3C2Tx results in mechanical flexibility of Ti3C2Tx/CPAM NC sensing film, which is missing in pristine Ti3C2Tx films. It is ascribed to the gluing action of CPAM, which bonds together Ti3C2Tx stacks, providing its mechanical flexibility. It is revealed that the sensitivity of a pristine Ti3C2Tx sensor varies under flexibility test. However, the sensitivity of the Ti3C2Tx/CPAM NC sensor is found stable with negligible descent. There is a slight increase in response on recovery time by approximately 2 s, but that is not significant compared to other sensors. Jin et al. [65] revealed the consistency of a PEDOT: PSS/Ti3C2Tx NCs NH3 sensor toward 100 ppm of NH3 under various bending angles (Figure 15). Various MXP-NCs-based NH3 sensors have shown stable sensitivity under the flexibility test, as illustrated in Table 5.
It has been observed that there are flexibility test reports on M3 × 2Tx-P NH3 sensors; however, M2XTx-P NH3 sensors have not been evaluated to date. Hence, MXP-NC NH3 sensors are easily processable to attain flexibility due to other polymer precursors or flexible substrates. However, the effects of type of substrate over NH3 sensing performance are still unexplored and possess vast scope to state-of-the-art sensors.

8.5. Stability-Based on the Effect of Varying Environmental Conditions

Stability is the most prominent factor in determining sensor performance in lifetime and consistency [11]. Stability is majorly affected due to degradation of sensing material on interaction with the analyte and variable surroundings [10]. Under ambient conditions, stability is measured by recording sensing response as a function of time (in days, weeks, or months) [34]. It is crucial to test the stability of MXP-NCs due to the possibility of degradation of its precursors. It has been revealed that polymers degrade with time due to interactions with surroundings, and MXene is prone to oxidation owing to its high reactivity [11,23,29]. However, stability is achieved in organic-inorganic NCs (including MXP-NCs) due to the formation of hetero-interfacial junctions or multi-interactions between its precursors [25,36,38]. Jin et al. [65] reported the stable sensing response (around 33%) of Ti3C2Tx/PEDOT: PSS sensor toward 100 ppm of NH3 for four weeks. Li et al. [52] and Wang et al. [57] also reported stability of a fabricated MXP-NC sensor for 35 consecutive days with stable sensing response toward 10 ppm of NH3. Hence, the MXP-NCs are reported stable in ambient conditions.
However, a sensor must also be stable for commercial purposes in varying temperature and humidity conditions, especially in harsh surroundings [11]. It has already been revealed that polymers (such as PAN) are prone to humidity due to bonding or swelling, which degrades their structure and sensing performance [10,97,98]. Abdulla et al. [99] revealed that the adsorption of water molecules over the PAN surface causes the release of electrons, which results in the formation of H3O+ ions and increases its ionic conductivity [100]. The interaction with NH3 molecules results in NH4OH or NH3∙H2O formation, which captures protons from PAN, resulting in increased sensitivity [99]. However, after a certain level of humidity, water molecules hinder the adsorption of the NH3 molecule limiting its sensitivity [52]. Consequently, there exists a dynamic equilibrium between humidity value and NH3 sensing response. Hence, the humidity test is performed by two routes: monitoring sensing response in the presence of NH3 [52] and monitoring change in sensing parameters in the absence of NH3 [3], with change in humidity.
Jin et al. [65] reported negligible sensing response toward humidity of Ti3C2Tx/PEDOT: PSS sensor. However, the NH3 sensitivity increases with relative humidity (RH: 20–90%). It is ascribed to increased charge transfer interactions due to a decrease in average separation between water molecules. However, Li et al. [52] revealed that the NH3 sensitivity first increases (0–40%RH) and then decreases (40–90%RH) with the increase in RH, and the threshold is found around 40%. The threshold is ascribed to the accumulation of a thin layer of water molecules on the sensor surface, hindering the NH3 adsorption sites. Li [52] and Jin et al. [65] used the same MXene precursor (Ti3C2Tx) to fabricate MXP-NCs; however, the sensitivity variation with RH is different. It is ascribed to the predominance of polymer precursor (PAN) in Ti3C2Tx/PAN NC reported by Li et al. [52], which possess a specific affinity toward NH3 [101].
In contrast, Wang et al. [56,57] revealed the threshold of RH for NH3 monitoring at 72.2% RH (Figure 16). The interaction of NH3 with water molecules present on the sensor surface results in an increase of NH3 sensitivity in the range of 41.0–72.2% RH; however, over 72.2% RH sensitivity decreases. It is ascribed to a higher specific surface area of Nb2CTx than Ti3C2Tx, shifting the threshold RH value to a higher magnitude. It reveals that the threshold RH value of the NH3 sensor depends on the nature and type of precursors (Polymer and MXenes). Hence, the influence of humidity on NH3 sensing performance is optimized by controlling the nature and composition of precursors in NCs.
Moreover, the study of variation of NH3 sensing response as a function of temperature is also vital to evaluate the performance of a sensor and record following the same strategy [11]. It is revealed that the increase in operational temperature results in a linear decrease of NH3 sensitivity of MXP-NCs [52,65,94]. It is ascribed to desorption of H3O+ ions from NC surface and the competition among desorption and exothermic adsorption of NH3 on sensor surface at higher temperature [102,103]. Additionally, the increase in the operational temperature of the sensor results in the formation of many oxygen radicals, which causes chemisorption dominance in contributing to the net sensing phenomenon. Similar trends are observed for MXP-NCs sensors by Li et al. [52] and Wang et al. [57], observing a linear decrease in NH3 with an increase in operational temperature.
Hence, MXP-NCs-based NH3 sensors operate with high stability in ambient and varying surroundings (humidity and varying temperature); however, the optimization of precursors is primarily required to record utmost sensitivity and stability.

9. Advancements in NH3 Detection for Applications Point-of-View

The inclusion of various technological developments is helping to extensively develop the application prospects of NH3 sensors with advanced sensing features. The current need for technological advancements has gained extensive interest in designing intelligent sensors, replacing the current complex, expensive and time-consuming techniques. For instance, monitoring NH3 in the agricultural field has advanced with intelligent sensors, reducing the requirement of human resources and efforts. This section highlights the advanced application of next-generation NH3 sensors using MXP-NCs as the sensing material.

9.1. Human Breath Analysis Based on NH3 Detection

The detection of urea/ammonium salt concentration is highly desirable for gastric ulcer/renal patients. It is executed through various complex techniques such as blood test, which are time-consuming, expensive, and requires dedicated human resources. Alternatively, the urea/ammonium salt concentration is detected by monitoring NH3 in exhaled human breath [104]. It has been revealed that the NH3 concentration in the breath of a renal patient is beyond 0.8 ppm, which can be potentially detected by an NH3 sensor with LDL less than 0.8 ppm [104,105]. MXP-NCs sensors are promising biomarkers in detecting NH3 in renal/gastric ulcer patients owing to their low detection limit (20 ppb: Table 1) with high sensitivity. Hence, airborne NH3 sensors have the potential to present an alternative to present-day monitoring techniques.
Wang et al. [57] revealed the use of an Nb2CTx/PAN sensor to detect NH3 in a human breath through an exhaled-breath analysis simulation system (Figure 17). The sensitivity of the Nb2CTx/PAN sensor is observed around 10% on exposure to the simulated breath of a patient containing 0.8 ppm of NH3. It is ascribed to the fascinating features of MXP-NCs toward NH3 detection such as higher specific surface area, large adsorption sites, specific surface functionalities, explicit affinity, and fast conducting pathways due to the presence of both the precursors. However, a study on other sensing characteristics for detailed breath NH3 analysis is missing and possesses scope for further advancements. The sensing analysis is required to associate with pattern recognition, array-sensing, and data from clinical trials for its commercial development in bio-medical applications.

9.2. Detecting Volatilization of Agricultural NH3

Reducing NH3 emission by employing advanced farm practices is a priority of environmental regulators in the agricultural field around the world. It has been revealed that the 90% of airborne NH3 is released from various agricultural sources such as livestock, fertilizers, etc. [106,107]. Hence, to sustain the implementation of NH3 emission mitigation practices, advanced sensors are developed to identify emission sources, individual contributions and assess the effectiveness of control measures. The MXP-NCs owing to their excellent NH3 detecting capabilities are promising to detect the volatilization of NH3 in the agricultural sector.
Li et al. [52] predicted the feasibility of a Ti3C2Tx/PAN sensor to detect volatilized NH3 in agricultural applications by designing an experiment based on an agricultural simulator. The NH3 volatilization results of the Ti3C2Tx/PAN sensor are compared with that of the conventional sulfuric acid adsorption method (SAAM) and Drager nitrogen tube method (DTM). Typically, a layer of fresh soil sprayed with aqueous urea solution is cultivated under dark and ambient conditions in a large vessel. A beaker containing sulfuric acid is placed in the vessel after fertilization, and the concentration of ammonium salt produced due to reaction among acid vapors and urea is recorded (SAAM). Additionally, a DTM tube is placed inside the vessel, which changes color on reacting with volatilized NH3 molecules. The color scale of the DTM tube gives a measurement of the concentration of volatilized NH3.
Moreover, a Ti3C2Tx/PAN sensor is also placed in the same vessel, and monitoring of volatilized NH3 is performed. The results obtained using the Ti3C2Tx/PAN sensor in predicting NH3 trends are similar to those obtained from conventional SAAM and DTM techniques [52] (Figure 18). The study proposes the potential of MXP-NCs sensor in detecting NH3 volatilization in agricultural sectors as an intelligent alternative for easy and precise detection with the reduced requirement of human resources. Hence, the MXP-NCs sensors can act as intelligent sensors that give rise to advancements in intelligence for agricultural applications. However, it requires dedicated research interest with advanced intelligent techniques to configure smart sensors with optimized performance.

9.3. Self-Driven NH3 Sensing

The energy crisis is another primary global concern, resulting in the advancement of self-drive smart sensors eliminating the requirement of conventional power supplies. Self-driven NH3 sensors operate through mechanical forces generated by nanogenerators such as TENG instead of conventional electric sources [94,108]. TENG operates on mechanical forces and acts as a voltage source for NH3 sensors [109,110]. Wang et al. [56] reported on a Nb2CTx/PAN–TENG sensor to detect a low trace of NH3. The fabricated Nb2CTx/PAN sensor (on the interdigitated PI substrate) is coupled to a TENG operated through a linear motor of a fixed frequency of 1 Hz (Figure 19A). The coupled TENG acts as a power source to drive the MXP-NC sensor for NH3 monitoring operation. The interaction between NH3 molecules and the sensor causes a change in real-time resistance of the NC sensor, which further varies the output voltage of externally loaded TENG.
In spite of recording real-time resistance of the sensor, variation in output voltage is measured as a sensing parameter (Figure 19B). The sensitivity of the TENG-coupled Nb2CTx/PAN sensor is many-fold higher than that of the pristine Nb2CTx/PAN sensor. It is revealed that the sensitivity of the TENG coupled Nb2CTx/PAN sensor is 2.57% per ppm of NH3 with a low response time of 105 s. It is ascribed to the various factors such as synergistic effects due to the formation of the p–n junction between the precursors, TENG presence, and choice of sensing parameter and measuring technique [56].
Hence, the MXP-NCs sensors perform better while coupled with TENG and possess potential application in smart self-driven NH3 sensors with fewer energy requirements. However, the design of compact and optimal TENG with advanced sensing configuration is required to address its commercial prospects.

9.4. Monitoring of Environmental Contaminated by Atmospheric NH3

Atmospheric NH3 is a critical environmental pollutant contributing to acidification of the ecosystem, secondary particles formulation, and eutrophication [111,112]. It plays a primary role in the formation of secondary particulate matter by reacting with airborne acidic species (such as oxides of sulfur and nitrogen) to form ammonium ion-based aerosols, which is the foremost constituent of PM2.5 [112,113]. Ammonium-particulate species contribute to degradation in air quality and visibility and affect atmospheric radiative balance [114]. The PM also acts as a potential carrier of coronavirus through aerosol, conveying the virus to a more considerable distance with increased speed [6,115]. Moreover, the inhalation of ammonium PM can induce damage to respiratory, nervous, and renal systems [7,116]. Hence, the monitoring of airborne ammonium contaminants is the primary concern of environmental stakeholders, especially in a pandemic. MXP-NCs sensors are revealed to detect a low concentration of NH3 (as low as 20 ppb) with advanced sensing characteristics such as flexibility, portable, and high selectivity (Table 2, Table 3 and Table 4). Wang et al. [57] revealed high-sensing NH3 performance of a flexible Nb2CTx/PAN sensor (with LDL 20 ppb) at room temperature. The compact size and mechanical flexibility of the designed sensor allow the sensor to be installed at every NH3 emission site, making it a potential candidate to monitor atmospheric NH3 contamination. Hence, MXP-NCs sensors have enormous latency to replace present-day temperature-assisted environmental NH3 sensors.

10. Challenges and Alternative Approaches

The applications of MXP-NCs for NH3 monitoring require imperative advancements related to precursor processing science and sensor fabrication. The main challenge is to develop moderate and safe synthesis approaches to prepare MXenes with low cost and high yield from the aspect of precursor processing [23,24,25,54,117,118,119]. Although HF-free synthesis routes are proven to be better than HF etching routes with fewer defects on the MXene surface, the presence of cations or extra water between the MXene layers may influence its physiochemical properties. The unstable nature of MXenes such as Ti3C2Tx due to prone to oxidation over long-term operations remains another challenge to address.

10.1. Optimization of Concentration of Precursors

A challenge in the fabrication of MXP-NCs is simultaneous optimization of mechanical flexibility with maintaining high conductivity. It is essential to maintain a significant concentration of MXene in MXP-NCs to maintain high electroconductivity. However, the increase in MXene concentration results in a decrease in mechanical stretchability and flexibility. Numerous tribological characterization tests must be performed to evaluate the threshold MXene concentration for each polymer precursor. These facts argue the presence of trade-offs between the NC properties and the precursor concentration for an NC sensor.
Finding the optimal concentration of MXene in NC for fabricating a sensor is challenging. A suitable concentration of MXene is required to form conductive pathways/channels in the polymer matrix. This conductive network can break in a particular strain range due to an increase in interlayer distance between MXene sheets. As the strain increases, the MXene sheets forming a conductive network moves away from each other. Further strain above the threshold will eventually result in the complete breaking of the conductive network. These variations in conductive networks arrangement are manifested by a steady decrease in the transient electrical current through the sensor and the extent of strain experienced by the sensor. However, the high concentration of MXene in MXP-NCs results in the formation of a continuous conductive network throughout the NC in an anticipated strain range. It results in no detectable change in conductance, including at high strains. Hence, the concentration of MXene precursor in NC highly determines the sensitivity of the sensor.
The surface functionalization of MXenes can address these challenges. In various MXP-NCs, surface functionalization is required to enhance the affinity between the precursors. Hydroxyl groups on the MXene surface are reactive and can endure various surface functionalization reactions. The formation of hydrogen bonds between hydroxyl groups on the MXene surface and polymer precursor improves lifespan and workability. It also reduces the percolation threshold of MXene precursor in the NC sensor due to forming a well-oriented ordered MXene network. It has been revealed by Guo et al. [120] that the percolation of the threshold of MXene in a nonreactive polymer is reduced from 40 to 6 wt% due to the formation of hydrogen bonds on surface modifications. These hydrogen bonds also improve the mechanical flexibility and stretchability of the sensor, as they work as stress transfer networks between the precursors. Similar observations were made by Zhang et al. [121] for MXene/polydimethylsiloxane sensor. Hence, surface modification of MXene precursor serves as an alternative approach for simultaneous optimization of electroconductivity and mechanical flexibility.
Another alternative approach is using secondary nanoparticles for surface modification of MXenes in NCs, which results in the formation of mixed-dimensional structures. The secondary nanoparticles can be zero-dimensional (size less than 100 nm: silver quantum dots), one-dimensional (such as carbon nanotubes), two-dimensional (graphene), and three-dimensional (modified graphene oxide). Generally, secondary nanoparticles are modified using coupling agents such as silane to provide them with a positive charge at the surface. It results in electrostatic interaction between MXene and secondary nanoparticles, causing the formation of a ternary NC.
Zhou et al. [122] revealed the improved carbon dioxide sensing performance of ternary nanocomposite of nitrogen-doped MXene, Ti3C2Tx (N-MXene), polyethyleneimine (PEI), and reduced graphene oxide (rGO). It is ascribed to the surface modification of MXene through nitrogen doping and the presence of secondary nanoparticles (rGO). Hence, the sensing performance of MXP-NCs can be improved by surface modification of an MXene precursor through any coupling agent or secondary nanoparticles due to hydrogen bonding/electrostatic interactions.

10.2. Slower Response

Additionally, MXP-NCs exhibit a slower response toward NH3, limiting its commercial prospects. It can be decreased using a hierarchically interlocked structure (such as Ti3C2Tx/CPAM) to fabricate an NH3 sensor. Based on reported studies, two strategies can be adopted to design state-of-the-art MXP-NC sensors. First, the form of well-distributed conductive channels of MXene in a polymer matrix with increased interlayer separation in NC. Second, improve the deformability of the sensing layer by improving contact between precursors, resulting in hetero-interfacial junctions.

10.3. Mass Production

The technical challenges related to mass production and process integration of MXP-NC sensors should be addressed to meet industrial development. It can be achieved through developing one-pot synthesis approaches, in which MXene flakes can be exfoliated during the in situ polymerization of monomers, aiding the scalable production of MXP-NCs.
Addressing the challenges with alternative technologies will further accelerate novel MXP-NCs sensors’ development and explore its potential in widespread application fields.

11. Conclusions, Prospects, and Viewpoints

The last several years have witnessed rapid advancement in the investigation of MXP-NCs for sensing applications due to synergies between precursors’ advantages (polymers and MXene). Generally, the stability and mechanical flexibility of MXene are improved, and the electrical, mechanical, and thermal properties of the polymer are enhanced in MXP-NCs, while other properties are still under examination. This review has focused on the advancements in fabrication strategies, fascinating structures, and unique properties of MXP-NCs NH3 sensors. The reported NH3 sensing performances of MXP-NC sensors have been comparatively analyzed in several comprehensive tables. In addition, diversified applications of MXP-NCs NH3 sensors in medical, industrial, environmental, and agricultural industries have been summarized. As per published literature, MXP-NCs have great potential in NH3 monitoring, and their overall sensing performances are better than those of other sensing materials. Nevertheless, the development of MXP-NCs NH3 sensors is in its infancy, and there are still many scientific quests remaining before revealing the commercial applications of these materials.
MXP-NCs sensors open a prospect for NH3-monitoring applications in diversified fields, including bio-medical, industrial, and agricultural. Contrary to conventional metal oxide-based sensors, their operation at room temperature makes them cost effective and energy efficient due to the exclusion of micro-heating components. These NH3 sensors are user friendly due to the exclusion of any toxic heavy metal and contribute to negligible toxic nano-waste. Since they can be operated in chemiresistive mode, they are portable, compact, easily handled, simple to configure, and can be produced on a large scale, dissimilar to spectroscopic or electrochemical sensors. However, studies on optimizing large electrical conductivity for mass production of MXene still require ordinate attention. Their hydrophilic nature makes it easier to fabricate flexible and compact NH3 sensors, which can be utilized at every emission site. MXP-NCs NH3 sensors also possess the potential to fabricate smart and intelligent technologies to detect NH3 volatilization in agricultural fields. The high sensitivity makes MXP-NC sensors rapidly detect minute NH3 leakage for workplace safety. However, the optimization of response and recovery time still requires improvement. MXPs-NC NH3 sensors can substitute expensive and sophisticated clinical tests to detect urea concentration in gastric or renal patients owing to their low detection limit. They can detect the low concentration of NH3 (as low as 20 ppb) in exhaled human breath and can result in an optimized real-time NH3 monitoring system with low cost and simple configuration. However, more emphasis on clinical trial data is required to explore its exact potential to replace other urea detecting methods and tests. Hence, MXP-NC sensors give new prospects to NH3 monitoring in rapid and low-level detection, including intelligent technologies, environment, and user-friendly, compact, flexible, and portable nature, and high sensitivity.
The advanced NH3 sensing features of MXP-NCs are ascribed to their unique structural and functional properties such as sizeable interlayer distance, exfoliation, surface functionalities, formation of hetero-interfacial junctions between the precursors, synergistic effects, enhanced porosity, improved charge transport, high thermal and mechanical stability, and high specific surface area. However, the reports on MXP-NCs-based NH3 sensors are scarce and require dedicated attention from the scientific community.
Additionally, polymers are abundant in nature, which has been used for NH3 monitoring. The use of advanced and functional polymers can further facilitate the NH3 detection characteristics of MXP-NCs. Hence, various combinations of MXene and polymers are yet to fabricate and utilized for NH3 sensing.
Moreover, there is a vast scope of research in optimizing the properties of MXP-NCs for NH3 sensing through optimizing the nature and concentration of precursors. The NH3 sensing mechanism can be experimentally explored using advanced spectroscopic/microscopy systems, including in situ FTIR or in situ Raman or kelvin probe force microscopy (KPFM) techniques in the presence of NH3.
Consequently, challenges in the processing of MXP-NCs NH3 sensor can be summarized as a requirement of large-scale production without compromising the electrical conductivity, optimizing the concentration of precursors (MXene and polymer) for better sensing characteristics, controlling synergistic effects, optimizing humidity with sensing performance, managing durability along with flexibility, choice of precursors and surface modifications. Hence, the colossal possibilities in NH3 sensing are addressing these challenges for the development of novel MXP-NC sensors. Less explored fields such as food engineering and electric nose can also benefit from MXP-NC sensors by incorporating recent advancements such as machine learning and pattern recognition. In the future, MXP-NC sensors (Figure 20) application fields are expected to expand and evolve using artificial intelligence.

Funding

This research received no external funding to support this review work.

Acknowledgments

The authors wish to acknowledge the Department of Science and Technology, Government of India, and Vice-Chancellor, University of Delhi, India for providing a facility of resources. Y.K.M. acknowledges funding by Interreg Deutschland-Denmark with money from the European Regional Development Fund, project number 096-1.1-18 (Access and Acceleration).

Conflicts of Interest

Authors declare no conflicts of interest.

References

  1. Valera-Medina, A.; Amer-Hatem, F.; Azad, A.K.; Dedoussi, I.C.; de Joannon, M.; Fernandes, R.X.; Glarborg, P.; Hashemi, H.; He, X.; Mashruk, S.; et al. Review on Ammonia as a Potential Fuel: From Synthesis to Economics. Energy Fuels 2021, 35, 6964–7029. [Google Scholar] [CrossRef]
  2. Malik, R.; Tomer, V.K.; Mishra, Y.K.; Lin, L. Functional gas sensing nanomaterials: A panoramic view. Appl. Phys. Rev. 2020, 7, 021301. [Google Scholar] [CrossRef] [Green Version]
  3. Chaudhary, V.; Kaur, A. Enhanced and selective ammonia sensing behaviour of poly(aniline co-pyrrole) nanospheres chemically oxidative polymerized at low temperature. J. Ind. Eng. Chem. 2015, 26, 143–148. [Google Scholar] [CrossRef]
  4. Insausti, M.; Timmis, R.; Kinnersley, R.; Rufino, M. Advances in sensing ammonia from agricultural sources. Sci. Total Environ. 2020, 706, 135124. [Google Scholar] [CrossRef]
  5. Wu, Y.; Gu, B.; Erisman, J.W.; Reis, A.; Fang, Y.; Lu, X.; Zhang, X. PM2.5 pollution is substantially affected by ammonia emissions in China. Environ. Pollut. 2016, 218, 86–94. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Comunian, S.; Dongo, D.; Milani, C.; Palestini, P. Air Pollution and COVID Total 19: The Role of Particulate Matter in the Spread and Increase of COVID-19’s Morbidity and Mortality. Int. J. Environ. Res. Public Health 2020, 17, 4487. [Google Scholar] [CrossRef] [PubMed]
  7. Ricci, P.P.; Gregory, O.J. Sensors for the detection of ammonia as a potential biomarker for health screening. Sci. Rep. 2021, 11, 1–7. [Google Scholar] [CrossRef] [PubMed]
  8. Chauhan, N.; Gupta, S.; Avasthi, D.K.; Adelung, R.; Mishra, Y.K.; Jain, U. Zinc Oxide Tetrapods Based Biohybrid Interface for Voltammetric Sensing of Helicobacter pylori. ACS Appl. Mater. Interfaces 2018, 10, 30631–30639. [Google Scholar] [CrossRef]
  9. Seekaew, Y.; Pon-On, W.; Wongchoosuk, C. Ultrahigh Selective Room-Temperature Ammonia Gas Sensor Based on Tin–Titanium Dioxide/reduced Graphene/Carbon Nanotube Nanocomposites by the Solvothermal Method. ACS Omega 2019, 4, 16916–16924. [Google Scholar] [CrossRef] [Green Version]
  10. Kwak, D.; Lei, Y.; Maric, R. Ammonia gas sensors: A comprehensive review. Talanta 2019, 204, 713–730. [Google Scholar] [CrossRef]
  11. Aarya, S.; Kumar, Y.; Chahota, R.K. Recent Advances in Materials, Parameters, Performance and Technology in Ammonia Sensors: A Review. J. Inorg. Organomet. Polym. Mater. 2019, 30, 269–290. [Google Scholar] [CrossRef]
  12. Zhuang, X.; Zhang, D.; Wang, X.; Yu, X.; Yu, J. Biocompatible and degradable gelatin dielectric based low-operating voltage organic transistors for ultra-high sensitivity NH3 detection. Appl. Phys. Lett. 2018, 113, 263301. [Google Scholar] [CrossRef] [Green Version]
  13. Alrammouz, R.; Podlecki, J.; Abboud, P.; Sorli, B.; Habchi, R. A review on flexible gas sensors: From materials to devices. Sens. Actuators A Phys. 2018, 284, 209–231. [Google Scholar] [CrossRef]
  14. Yang, S.; Jiang, C.; Wei, S.-H. Gas sensing in 2D materials. Appl. Phys. Rev. 2017, 4, 021304. [Google Scholar] [CrossRef]
  15. Mishra, R.K.; Choi, G.J.; Mishra, Y.K.; Kaushik, A.; Sohn, Y.; Lee, S.H.; Gwag, J.S. A highly stable, selective, and high-performance VOC sensor using a SnS2 nano-lotus structure. J. Mater. Chem. C 2021, 9, 7713. [Google Scholar] [CrossRef]
  16. González-Garnica, M.; Galdámez-Martínez, A.; Malagón, F.; Ramos, C.; Santana, G.; Abolhassani, R.; Panda, P.K.; Kaushik, A.; Mishra, Y.K.; Karthik, T.V.; et al. One dimensional Au-ZnO hybrid nanostructures based CO2 detection: Growth mechanism and role of the seed layer on sensing performance. Sens. Actuators B Chem. 2021, 337, 129765. [Google Scholar] [CrossRef]
  17. Khan, S.B. Gas Sensors, 1st ed.; IntechOpen: London, UK, 2020. [Google Scholar]
  18. Anichini, C.; Czepa, W.; Pakulski, D.; Aliprandi, A.; Ciesielski, A.; Samorì, P. Chemical sensing with 2D materials. Chem. Soc. Rev. 2018, 47, 4860–4908. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  19. Wang, L.; Zhang, M.; Yang, B.; Tan, J.; Ding, X.; Li, W. Recent Advances in Multidimensional (1D, 2D, and 3D) Composite Sensors Derived from MXene: Synthesis, Structure, Application, and Perspective. Small Methods 2021, 5, 2100409. [Google Scholar] [CrossRef]
  20. Mathew, M.; Shinde, P.V.; Samal, R.; Rout, C.S. A review on mechanisms and recent developments in p-n heterojunctions of 2D materials for gas sensing applications. J. Mater. Sci. 2021, 56, 9575–9604. [Google Scholar] [CrossRef]
  21. Rathee, G.; Bartwal, G.; Rathee, J.; Mishra, Y.K.; Kaushik, A.; Solanki, P.R. Emerging Multimodel Zirconia Nanosystems for High-Performance Biomedical Applications. Adv. NanoBiomed Res. 2021, 1, 2100039. [Google Scholar] [CrossRef]
  22. Chaudhary, V. High performance X-band electromagnetic shields based on methyl-orange assisted polyaniline-silver core-shell nanocomposites. Polym. Technol. Mater. 2021, 60, 1–10. [Google Scholar] [CrossRef]
  23. VahidMohammadi, A.; Rosen, J.; Gogotsi, Y. A Review outlines the development and potential applications of a wide range of 2D metal carbides and nitrides. Science 2021, 372, 6547. [Google Scholar] [CrossRef]
  24. Aghaei, S.M.; Aasi, A.; Panchapakesan, B. Experimental and Theoretical Advances in MXene-Based Gas Sensors. ACS Omega 2021, 6, 2450–2461. [Google Scholar] [CrossRef] [PubMed]
  25. Riazi, H.; Taghizadeh, G.; Soroush, M. MXene-Based Nanocomposite Sensors. ACS Omega 2021, 6, 11103–11112. [Google Scholar] [CrossRef] [PubMed]
  26. Zamhuri, A.; Lim, G.P.; Ma, N.L.; Tee, K.S.; Soon, C.F. MXene in the lens of biomedical engineering: Synthesis, applications and future outlook. Biomed. Eng. Online 2021, 20, 1–24. [Google Scholar] [CrossRef] [PubMed]
  27. Zhan, X.; Si, C.; Zhou, J.; Sun, Z. MXene and MXene-based composites: Synthesis, properties and environment-related applications. Nanoscale Horiz. 2019, 5, 235–258. [Google Scholar] [CrossRef]
  28. Wong, Y.C.; Ang, B.C.; Haseeb, A.S.M.A.; Baharuddin, A.A.; Wong, Y.H. Review—Conducting Polymers as Chemiresistive Gas Sensing Materials: A Review. J. Electrochem. Soc. 2019, 167, 037503. [Google Scholar] [CrossRef]
  29. Nazemi, H.; Joseph, A.; Park, J.; Emadi, A. Advanced Micro- and Nano-Gas Sensor Technology: A Review. Sensors 2019, 19, 1285. [Google Scholar] [CrossRef] [Green Version]
  30. Ansari, N.; Lone, M.Y.; Shumaila; Ali, J.; Zulfequar, M.; Husain, M.; Islam, S.S.; Husain, S. Trace level toxic ammonia gas sensing of single-walled carbon nanotubes wrapped polyaniline nanofibers. J. Appl. Phys. 2020, 127, 044902. [Google Scholar] [CrossRef]
  31. Baker, C.O.; Huang, X.; Nelson, W.; Kaner, R.B. Polyaniline nanofibers: Broadening applications for conducting polymers. Chem. Soc. Rev. 2017, 46, 1510–1525. [Google Scholar] [CrossRef]
  32. Chaudhary, V.; Royal, A.; Chavali, M.; Yadav, S.K. Advancements in research and development to combat COVID-19 using nanotechnology. Nanotechnol. Environ. Eng. 2021, 6, 1–15. [Google Scholar] [CrossRef]
  33. Fu, S.-Y.; Sun, Z.; Huang, P.; Li, Y.-Q.; Hu, N. Some basic aspects of polymer nanocomposites: A critical review. Nano Mater. Sci. 2019, 1, 2–30. [Google Scholar] [CrossRef]
  34. Chaudhary, V.; Kaur, A. Enhanced room temperature sulfur dioxide sensing behaviour of in situ polymerized polyaniline–tungsten oxide nanocomposite possessing honeycomb morphology. RSC Adv. 2015, 5, 73535–73544. [Google Scholar] [CrossRef]
  35. Zhang, A.; Liu, R.; Tian, J.; Huang, W.; Liu, J. MXene-Based Nanocomposites for Energy Conversion and Storage Applications. Chem. A Eur. J. 2020, 26, 6342–6359. [Google Scholar] [CrossRef] [PubMed]
  36. Hasan, M.; Hossain, M.; Chowdhury, H.K. Two-dimensional MXene-based flexible nanostructures for functional nanodevices: A review. J. Mater. Chem. A 2021, 9, 3231–3269. [Google Scholar] [CrossRef]
  37. Ling, Z.; Ren, C.E.; Zhao, M.-Q.; Yang, J.; Giammarco, J.M.; Qiu, J.; Barsoum, M.W.; Gogotsi, Y. Flexible and conductive MXene films and nanocomposites with high capacitance. Proc. Natl. Acad. Sci. USA 2014, 111, 111–16676. [Google Scholar] [CrossRef] [Green Version]
  38. Carey, M.; Barsoum, M. MXene polymer nanocomposites: A review. Mater. Today Adv. 2021, 9, 100120. [Google Scholar] [CrossRef]
  39. Chen, X.; Zhao, Y.; Li, L.; Wang, Y.; Wang, J.; Xiong, J.; Du, S.; Zhang, P.; Shi, X.; Yu, J. MXene/Polymer Nanocomposites: Preparation, Properties, and Applications. Polym. Rev. 2021, 61, 80–115. [Google Scholar] [CrossRef]
  40. Kausar, A. Polymer/MXene nanocomposite—A new age for advanced materials. Polym. Technol. Mater. 2021, 60, 1377–1392. [Google Scholar] [CrossRef]
  41. Lee, E.; VahidMohammadi, A.; Prorok, B.C.; Yoon, Y.S.; Beidaghi, M.; Kim, D.-J. Room Temperature Gas Sensing of Two-Dimensional Titanium Carbide (MXene). ACS Appl. Mater. Interfaces 2017, 9, 37184–37190. [Google Scholar] [CrossRef] [PubMed]
  42. Ho, D.H.; Choi, Y.Y.; Jo, S.B.; Myoung, J.; Cho, J.H. Sensing with MXenes: Progress and Prospects. Adv. Mater. 2021, 2005846. [Google Scholar] [CrossRef] [PubMed]
  43. Naguib, M.; Halim, J.; Lu, J.; Cook, K.M.; Hultman, L.; Gogotsi, Y.; Barsoum, M.W. New Two-Dimensional Niobium and Vanadium Carbides as Promising Materials for Li-Ion Batteries. J. Am. Chem. Soc. 2013, 135, 15966–15969. [Google Scholar] [CrossRef] [PubMed]
  44. Kim, S.J.; Koh, H.-J.; Ren, C.E.; Kwon, O.; Maleski, K.; Cho, S.; Anasori, B.; Kim, C.-K.; Choi, Y.-K.; Kim, J.; et al. Metallic Ti3C2Tx MXene Gas Sensors with Ultrahigh Signal-to-Noise Ratio. ACS Nano 2018, 12, 986–993. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Zhao, L.; Wang, K.; Wei, W.; Wang, L.; Han, W. High-performance flexible sensing devices based on polyaniline/MXene nanocomposites. InfoMat 2019, 1, 407–416. [Google Scholar] [CrossRef] [Green Version]
  46. Kaushik, A.; Kumar, R.; Arya, S.K.; Nair, M.; Malhotra, B.D.; Bhansali, S. Organic–Inorganic Hybrid Nanocomposite-Based Gas Sensors for Environmental Monitoring. Chem. Rev. 2015, 115, 4571–4606. [Google Scholar] [CrossRef]
  47. Chen, J.; Chen, K.; Tong, D.; Huang, Y.; Zhang, J.; Xue, J.; Huang, Q.; Chen, T. CO2 and temperature dual responsive “Smart” MXene phases. Chem. Commun. 2015, 51, 314–317. [Google Scholar] [CrossRef] [PubMed]
  48. Fu, J.; Yun, J.; Wu, S.; Li, L.; Yu, L.; Kim, K.H. Architecturally Robust Graphene-Encapsulated MXene Ti2CTx@Polyaniline Composite for High-Performance Pouch-Type Asymmetric Supercapacitor. ACS Appl. Mater. Interfaces 2018, 10, 34212–34221. [Google Scholar] [CrossRef]
  49. Wang, X.; Sun, K.; Li, K.; Li, X.; Gogotsi, Y. Ti3C2T/PEDOT:PSS hybrid materials for room-temperature methanol sensor. Chin. Chem. Lett. 2020, 31, 1018–1021. [Google Scholar] [CrossRef]
  50. Zhu, M.; Huang, Y.; Deng, Q.; Zhou, J.; Pei, Z.; Xue, Q.; Wang, Z.; Li, H.; Huang, Q.; Zhi, C. Highly Flexible, Freestanding Supercapacitor Electrode with Enhanced Performance Obtained by Hybridizing Polypyrrole Chains with MXene. Adv. Energy Mater. 2016, 6, 1600969. [Google Scholar] [CrossRef]
  51. Zhang, P.; Yang, X.-J.; Li, P.; Zhao, Y.; Niu, Q.J. Fabrication of novel MXene (Ti3C2)/polyacrylamide nanocomposite hydrogels with enhanced mechanical and drug release properties. Soft Matter 2020, 16, 162–169. [Google Scholar] [CrossRef]
  52. Li, X.; Xu, J.; Jiang, Y.; He, Z.; Liu, B.; Xie, H.; Li, H.; Li, Z.; Wang, Y.; Tai, H. Toward agricultural ammonia volatilization monitoring: A flexible polyaniline/Ti3C2T hybrid sensitive films based gas sensor. Sens. Actuators B Chem. 2020, 316, 128144. [Google Scholar] [CrossRef]
  53. Jiang, X.; Kuklin, A.; Baev, A.; Ge, Y.; Ågren, H.; Zhang, H.; Prasad, P.N. Two-Dimensional MXenes: From morphological to optical, electric, and magnetic properties and applications. Phys. Rep. 2020, 848, 1–58. [Google Scholar] [CrossRef]
  54. Naguib, M.; Kurtoglu, M.; Presser, V.; Lu, J.; Niu, J.; Heon, M.; Hultman, L.; Gogotsi, Y.; Barsoum, M.W. Two-Dimensional Nanocrystals Produced by Exfoliation of Ti3AlC. Adv. Mater. 2011, 23, 4248–4253. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Lee, E.; Kim, D.-J. Review—Recent Exploration of Two-Dimensional MXenes for Gas Sensing: From a Theoretical to an Experimental View. J. Electrochem. Soc. 2020, 167, 037515. [Google Scholar] [CrossRef] [Green Version]
  56. Wang, S.; Liu, B.; Duan, Z.; Zhao, Q.; Zhang, Y.; Xie, G.; Jiang, Y.; Li, S.; Tai, H. PANI nanofibers-supported Nb2CTx nanosheets-enabled selective NH3 detection driven by TENG at room temperature. Sens. Actuators B Chem. 2021, 327, 128923. [Google Scholar] [CrossRef]
  57. Wang, S.; Jiang, Y.; Liu, B.; Duan, Z.; Pan, H.; Yuan, Z.; Xie, G.; Wang, J.; Fang, Z.; Tai, H. Ultrathin Nb2CT nanosheets-supported polyaniline nanocomposite: Enabling ultrasensitive NH3 detection. Sens. Actuators B Chem. 2021, 343, 130069. [Google Scholar] [CrossRef]
  58. Gogotsi, Y.; Anasori, B. The Rise of MXenes. ACS Nano 2019, 13, 8491–8494. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  59. Xie, J.; Wang, X.; Li, A.; Li, F.; Zhou, Y. Corrosion behavior of selected Mn+1AXn phases in hot concentrated HCl solution: Effect of A element and MX layer. Corros. Sci. 2012, 60, 129–135. [Google Scholar] [CrossRef]
  60. Peng, J.; Chen, X.; Ong, W.-J.; Zhao, X.; Li, N. Surface and Heterointerface Engineering of 2D MXenes and Their Nanocomposites: Insights into Electro- and Photocatalysis. Chem 2019, 5, 18–50. [Google Scholar] [CrossRef] [Green Version]
  61. Zhao, L.; Zheng, Y.; Wang, K.; Lv, C.; Wei, W.; Wang, L.; Han, W. Highly Stable Cross-Linked Cationic Polyacrylamide/Ti3C2Tx MXene Nanocomposites for Flexible Ammonia-Recognition Devices. Adv. Mater. Technol. 2020, 5, 2000248. [Google Scholar] [CrossRef]
  62. Naguib, M.; Saito, T.; Lai, S.; Rager, M.S.; Aytug, T.; Paranthaman, M.P.; Zhao, M.-Q.; Gogotsi, Y. Ti3C2Tx (MXene)–Polyacrylamide nanocomposite films. RSC Adv. 2016, 6, 72069–72073. [Google Scholar] [CrossRef]
  63. Yu, B.; Tawiah, B.; Wang, L.-Q.; Yuen, A.C.Y.; Zhang, Z.-C.; Shen, L.-L.; Lin, B.; Fei, B.; Yang, W.; Li, A.; et al. Interface decoration of exfoliated MXene ultra-thin nanosheets for fire and smoke suppressions of thermoplastic polyurethane elastomer. J. Hazard. Mater. 2019, 374, 110–119. [Google Scholar] [CrossRef]
  64. Sheng, X.; Zhao, Y.; Zhang, L.; Lu, X. Properties of two-dimensional Ti3C2 MXene/thermoplastic polyurethane nanocomposites with effective reinforcement via melt blending. Compos. Sci. Technol. 2019, 181, 107710. [Google Scholar] [CrossRef]
  65. Jin, L.; Wu, C.; Wei, K.; He, L.; Gao, H.; Zhang, H.; Zhang, K.; Asiri, A.M.; Alamry, K.A.; Yang, L.; et al. Polymeric Ti3C2Tx MXene Composites for Room Temperature Ammonia Sensing. ACS Appl. Nano Mater. 2020, 3, 12071–12079. [Google Scholar] [CrossRef]
  66. Qin, L.; Tao, Q.; Liu, X.; Fahlman, M.; Halim, J.; Persson, P.O.; Rosen, J.; Zhang, F. Polymer-MXene composite films formed by MXene-facilitated electrochemical polymerization for flexible solid-state microsupercapacitors. Nano Energy 2019, 60, 734–742. [Google Scholar] [CrossRef]
  67. Carey, M.; Hinton, Z.; Sokol, M.; Alvarez, N.J.; Barsoum, M.W. Nylon-6/Ti3C2Tz MXene Nanocomposites Synthesized by in Situ Ring Opening Polymerization of ε-Caprolactam and Their Water Transport Properties. ACS Appl. Mater. Interfaces 2019, 11, 20425–20436. [Google Scholar] [CrossRef]
  68. Yan, K.; Li, J.; Pan, L.; Shi, Y. Inkjet printing for flexible and wearable electronics. APL Mater. 2020, 8, 120705. [Google Scholar] [CrossRef]
  69. Chaudhary, V.; Chavali, M. Novel methyl-orange assisted core-shell polyaniline-silver nanosheets for highly sensitive ammonia chemiresistors. J. Appl. Polym. Sci. 2021, 138, 51288. [Google Scholar] [CrossRef]
  70. Chaudhary, V.; Singh, H.; Kaur, A. Effect of charge carrier transport on sulfur dioxide monitoring performance of highly porous polyaniline nanofibres. Polym. Int. 2017, 66, 699–704. [Google Scholar] [CrossRef]
  71. Chaudhary, V.; Kaur, A. Solitary surfactant assisted morphology dependent chemiresistive polyaniline sensors for room temperature monitoring of low parts per million sulfur dioxide. Polym. Int. 2015, 64, 1475–1481. [Google Scholar] [CrossRef]
  72. Chaudhary, V.; Kaur, A. Surfactant directed polyaniline nanostructures for high performance sulphur dioxide chemiresistors: Effect of morphologies, chemical structure and porosity. RSC Adv. 2016, 6, 95349–95357. [Google Scholar] [CrossRef]
  73. Liu, Y.; Zhang, J.; Zhang, X.; Li, Y.; Wang, J. Ti3C2Tx Filler Effect on the Proton Conduction Property of Polymer Electrolyte Membrane. ACS Appl. Mater. Interfaces 2016, 8, 20352–20363. [Google Scholar] [CrossRef]
  74. Fei, M.; Lin, R.; Deng, Y.; Xian, H.; Bian, R.; Zhang, X.; Cheng, J.; Xu, C.; Cai, D. Polybenzimidazole/Mxene composite membranes for intermediate temperature polymer electrolyte membrane fuel cells. Nanotechnology 2018, 29, 035403. [Google Scholar] [CrossRef]
  75. Cao, Y.; Deng, Q.; Liu, Z.; Shen, D.; Wang, T.; Huang, Q.; Du, S.; Jiang, N.; Lin, C.-T.; Yu, J. Enhanced thermal properties of poly (vinylidene fluoride) composites with ultrathin nanosheets of MXene. RSC Adv. 2017, 7, 20494–20501. [Google Scholar] [CrossRef] [Green Version]
  76. Kang, R.; Zhang, Z.; Guo, L.; Cui, J.; Chen, Y.; Hou, X.; Wang, B.; Lin, C.-T.; Jiang, N.; Yu, J. Enhanced Thermal Conductivity of Epoxy Composites Filled with 2D Transition Metal Carbides (MXenes) with Ultralow Loading. Sci. Rep. 2019, 9, 1–14. [Google Scholar] [CrossRef] [Green Version]
  77. Liu, R.; Li, W. High-Thermal-Stability and High-Thermal-Conductivity Ti3C2Tx MXene/Poly (vinyl alcohol) (PVA) Composites. ACS Omega 2018, 3, 2609–2617. [Google Scholar] [CrossRef] [Green Version]
  78. Yi, N.; Cheng, Z.; Li, H.; Yang, L.; Zhu, J.; Zheng, X.; Chen, Y.; Liu, Z.; Zhu, H.; Cheng, H. Stretchable, ultrasensitive, and low-temperature NO2 sensors based on MoS2@rGO nanocomposites. Mater. Today Phys. 2020, 15, 100265. [Google Scholar] [CrossRef]
  79. Chaudhary, V. One-dimensional variable range charge carrier hopping in polyaniline–tungsten oxide nanocomposite-based hydrazine chemiresistor. Appl. Phys. A 2021, 127, 1–7. [Google Scholar] [CrossRef]
  80. Debbarma, R.; Nguyen, N.H.L.; Berry, V. Defect guided conduction in graphene-derivatives and MoS2: Two-Dimensional nanomaterial models. Appl. Mater. Today 2021, 23, 101072. [Google Scholar] [CrossRef]
  81. Hajian, S.; Khakbaz, P.; Moshayedi, M.; Maddipatla, D.; Narakathu, B.B.; Turkani, V.S.; Bazuin, B.J.; Pourfath, M.; Atashbar, M.Z. Impact of Different Ratios of Fluorine, Oxygen, and Hydroxyl Surface Terminations on Ti3C2Tx MXene as Ammonia Sensor: A First-Principles Study. In Proceedings of the 2018 IEEE Sensors, New Delhi, India, 28–31 October 2018; Volume 1, pp. 1–4. [Google Scholar]
  82. Zhu, C.; Cheng, X.; Dong, X.; Xu, Y.M. Enhanced Sub-ppm NH3 Gas Sensing Performance of PANI/TiO2 Nanocomposites at Room Temperature. Front. Chem. 2018, 6, 493. [Google Scholar] [CrossRef]
  83. Kumar, L.; Rawal, I.; Kaur, A.; Annapoorni, S. Flexible room temperature ammonia sensor based on polyaniline. Sens. Actuators B Chem. 2017, 240, 408–416. [Google Scholar] [CrossRef]
  84. Šetka, M.; Calavia, R.; Vojkůvka, L.; Polčák, J.; Llobet, E.; Drbohlavová, J.; Vallejos, S. Electrochemically deposited polypyrrole nanorods and study of their ammonia sensing properties. Mater. Today Proc. 2020, 20, 305–310. [Google Scholar] [CrossRef]
  85. Gund, G.; Park, J.H.; Harpalsinh, R.; Kota, M.; Shin, J.H.; Kim, T.-I.; Gogotsi, Y.; Park, H.S. MXene/Polymer Hybrid Materials for Flexible AC-Filtering Electrochemical Capacitors. Joule 2019, 3, 164–176. [Google Scholar] [CrossRef] [Green Version]
  86. Thirumalairajan, S.; Girija, K.; Mastelaro, V.; Ponpandian, N. Surface Morphology-Dependent Room-Temperature LaFeO3 Nanostructure Thin Films as Selective NO2 Gas Sensor Prepared by Radio Frequency Magnetron Sputtering. ACS Appl. Mater. Interfaces 2014, 6, 13917–13927. [Google Scholar] [CrossRef]
  87. Gao, J.; Qin, J.; Chang, J.; Liu, H.; Wu, Z.-S.; Feng, L. NH3 Sensor Based on 2D Wormlike Polypyrrole/Graphene Heterostructures for a Self-Powered Integrated System. ACS Appl. Mater. Interfaces 2020, 12, 38674–38681. [Google Scholar] [CrossRef]
  88. Wu, Z.; Chen, X.; Zhu, S.; Zhou, Z.; Yao, Y.; Quan, W.; Liu, B. Enhanced sensitivity of ammonia sensor using graphene/polyaniline nanocomposite. Sens. Actuators B Chem. 2013, 178, 485–493. [Google Scholar] [CrossRef]
  89. Patil, R.; Gaikwad, S.; Karanjekar, A.; Khanna, P.; Jain, G.; Gaikwad, V.; More, P.; Bisht, N. Optimization of strontium-doping concentration in BaTiO3 nanostructures for room temperature NH3 and NO2 gas sensing. Mater. Today Chem. 2020, 16, 100240. [Google Scholar] [CrossRef]
  90. Lee, E.; VahidMohammadi, A.; Yoon, Y.S.; Beidaghi, M.; Kim, D.-J. Two-Dimensional Vanadium Carbide MXene for Gas Sensors with Ultrahigh Sensitivity Toward Nonpolar Gases. ACS Sens. 2019, 4, 1603–1611. [Google Scholar] [CrossRef]
  91. Seekaew, Y.; Lokavee, S.; Phokharatkul, D.; Wisitsoraat, A.; Kerdcharoen, T.; Wongchoosuk, C. Low-cost and flexible printed graphene–PEDOT:PSS gas sensor for ammonia detection. Org. Electron. 2014, 15, 2971–2981. [Google Scholar] [CrossRef]
  92. Yang, Z.; Liu, A.; Wang, C.; Liu, F.; He, J.; Li, S.; Wang, J.; You, R.; Yan, X.; Sun, P.; et al. Improvement of Gas and Humidity Sensing Properties of Organ-like MXene by Alkaline Treatment. ACS Sens. 2019, 4, 1261–1269. [Google Scholar] [CrossRef]
  93. 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 Chem. 2018, 273, 726–734. [Google Scholar] [CrossRef]
  94. Wang, S.; Xie, G.; Tai, H.; Su, Y.; Yang, B.; Zhang, Q.; Du, X.; Jiang, Y. Ultrasensitive flexible self-powered ammonia sensor based on triboelectric nanogenerator at room temperature. Nano Energy 2018, 51, 231–240. [Google Scholar] [CrossRef]
  95. Du, C.-F.; Zhao, X.; Wang, Z.; Yu, H.; Ye, Q. Recent Advanced on the MXene–Organic Hybrids: Design, Synthesis, and Their Applications. Nanomaterials 2021, 11, 166. [Google Scholar] [CrossRef]
  96. Yu, X.-F.; Li, Y.-C.; Cheng, J.-B.; Liu, Z.-B.; Li, Q.-Z.; Li, W.-Z.; Yang, X.; Xiao, B. Monolayer Ti2CO2: A Promising Candidate for NH3 Sensor or Capturer with High Sensitivity and Selectivity. ACS Appl. Mater. Interfaces 2015, 7, 13707–13713. [Google Scholar] [CrossRef] [PubMed]
  97. Kulkarni, S.B.; Navale, Y.H.; Navale, S.T.; Stadler, F.; Patil, V.B. Room temperature ammonia gas sensing properties of polyaniline nanofibers. J. Mater. Sci. Mater. Electron. 2019, 30, 8371–8380. [Google Scholar] [CrossRef]
  98. Pandey, S. Highly sensitive and selective chemiresistor gas/vapor sensors based on polyaniline nanocomposite: A comprehensive review. J. Sci. Adv. Mater. Devices 2016, 1, 431–453. [Google Scholar] [CrossRef] [Green Version]
  99. Abdulla, S.; Mathew, T.L.; Pullithadathil, B. Highly sensitive, room temperature gas sensor based on polyaniline-multiwalled carbon nanotubes (PANI/MWCNTs) nanocomposite for trace-level ammonia detection. Sens. Actuators B Chem. 2015, 221, 1523–1534. [Google Scholar] [CrossRef]
  100. Mogera, U.; Sagade, A.; George, S.J.; Kulkarni, G.U. Ultrafast response humidity sensor using supramolecular nanofibre and its application in monitoring breath humidity and flow. Sci. Rep. 2015, 4, 4103. [Google Scholar] [CrossRef] [Green Version]
  101. Tanguy, N.R.; Thompson, M.; Yan, N. A review on advances in application of polyaniline for ammonia detection. Sensors Actuators B Chem. 2018, 257, 1044–1064. [Google Scholar] [CrossRef]
  102. Jha, R.K.; Wan, M.; Jacob, C.; Guha, P.K. Ammonia vapour sensing properties of in situ polymerized conducting PANI-nanofiber/WS2 nanosheet composites. New J. Chem. 2018, 42, 735–745. [Google Scholar] [CrossRef]
  103. Matsuguchi, M.; Io, J.; Sugiyama, G.; Sakai, Y. Effect of NH3 gas on the electrical conductivity of polyaniline blend films. Synth. Met. 2002, 128, 15–19. [Google Scholar] [CrossRef]
  104. Ly, T.N.; Park, S. Highly sensitive ammonia sensor for diagnostic purpose using reduced graphene oxide and conductive polymer. Sci. Rep. 2018, 8, 1–12. [Google Scholar] [CrossRef] [PubMed]
  105. Li, H.-Y.; Lee, C.-S.; Kim, D.H.; Lee, J.-H. Flexible Room-Temperature NH3 Sensor for Ultrasensitive, Selective, and Humidity-Independent Gas Detection. ACS Appl. Mater. Interfaces 2018, 10, 27858–27867. [Google Scholar] [CrossRef]
  106. Ma, Z.; Yue, Y.; Feng, M.; Li, Y.; Ma, X.; Zhao, X.; Wang, S. Mitigation of ammonia volatilization and nitrate leaching via loss control urea triggered H-bond forces. Sci. Rep. 2019, 9, 15140. [Google Scholar] [CrossRef] [Green Version]
  107. Wei, S.; Bai, Z.; Chadwick, D.; Hou, Y.; Qin, W.; Zhao, Z.; Jiang, R.; Ma, L. Greenhouse gas and ammonia emissions and mitigation options from livestock production in peri-urban agriculture: Beijing—A case study. J. Clean. Prod. 2018, 178, 515–525. [Google Scholar] [CrossRef]
  108. Zhu, J.; Zhu, M.; Shi, Q.; Wen, F.; Liu, L.; Dong, B.; Haroun, A.; Yang, Y.; Vachon, P.; Guo, X.; et al. Progress in TENG technology—A journey from energy harvesting to nanoenergy and nanosystem. EcoMat 2020, 2, 12058. [Google Scholar] [CrossRef]
  109. Liu, Y.; Zheng, Y.; Wu, Z.; Zhang, L.; Sun, W.; Li, T.; Wang, D.; Zhou, F. Conductive elastic sponge-based triboelectric nanogenerator (TENG) for effective random mechanical energy harvesting and ammonia sensing. Nano Energy 2021, 79, 105422. [Google Scholar] [CrossRef]
  110. Kim, D.W.; Lee, J.H.; Kim, J.K.; Jeong, U. Material aspects of triboelectric energy generation and sensors. NPG Asia Mater. 2020, 12, 6. [Google Scholar] [CrossRef]
  111. Sutton, M.A.; Erisman, J.W.; Dentener, F.; Möller, D. Ammonia in the environment: From ancient times to the present. Environ. Pollut. 2008, 156, 583–604. [Google Scholar] [CrossRef]
  112. Wang, S.; Nan, J.; Shi, C.; Fu, Q.; Gao, S.; Wang, D.; Cui, H.; Saiz-Lopez, A.; Zhou, B. Atmospheric ammonia and its impacts on regional air quality over the megacity of Shanghai, China. Sci. Rep. 2015, 5, 15842. [Google Scholar] [CrossRef] [Green Version]
  113. Updyke, K.M.; Nguyen, T.; Nizkorodov, S.A. Formation of brown carbon via reactions of ammonia with secondary organic aerosols from biogenic and anthropogenic precursors. Atmos. Environ. 2012, 63, 22–31. [Google Scholar] [CrossRef]
  114. Park, R.S.; Lee, S.; Shin, S.-K.; Song, C.H. Contribution of ammonium nitrate to aerosol optical depth and direct radiative forcing by aerosols over East Asia. Atmos. Chem. Phys. Discuss. 2014, 14, 2185–2201. [Google Scholar] [CrossRef] [Green Version]
  115. Nor, N.S.M.; Yip, C.W.; Ibrahim, N.; Jaafar, M.H.; Rashid, Z.Z.; Mustafa, N.; Hamid, H.H.A.; Chandru, K.; Latif, M.T.; Saw, P.E.; et al. Particulate matter (PM2.5) as a potential SARS-CoV-2 carrier. Sci. Rep. 2021, 11, 1–6. [Google Scholar] [CrossRef]
  116. Mendy, A.; Wu, X.; Keller, J.L.; Fassler, C.S.; Apewokin, S.; Mersha, T.B.; Xie, C.; Pinney, S.M. Long-term exposure to fine particulate matter and hospitalization in COVID-19 patients. Respir. Med. 2021, 178, 106313. [Google Scholar] [CrossRef] [PubMed]
  117. Khunger, A.; Kaur, N.; Mishra, Y.K.; Chaudhary, G.R.; Kaushik, A. Perspective and prospects of 2D MXenes for smart biosensing. Mater. Lett. 2021, 304, 130656. [Google Scholar] [CrossRef]
  118. Kaur, N.; Khunger, A.; Wallen, S.L.; Kaushik, A.; Chaudhary, G.R.; Varma, R.S. Advanced green analytical chemistry for environmental pesticide detection. Curr. Opin. Green Sustain. Chem. 2021, 30, 100488. [Google Scholar] [CrossRef]
  119. Chaudhary, V. Charge Carrier Dynamics of Electrochemically Synthesized Poly (Aniline Co-Pyrrole) Nanospheres Based Sulfur Dioxide Chemiresistor. Polym. Plast. Technol. Mater. 2021, 1–9. [Google Scholar] [CrossRef]
  120. Guo, Q.; Zhang, X.; Zhao, F.; Song, Q.; Su, G.; Tan, Y.; Tao, Q.; Zhou, T.; Yu, Y.; Zhou, Z.; et al. Protein-Inspired Self-Healable Ti3C2 MXenes/Rubber-Based Supramolecular Elastomer for Intelligent Sensing. ACS Nano 2020, 14, 2788–2797. [Google Scholar] [CrossRef]
  121. Zhang, K.; Sun, J.; Song, J.; Gao, C.; Wang, Z.; Song, C.; Wu, Y.; Liu, Y. Self-Healing Ti3C2 MXene/PDMS Supramolecular Elastomers Based on Small Biomolecules Modification for Wearable Sensors. ACS Appl. Mater. Interfaces 2020, 12, 45306. [Google Scholar] [CrossRef] [PubMed]
  122. Zhou, Y.; Wang, Y.; Wang, Y.; Li, X. Humidity-Enabled Ionic Conductive Trace Carbon Dioxide Sensing of Nitrogen-Doped Ti3C2Tx MXene/Polyethyleneimine Composite Films Decorated with Reduced Graphene Oxide Nanosheets. Anal. Chem. 2020, 92, 16033–16042. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Schematic of state-of-the-art NH3 sensing techniques using different sensing signals.
Figure 1. Schematic of state-of-the-art NH3 sensing techniques using different sensing signals.
Nanomaterials 11 02496 g001
Figure 2. Schematic of various ammonia sensing techniques using different sensing signals.
Figure 2. Schematic of various ammonia sensing techniques using different sensing signals.
Nanomaterials 11 02496 g002
Figure 3. Statistics on yearly publication in the field of MXene–polymer nanocomposites (especially properties and processing of NCs). Reprinted from ref. [38].
Figure 3. Statistics on yearly publication in the field of MXene–polymer nanocomposites (especially properties and processing of NCs). Reprinted from ref. [38].
Nanomaterials 11 02496 g003
Figure 4. Possible interactions in polymer nanocomposites. Reprinted with permission from ref. [46]. Copyright 2015 American Chemical Society.
Figure 4. Possible interactions in polymer nanocomposites. Reprinted with permission from ref. [46]. Copyright 2015 American Chemical Society.
Nanomaterials 11 02496 g004
Figure 5. Schematic of fabrication of MXene using selective Etching of MAX phase through wet and dry route.
Figure 5. Schematic of fabrication of MXene using selective Etching of MAX phase through wet and dry route.
Nanomaterials 11 02496 g005
Figure 6. Schematic of synthesis routes for MXP-NCs. Ex situ route: (a) blending, (b) alternate deposition. In situ route: (c) in situ polymerization.
Figure 6. Schematic of synthesis routes for MXP-NCs. Ex situ route: (a) blending, (b) alternate deposition. In situ route: (c) in situ polymerization.
Nanomaterials 11 02496 g006
Figure 7. Different reported strategies for fabrication of an ammonia-sensing device using MXP-NCs.
Figure 7. Different reported strategies for fabrication of an ammonia-sensing device using MXP-NCs.
Nanomaterials 11 02496 g007
Figure 8. Morphological, structural, and chemical element analysis of PAN/Ti3C2Tx NC with (a) Morphology of Pristine Ti3C2Tx, (b,c) Morphology of PAN/Ti3C2Tx NC, (d) Structure of PAN/Ti3C2Tx NC through XRD Analysis, (e,f) XPS of PANI and PAN/Ti3C2Tx NC for chemical element analysis. Reprinted with permission from ref. [52] Copyright 2020 Elsevier.
Figure 8. Morphological, structural, and chemical element analysis of PAN/Ti3C2Tx NC with (a) Morphology of Pristine Ti3C2Tx, (b,c) Morphology of PAN/Ti3C2Tx NC, (d) Structure of PAN/Ti3C2Tx NC through XRD Analysis, (e,f) XPS of PANI and PAN/Ti3C2Tx NC for chemical element analysis. Reprinted with permission from ref. [52] Copyright 2020 Elsevier.
Nanomaterials 11 02496 g008
Figure 9. Comparative TGA analysis of MXP-NC with its precursors in (a) argon atmosphere, and (b) ambient Atmosphere. Reprinted with permission from ref. [65]. Copyright 2020 American Chemical Society.
Figure 9. Comparative TGA analysis of MXP-NC with its precursors in (a) argon atmosphere, and (b) ambient Atmosphere. Reprinted with permission from ref. [65]. Copyright 2020 American Chemical Society.
Nanomaterials 11 02496 g009
Figure 10. Hoping of charge carrier along 1D in conducting polymers due to addition of inorganic nanomaterials.
Figure 10. Hoping of charge carrier along 1D in conducting polymers due to addition of inorganic nanomaterials.
Nanomaterials 11 02496 g010
Figure 11. Schematic of ammonia sensing set-up. Reprinted with permission from ref. [69]. Copyright 2021 Wiley.
Figure 11. Schematic of ammonia sensing set-up. Reprinted with permission from ref. [69]. Copyright 2021 Wiley.
Nanomaterials 11 02496 g011
Figure 12. Chemisorption of ammonia in Ti3C2Tx/PEDOT: PSS sensors: Reprinted with permission ref. [65]. Copyright 2020 American Chemical Society.
Figure 12. Chemisorption of ammonia in Ti3C2Tx/PEDOT: PSS sensors: Reprinted with permission ref. [65]. Copyright 2020 American Chemical Society.
Nanomaterials 11 02496 g012
Figure 13. Ammonia-sensing mechanism in p–p-type MXP-NCs due to formation of Schottky barrier.
Figure 13. Ammonia-sensing mechanism in p–p-type MXP-NCs due to formation of Schottky barrier.
Nanomaterials 11 02496 g013
Figure 14. Ammonia-sensing mechanism in p–n-type MXP-NCs due to formation of p–n junctions.
Figure 14. Ammonia-sensing mechanism in p–n-type MXP-NCs due to formation of p–n junctions.
Nanomaterials 11 02496 g014
Figure 15. Flexibility of PEDOT: PSS/Ti3C2Tx NC ammonia sensor under different bending angles. Reprinted with permission from ref. [65]. Copyright 2020 American Chemical Society.
Figure 15. Flexibility of PEDOT: PSS/Ti3C2Tx NC ammonia sensor under different bending angles. Reprinted with permission from ref. [65]. Copyright 2020 American Chemical Society.
Nanomaterials 11 02496 g015
Figure 16. Ammonia-sensing in high humidity conditions. Reprinted with permission from ref. [56] Copyright 2021 Elsevier.
Figure 16. Ammonia-sensing in high humidity conditions. Reprinted with permission from ref. [56] Copyright 2021 Elsevier.
Nanomaterials 11 02496 g016
Figure 17. (a) Schematic of human breath analysis simulator, (b) monitoring of ammonia in exhaled human breath using MXP-NCs. Reprinted with permission from ref. [57]. Copyright 2021 Elsevi.er.
Figure 17. (a) Schematic of human breath analysis simulator, (b) monitoring of ammonia in exhaled human breath using MXP-NCs. Reprinted with permission from ref. [57]. Copyright 2021 Elsevi.er.
Nanomaterials 11 02496 g017
Figure 18. (a-c) Set-up to monitoring the volatilization of ammonia using MXP-NCs and (d) comparison of MXP-NCs performance in detecting volatilizing ammonia with results from DTM and SAAM techniques. Reprinted with permission from ref. [52]. Copyright 2020 Elsevier.
Figure 18. (a-c) Set-up to monitoring the volatilization of ammonia using MXP-NCs and (d) comparison of MXP-NCs performance in detecting volatilizing ammonia with results from DTM and SAAM techniques. Reprinted with permission from ref. [52]. Copyright 2020 Elsevier.
Nanomaterials 11 02496 g018
Figure 19. (A) Schematic of TENG coupled Nb2CTx/PAN ammonia sensor and (B) External load voltage as sensing parameter for (a,c) Polyaniline sensor, and (b,d) Nb2CTx/PAN sensor. Reprinted with permission from ref. [56]. Copyright 2021 Elsevier.
Figure 19. (A) Schematic of TENG coupled Nb2CTx/PAN ammonia sensor and (B) External load voltage as sensing parameter for (a,c) Polyaniline sensor, and (b,d) Nb2CTx/PAN sensor. Reprinted with permission from ref. [56]. Copyright 2021 Elsevier.
Nanomaterials 11 02496 g019
Figure 20. Technological and sustainable prospects of MXP-based ammonia sensor.
Figure 20. Technological and sustainable prospects of MXP-based ammonia sensor.
Nanomaterials 11 02496 g020
Table 1. Various types of NH3 sensors with respective advantages and disadvantages.
Table 1. Various types of NH3 sensors with respective advantages and disadvantages.
Type of SensorsAdvantagesDisadvantagesApplications
Electrochemical Sensors
[10,11,17]
Low detection limitShort livesTypical laboratory analysis
Insensitive to environment changeMalfunction of Electrodes
High and accurate sensitivityRequirement of special design for selective detectionWorkplaces such as chemical industries
Spectroscopic Sensors
[10,11,17]
Long lifetimeDifficulty in miniaturization to be installed at every emission siteRemote air quality monitoring
Expensive
Insensitive to environment changeComplicated designGas leak detection systems with accuracy and safety
High sensitivity, selectivity, and stabilityTime consuming
Calorimetric Sensors
[10,11,17]
Cost effectiveSelectivityMost combustible gases under industrial environment
Adequate sensitivity for industrial detectionLow versatility
Stable at ambient temperatureRisk of catalyst poisoning and explosionPetrochemical plants
Mass sensitive sensors
[10,11,17]
Long lifetimeLow sensitivityComponents of Wireless Sensor Networks.
Sensitive to environmental change
Avoiding secondary pollutionSelectivity
Sophisticated equipment
FET and Diode based Sensors
[10,11,17]
Adequate sensitivityFabrication complexityIndustrial applications and civil use
Cost effectiveMiniaturization
Detailed AnalysisCharacterization
Chemiresistors
[10,11,17,18]
Cost effectiveSensitiveness to surrounding environmentHumidity monitoring
Long lifetime
Eco-friendlyIndustrial applications
Energy efficient
Table 2. Comparison of Sensitivity of MXP-NCs with parent constituents and other materials toward NH3.
Table 2. Comparison of Sensitivity of MXP-NCs with parent constituents and other materials toward NH3.
Sensing MaterialSensing ParameterSensitivity @ Lowest Detection RangeTemperature of DetectionReference
Ti3C2Tx/PEDOT:PSSResistance4.94% @ 10 ppm27 °C[55]
Ti3C2Tx/PANResistance0.05% @ 25 ppbRT[39]
Nb2CTx/PAN-TENGVoltage9.33% @ 1 ppmRT[45]
Nb2CTx/PANResistance1.19% @ 20 ppb25 °C[46]
Ti3C2Tx/CPAMCurrent1.5% @ 50 ppmRT[51]
Nb2CTxVoltage8.15% @ 100 ppm25 °C[45]
PANResistance11% @ 1 ppmRT[77]
V2CTxResistance1.7% @100 ppmRT[78]
PEDOT:PSS: GrapheneResistance0.9% @ 5 ppmRT[79]
PAPResistance9% @ 10 ppm25 °C[2]
Alkalise-Ti3C2TxResistance11% @ 10 ppm25 °C[80]
PAN-AgResistance47.1% @ 1 ppm27 °C[58]
PAN-GOResistance21.8% @ 0.5 ppm20 °C[81]
PAN-MWCNTVoltage10% @ 0.01 ppmRT[82]
Table 3. Comparison of sensing characteristics of MXP-NCs and other NH3 sensors.
Table 3. Comparison of sensing characteristics of MXP-NCs and other NH3 sensors.
MXP-NCResponse TimeRecovery TimeLinear Regression ValueRepeatabilityReference
PEDOT:PSS/Ti3C2Tx116 s for 100 ppm40 s for 100 ppm0.957 for 10–100 ppm range3 cycles[55]
0.983 for 100–1000 ppm range
PAN/Ti3C2Tx~600 s for 25 ppb~1400 s for 25 ppb0.997 for 2–10 ppm4 cycles[39]
CPAM/Ti3C2Tx12.7 s for 150 ppm14.6 s for 150 ppm0–2000 ppm10 cycles[51]
Nb2CTx/PAN-TENG105 s for 100 ppm143 s for 100 ppm0.9655 for
1–100 ppm
3 cycles[45]
Nb2CTx/PAN218 s for 10 ppm300 for 10 ppm0.9951 for 0.1–10 ppm3 cycles[46]
V2CTx105 s for 100 ppm120 s for 100 ppmNot mentioned3 cycles[78]
PAN/MWCNT-TENG120 s for 100 ppm127 s for 100 ppm0.9928 for 20–100 ppm3 cycles[82]
Alkaised-Ti3C2Tx1 s for 100 ppm201 s for 100 ppmNot mentioned5 Cycles[80]
PAN-Ag271 s for 5 ppm575 s for 5 ppm1–100 ppm4 cycles[58]
Table 4. Summary of selectivity demonstrated by MXP-NCs-based NH3 sensors.
Table 4. Summary of selectivity demonstrated by MXP-NCs-based NH3 sensors.
AnalytesPEDOT:PSS/Ti3C2Tx
[55]
PAN/Ti3C2Tx
[39]
CPAM/Ti3C2Tx
[51]
Nb2CTx/PAN-TENG
[45]
Nb2CTx/PAN
[46]
Ammonia36.6%
@100 ppm
~1.7%
@10 ppm
4.7%
@200 ppm
~59%
@10 ppm
74.46%
@10 ppm
Toluene1.2%
@100 ppm
----
Methanol14%
@100 ppm
-~15%
@2000 ppm
--
Ethanol4.6%
@100 ppm
-~10%
@2000 ppm
~3%
@10 ppm
~2%
@10 ppm
Acetone3.4%
@100 ppm
-~10%
@2000 ppm
~2%
@10 ppm
~2%
@10 ppm
Sulfur dioxide-~0.02%
@25 ppm
-~6%
@10 ppm
~9%
@10 ppm
Hydrogen Sulfide-~1%
@25 ppm
-~2%
@10 ppm
~1%
@10 ppm
Formaldehyde-~0.02%
@25 ppm
-~0.5%
@10 ppm
~0.5%
@10 ppm
Carbon Monoxide-~0.05%
@25 ppm
-~1%
@10 ppm
Negligible
@10 ppm
Methane- -~2.5%
@10 ppm
Table 5. Flexibility feature demonstrated by MXP-NCs-based NH3 sensors.
Table 5. Flexibility feature demonstrated by MXP-NCs-based NH3 sensors.
MXP-NCBending Angle/Number of Folds
PEDOT: PSS/Ti3C2Tx over PI substrate: [55]
Bending Angle
(in degrees)
060120180240
Sensitivity @100 ppm36.4%~35–37%
PAN/Ti3C2Tx over PI substrate: [39]
Bending Angle
(in degrees)
0203040
Bending Cycles: (in number)0100300500
Sensitivity @10 ppm~1.6%
CPAM/Ti3C2Tx over PET substrate: [51]
Bending Angle
(in degrees)
0406080100
Bending Cycles: (in number)03
Sensitivity @2000 ppm~45%~43–45%
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Chaudhary, V.; Gautam, A.; Mishra, Y.K.; Kaushik, A. Emerging MXene–Polymer Hybrid Nanocomposites for High-Performance Ammonia Sensing and Monitoring. Nanomaterials 2021, 11, 2496. https://doi.org/10.3390/nano11102496

AMA Style

Chaudhary V, Gautam A, Mishra YK, Kaushik A. Emerging MXene–Polymer Hybrid Nanocomposites for High-Performance Ammonia Sensing and Monitoring. Nanomaterials. 2021; 11(10):2496. https://doi.org/10.3390/nano11102496

Chicago/Turabian Style

Chaudhary, Vishal, Akash Gautam, Yogendra K. Mishra, and Ajeet Kaushik. 2021. "Emerging MXene–Polymer Hybrid Nanocomposites for High-Performance Ammonia Sensing and Monitoring" Nanomaterials 11, no. 10: 2496. https://doi.org/10.3390/nano11102496

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