Next Article in Journal
Development and Characterization of Sustainable Biocomposites from Wood Fibers, Spent Coffee Grounds, and Ammonium Lignosulfonate
Previous Article in Journal
Physical Characterization of Multiwire Polystyrene Produced by Electrospinning Technique
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis and Studies of PAM-Ag-g/WS2/Ti3C2Tx Hydrogel and Its Possible Applications

1
Laboratory of Photovoltaic Phenomena and Devices, Institute of Physics and Technology, Satbayev University, Almaty 050013, Kazakhstan
2
Department of Chemistry, Nanjing Forestry University, Nanjing 210037, China
3
Department of Materials, National Center of Space Research and Technology JSC, Almaty 050010, Kazakhstan
*
Author to whom correspondence should be addressed.
Polymers 2025, 17(19), 2588; https://doi.org/10.3390/polym17192588
Submission received: 12 August 2025 / Revised: 15 September 2025 / Accepted: 19 September 2025 / Published: 24 September 2025
(This article belongs to the Section Polymer Composites and Nanocomposites)

Abstract

In this study, a new hybrid hydrogel based on PAM (polyacrylamide)-Ag-g/WS2/Ti3C2Tx was synthesized by radical polymerization using a conductive heterostructural nanocomposite WS2/Ti3C2Tx. The synergy between the polymer matrix and the interface between two-dimensional nanomaterials ensured the production of a hydrogel with high extensibility and conductivity, as well as sensory characteristics. The composite hydrogel exhibited excellent strain-sensing capabilities, with gauge factors of 1.4 at low strain and 2.8 at higher strain levels. In addition, the material showed a fast response time of 2.17 s and a short recovery time of 0.46 s under cyclic stretching, which confirms its high reliability and reproducibility. The integration of Ti3C2Tx and WS2 promoted the formation of a conductive network in the hydrogel structure, which simultaneously increased its mechanical strength and signal stability under variable loads. Measurements confirm some potential of the PAM-Ag-g/WS2/Ti3C2Tx composite hydrogel as a flexible wearable strain sensor. Based on measured numbers, we discussed the impact of the WS2/Ti3C2Tx interface on the Gauge factor and conductivity of the composite. Theoretical modeling demonstrates significant changes in the electronic structure of the WS2/Ti3C2Tx interface, and especially the WS2 surface, induced by substrate strain. Possible applications of the peculiar properties of PAM-Ag-g/WS2/Ti3C2Tx composite were proposed.

1. Introduction

Transition metal dichalcogenides (TMDs), which have excellent mechanical and electrical conductivity, are among the most promising 2D materials for the creation of a generation of smart and flexible electronic devices [1]. Among TMDs, tungsten disulfide WS2 is a notable candidate due to its large number of active sites and good electrical properties, and is also less toxic than graphene oxide and can therefore be used as a safer alternative in future applications [2]. The properties of TMDs can be improved by chemical modification. Various methods, such as cocatalyst loading, doping, and heterostructure design, have been used to improve the electrical conductivity of TMDs and expand their application [3,4,5]. WS2 TMDs possess out-of-plane sulfur atoms that readily interact with surrounding chemical groups, making it easy to form heterostructures with other 2D or 1D nanomaterials, determining the growth and morphology of heterostructured nanocomposites [6]. MXene, a 2D multilayer material, exhibits excellent conductivity and can serve as a good conductive substrate material for the uniform growth of 2D dichalcogenide sheets on top of and between layers. The development of MXene-based hybrid systems is in its infancy compared to other established 2D materials. MXene-based hybrids with interesting hierarchical structures and excellent performances have attracted considerable interest [7,8].
Conductive nanomaterial-filled hydrogels offer excellent conductivity, elasticity, ease of fabrication, and durability, making them ideal for flexible sensors [6]. Driven by Internet of Things (IoT) advancements, the market for flexible electronics like wearable sensors has expanded significantly. Conductive hydrogel-based sensors translate human mechanical motion (pressure, strain, axial displacement) into electrical signals for real-time monitoring [7]. These sensors, combined with AI, enable data analysis and predictive capabilities. However, the limited mechanical strength of traditional conductive hydrogels has hindered their broader use in flexible electronics. For instance, Jia Pan et al. [1] successfully integrated MoS2 into a flexible hydrogel-based sensor with a quick response time of 150 ms. Integrating WS2/Ti3C2Tx heterostructure into a hydrogel matrix is a promising approach for developing next-generation flexible wearable strain sensors. Mxene has exceptional metallic conductivity, hydrophilicity, and tunable surface end groups (–O, –OH, –F), which facilitate strong interfacial interaction with other nanomaterials and hydrogel networks. The resulting WS2/Ti3C2Tx heterointerface creates a 2D conductive network with an electric field, improving charge carrier mobility and sensor sensitivity. Recent reports demonstrate that interfacing WS2 with Ti3C2Tx can yield 2D-2D heterostructures with improved charge transport and interfacial contact and synergistic sensing/electrocatalytic behaviors [9]. However, most WS2/Ti3C2Tx studies focus on dry films/electrodes embedding a pre-formed WS2/Ti3C2Tx heterostructures nanofiller directly into hydrogel networks remains comparatively underexplored.
PAM-Ag-g hydrogel, a composite of synthetic and natural polymers, is promising for flexible sensors, biomedical devices [10], and wearable technology [11]. Agar-agar (Ag-g) provides structural integrity and prevents dehydration by forming a physical gel network that retains significant water [12]. This combination creates a dual network structure exhibiting high stretchability, elasticity, structural stability, and water retention capacity [13]. In this work, we aimed to synthesize and characterize a PAM-Ag-g/WS2/Ti3C2Tx hydrogel strain sensor with enhanced mechanical strength and sensitivity through interfacial engineering of a WS2/Ti3C2Tx heterostructure.

2. Materials and Methods

2.1. Materials

Titanium aluminum carbide 312 MAX Phase (Ti3AlC2, ≥90%, ≤40 μm particle size, Sigma-Aldrich, Merck KGaA, Darmstadt, Germany), tangsten hexachloride (WCl2, ≥99% Sigma-Aldrich, MilliporeSigma, Urbana, IL, USA), L-cysteine (C3H7NO2S, Sigma-Aldrich, Buchs, Switzerland), ethanol (C2H5OH, Sigma-Aldrich, Taufkirchen, Germany), phosphoric acid (≥89%, Sigma-Aldrich, Buchs, Switzerland), Acrylamide (AM, CH2=CHCONH2, 99%, Sigma-Aldrich, Buchs, Switzerland), N,N′-methylene bisacrylamide (MBA, 99%, Sigma-Aldrich, Wuxi, China), Ammonium persulphate (APS, 98.5%, Sigma-Aldrich, Milwaukee, WI, USA), Agar (Ag-g, (C12H18O9)n Sigma-Aldrich, Buchs, Switzerland).

2.2. Materials Characterization

The chemical structure of the samples was analyzed using Fourier transform infrared spectroscopy (FTIR, Nicolet iS10 FT-IR Spectrometer, Thermo Fisher Scientific Inc., Ogden, UT, USA). The surface chemical composition of MXene/WS2 was investigated via X-ray Photoelectron Spectroscopy (XPS, NEXSA Thermo Scientific). Scanning electron microscopy (SEM-EDS, JSM-7500F, JEOL Ltd., Yamagata, Japan) was used to study the morphology of sample surfaces. The crystal structures of samples were analyzed by XRD (SmartLab, Rigaku Co., Takatsuki, Japan, Cu Kα radiation, λ = 0.154056 nm). XRD data were obtained in the 2θ range from 20 to 80 °C at a scan rate of 6 deg. min−1 using X-ray radiation 40 kV, 30 mA. Raman spectroscopy (The Horiba LabRam Evolution, Palaiseau, France) was used to study the structure and phase composition of the hybrid WS2/Ti3C2Tx.

2.3. Synthesis

Synthesis of Ti3C2Tx MXene. Ti3C2Tx was prepared by selectively etching 1.0 g Ti3AlC2 powder was slowly added to 20 mL 48 wt% HF solution in a Teflon beaker under a fume hood with appropriate protective equipment. The mixture was stirred for 48 h at room temperature. The resulting mixture was then washed with deionized (DI) water until a neutral pH was reached. Finally, the Ti3C2 MXene was dried at 60 °C overnight in a vacuum oven.
Synthesis of WS2. During the synthesis, 0.099 g (0.01 M) of WCl6 and 0.061 g (0.02 M) L-cysteine were dissolved in 25 mL of distilled water under magnetic stirring at 400 rpm for 30 min at room temperature. The above solution was transferred to a 50 mL autoclave chamber. Then, the teflon-lined stainless-steel autoclave was heated at 180 °C for 24 h in a vacuum oven. The product was collected by centrifugation and repeated washing with ethanol and distilled water. The sample was dried at 60 °C overnight in a hot air oven. The powder was collected by centrifugation and repeated washing with ethanol and distilled water. The sample was dried at 600 °C overnight in a hot air oven.
Synthesis of WS2/Ti3C2Tx. The layered hybrid structure WS2/Ti3C2 was synthesized via a one-step hydrothermal method. A solution containing 0.005 M WCl6, 0.02 M cysteine, and Ti3C2 was added to distilled water and stirred for 30 min at room temperature. The resulting mixture was then heated at 180 °C for 24 h in a 50 mL Teflon-lined autoclave. The black precipitate was collected by centrifugation, washed three times with distilled water, and dried in a vacuum oven at 60 °C for 12 h. A schematic illustration is presented in Figure 1a.
Synthesis PAM-Ag-g/WS2/Ti3C2Tx Hydrogel. PAM-Ag-g/WS2/Ti3C2Tx hydrogel was synthesized via free radical polymerization. Acrylamide (1.5 g) and agar (400 mg) were dissolved in 10 mL of deionized water, followed by the addition of N,N′-methylenebisacrylamide (0.03 g) as a crosslinker and ammonium persulfate (0.03 g) as an initiator. After homogenization, WS2/Ti3C2Tx nanofillers (0.2–1 wt% relative to total polymer solids), pre-dispersed in 2 mL of deionized water by bath sonication for 30 min was introduced. The mixture was then polymerized in a mold at 60 °C for 12 h to form the hydrogel (Figure 1b). A schematic illustration is presented in Figure 1b.

2.4. Mechanical Testing of PAM-Ag-hg/WS2/Ti3C2Tx Hydrogel

The mechanical properties of the sample were evaluated using a tensile testing machine (Materials Testing Machine Z010/TN2S, Kennesaw, GA, USA), and the tests were performed at a tensile rate of 10 mm/min. For tensile testing, the hydrogels were made in the form of a rectangle (45 mm long, 20 mm high) and stretched at a strain rate of 5 mm min−1. The strength was calculated by integrating the area under the stress–strain curve.

2.5. Sensor Characteristic of PAM-Ag-g/WS2/Ti3C2Tx Hydrogel

The sensor characteristics of the hydrogel were evaluated using a custom-made platform with a linear motor (LinMot PS01-37x120F-HP-C, Dongguan, China) operating at 5–11 Hz in contact-opening mode to simulate compression and decompression. Each end of the hydrogel was inserted into a copper wire connected to a TH LCR meter, and the sensors were adjusted to the LCR meter (LCR-106X, IVYTECH, Changzhou, China) at different resistor deformations. The output resistance was measured using Kitstar software (version 5.2, Kitware Inc., Clifton Park, NY, USA) at different strains, and a force sensor (Vernier LabQuest Mini, Beaverton, OR, USA) monitored the applied strain.

2.6. First-Principles Modeling

The atomic structure and energetics of various configurations were studied using DFT with the QUANTUM-ESPRESSO code [14] and the GGA-PBE [15] functional, taking into account van der Waals forces corrections [16]. For all calculations, we used ultrasoft pseudopotentials [17]. The values of energy cutoffs of 35 Ry and 400 Ry for the plane-wave expansion of the wave functions and the charge density, respectively.
Figure 1. Schematic illustration of the synthesis process of (a) WS2/Ti3C2Tx and (b) PAM-Ag-g/WS2/Ti3C2Tx Hydrogel.
Figure 1. Schematic illustration of the synthesis process of (a) WS2/Ti3C2Tx and (b) PAM-Ag-g/WS2/Ti3C2Tx Hydrogel.
Polymers 17 02588 g001

3. Results and Discussion

3.1. Chemical Characteristics of WS2/Ti3C2Tx

The FTIR spectroscopy analysis (Figure 2a) allowed us to identify the functional groups in the synthesized PAM-Ag-g/WS2/Ti3C2Tx hydrogel composite, and to compare its spectral characteristics with those of pristine Ti3C2Tx, WS2 and WS2/Ti3C2Tx nanocomposite. The composite spectrum exhibits enhancement and broadening in the higher-wavenumber region 3200–3500 cm−1, indicating the –NH2 group of PAM [18], –OH group of Ti3C2Tx [19], and –OH group of agar-agar. The asymmetric stretching vibration of –CH2 for PAM is evident at ~3050 cm−1 [20]. A noticeable shift in this band, as well as shifts in other characteristics peaks, can be attributed to the strong interfacial interactions and hydrogen bonding between PAM chains and the WS2/Ti3C2Tx [18]. The peaks at 2990 cm−1 and 2890 cm−1 were apparently due to the –C–H stretching from residual cysteine originating from the WS2 synthesis. A slight shift the C=O stretching vibration ~1650 cm−1 and ~1070 cm−1 the C-O-C of agar-agar matrix [20]. The peak at ~1400 cm−1 reflects the O–H deformation vibrations, and the peak at ~1300 cm−1 reflects the C–F stretching vibrations [19] of Ti3C2Tx. The W-S stretching band near 600–700 cm−1 that confirms the presence od WS2 [21]. Taken together, these features indicate incorporation of the WS2/Ti3C2Tx hybrid filler into the hydrogel matrix and the formation of an effective interfacial contact without of degradation of the polymer network.
To study the phase composition and crystal structure, XRD analysis was performed (Figure 2b). The obtained diffraction data for the WS2/Ti3C2Tx composite were compared with the spectra of its individual components. For pure MXene, diffraction peaks were observed at 2θ ≈ 9.5°, 19.1°, 27.5°, 35.8° and 39.1°, corresponding to the (002), (004), (111), (200) and (110) planes of the Ti3C2Tx crystalline phase, indicating successful etching and exfoliation of the starting material. These data confirm the successful etching and exfoliation of MXene [22]. WS2 showed clear diffraction peaks at 2θ ≈ 14.2°, 28.3°, 32.8°, 44.4°, 58.4°, and 63.2°, indexed to the (002), (004), (001), (103), (008), and (112) planes of hexagonal WS2 (JCPDS Card No. 08-0237), indicating high crystallinity [23]. The WS2/Ti3C2Tx composite exhibited a superposition of the individual components’ diffraction peaks, without any additional peaks, indicating successful composite formation without structural degradation. Minor peak shifts and intensity variations (particularly in the 25–35° region) suggest interphase interactions or partial WS2 intercalation within the MXene layers. The X-ray diffraction results indicate the preservation of the crystalline structure of both components in the composite, which also points to the possible presence of synergistic interactions between the phases. The PAM-Ag-g/WS2/Ti3C2Tx hydrogel spectrum exhibits characteristic peaks of MXene and WS2, along with a broad amorphous peak at 2θ ≈ 23° from the PAM matrix [24], confirming the preserved crystallinity of the fillers and the polymer’s contribution.
Raman spectra (Figure 2c) showed characteristic vibrations for WS2. The out-of-plane mode A1g of sulfur at 458 cm−1 and the in-plane mode E1g associated with tungsten and sulfur atoms at 336 cm−1, which confirms the presence of the WS2 phase. The absence of distinct peaks for Ti3C2Tx phase could be attributed to the strong presence of Ti3C2Tx that reduces the overall crystallinity of the composite and leads to peak suppression in the Raman spectrum [25]. The presence of WS2/Ti3C2Tx2 in the sample is complemented by XRD and FTIR results.
XPS was used to investigate the surface chemical composition and electronic states of the synthesized samples. The spectra (Figure 2d) demonstrate the presence of characteristic peaks of Ti, C, and O elements in the structure of pure Ti3C2Tx. All spectra were calibrated using the C 1s peak at 284.8 eV. The spectrum of the WS2/Ti3C2Tx composite additionally contains peaks corresponding to W 4f and S 2p, which confirms the successful incorporation of WS2 into the Ti3C2Tx structure. The XPS data allowed us to clarify the elemental composition and types of chemical bonds in the WS2/Ti3C2Tx hybrid system. The Ti 2p spectrum contains peaks corresponding to Ti–O bonds at ~458.5 eV and ~464.3 eV, as well as S–Ti–C at ~455.1 eV, which indicate the existence of interphase interactions. The presence of the Ti–C peak confirms the preservation of the main MXene crystal structure. The C 1s spectrum contains signals corresponding to C–C, C–O/CHx, and C–Ti–Tx bonds, which are typical for functionalized Ti3C2Tx, corresponding to ~284.8 eV, 286.2 eV, and 288.5 eV, respectively. The O 1s spectrum shows signals indicating the presence of chemical bonds C–Ti–O, O–Ti–O, and C–Ti–OH at ~529.9 eV, ~531.2 eV, and ~532.4 eV, correspondingly, which indicates the presence of –O and –OH functional groups on the surface of the material.
Analysis of the W 4f region reveals characteristic peaks corresponding to W–S bonds (~32.5 eV for W 4f7/2 and ~34.7 eV for W 4f5/2), as well as the W 5p3/2 peak at ~38.6 eV, which confirms the presence of WS2 in the composite. The S 2p spectrum contains peaks associated with S–W at ~162.3 eV (S 2p3/2) and ~163.5 eV (S 2p1/2) [26], as well as a weak peak at about ~165.6 eV, which is associated with the formation of Si–T or S–Ti bonds and indicates interfacial interactions [27]. These data confirm the formation of the WS2/Ti3C2Tx hybrid structure with the preservation of the crystallinity of the original components and the presence of interphase interactions.
The morphology of the obtained Ti3C2Tx, WS2, and WS2/Ti3C2Tx structures was analyzed by SEM, which confirmed the successful growth of both Ti3C2Tx and WS2 layered structures (Figure 3a–d). Figure 3a of pure WS2 shows a flower-like morphology consisting of numerous individual flower shapes. Ti3C2Tx shows an accordion-like morphology, as shown in Figure 3b, where the Ti3C2Tx layers can be seen separated from each other. The cross-section of the WS2/Ti3C2Tx heterostructure sample is shown in Figure 3c,d, which depicts the uniform distribution and inclusion of WS2 nanoflowers within the interlayer spaces of the MXene layers, leading to the formation of wrinkled WS2/Ti3C2Tx heterostructure nanosheets.
The TEM images at different magnifications (Figure 4a–c) show that the flower-shaped WS2 nanostructures are uniformly distributed both on the surface and between the Ti3C2Tx layers. The measured interlayer distance in the heterostructure is 0.26 nm (Figure 4c), which corresponds to the Ti3C2Tx crystallographic plane [28].

3.2. Mechanical Properties of PAM-Ag-g/WS2/Ti3C2Tx Hydrogel

Reprehensive uniaxial tensile tests showed that the PAM-Ag-g/WS2/Ti3C2Tx hydrogel has improved mechanical properties compared to PAM-Ag-g (Figure 5a). This trend is consistent with hydrogen-bonding interactions formed upon adding Ti3C2Tx to the PAM-Ag-g matrix, with can enhance the hydrogels elasticity and ductility. Figure 5b shows that the composition retains high flexibility, strength, and resistance to destruction, which makes it promising for use in smart sensors, soft robotics, and biodegradable wearable devices.
The addition of the WS2/Ti3C2Tx heterostructured nanocomposite to the PAM-Ag-g hydrogel made it possible to obtain an electrically conductive composite [29]. To determine the optimal conductivity, the WS2/Ti3C2Tx content was varied in the range of 0–1 wt.% (0, 0.2, 0.4, 0.6, 0.8, 1 wt.%). Based on the results of four-probe measurements, it was found that the maximum electrical conductivity (1.02 S m−1) was achieved at a content of 0.4 wt.MXene/WS2 (Figure 6a) exhibits enhanced performance with a further increase in concentration. Note that a similar decrease in conductivity was observed for other MXenes-based composites [30]. Consequently, the PAM-Ag-g/WS2/Ti3C2Tx hydrogel with 0.4 wt% WS2/Ti3C2Tx was selected for further investigation.
The strain Gauge factor (GF) of the strain gauge is calculated using the equation:
G F = ( R R 0 ) ε
where ΔR = R − R0, R0 is the initial resistance, and R is the resistance at a certain deformation, respectively, ε is the corresponding deformation [29].
The dependence of the relative change in electrical resistance (ΔR/R0) on the applied strain for the PAM-Ag-g/WS2/Ti3C2Tx composite demonstrates a linear character with calculated values of the sensitivity coefficient (Figure 6b) equal to 1.4 in the region of low strains and 2.8 at high degrees of stretching. The obtained indicators indicate the sensitivity of the material to mechanical effects, which confirms its potential as a strain gauge sensor. However, the magnitude of the Gauge factor is far beyond what has been reported for MXene-based sensors [30].
As shown in Figure 6c, an apparent change in resistance is observed during five extension-relaxation cycles carried out at different strain levels (10%, 30%, 60%, 80%, 100%). The relative resistance (ΔR/R0) increases gradually with increasing strain, with each loading stage being reproduced with high accuracy. This indicates that the PAM-Ag-g/WS2/Ti3C2Tx -based hydrogel effectively responds to mechanical loads over a wide range and demonstrates stable and reliable operation even under multiple strain cycles. The time plots of ΔR/R0 (%) change confirm the stability of the sensory response and its repeatability during long-term use.
When stretched to 100%, the sensor element showed a response time of 2.17 s and a rapid recovery of 0.46 s (Figure 6d), indicating the sensitivity of the material to external mechanical influences and its ability to quickly return to its original characteristics after the load is removed.
The working mechanism of the PAM-Ag-g/WS2/Ti3C2Tx strain sensor can be explained by the redistribution of charge carriers at the WS2/Ti3C2Tx heterointerface, driven by electron transfer from the WS2 layers to the highly conductive MXene surface, which is accompanied by the formation of a built-in electric field. This process is additionally enhanced by the interfacial dipole–dipole interactions between the –O/–OH groups of MXene and the external sulfur atoms of WS2, leading to some reduction in interface resistance, an increase in the initial conductivity, and a substantial enhancement of the sensor’s sensitivity to mechanical deformation.

3.3. Theoretical Simulations

To unveil the nature of the relatively modest values of the gauge factor and conductivity, we performed the simulation of the changes in the electronic structure upon formation of WS2/Ti3C2Tx composite and following substrate-induced stretching. For this purpose, we built a model interface between the bilayer of WS2 and the Ti3C2(OH)2 monolayer (Figure 7a). The formation of the interface led to the transfer of the electron density from the MXene substrate to WS2. Thus, a decrease in the conductivity can be associated with the formation of the interface between highly conductive MXene and a semiconductor.
Next, we checked how the substrate-induced strain affects the electronic structure of the composite. Figure 7b,c depict MXene to WS2 charge transfer for 5% and 10% in-plane uniaxial strain. As one can see, in-plane stretching leads to an increase in electron transfer from MXene to WS2. This massive redistribution of the charge density leads to the appearance of a pseudo-gap on the Fermi level (see the area with zero energy in Figure 7d). This explains the relatively modest values of the Gauge factor observed in the experiment. Note that visible changes are observed even in the second layer of WS2. This doping led to the appearance of distinct states inside the band gap (about −1 eV and +0.5 eV in Figure 7e). These and other states can be the source of distinct optical transitions, which are shown by arrows in Figure 7e. This arising of the particular states on the edges of conductive and valence bands can also be the source of catalytic activity in diselenides as it was shown in our recent work [31]. The combination of mechanical stability, relatively good conductivity, and strain-induced doping can be the source of manipulating the optical, catalytic, and photo-catalytic properties of semiconductive dichalcogenides attached to MXenes incorporated in polymer matrices.

4. Conclusions

In conclusion, the WS2/Ti3C2Tx hybrid structure was successfully synthesized. The combination of the conductive properties of Ti3C2Tx and the catalytically active surface of WS2 allowed the formation of a stable conductive network in the polymer matrix, which ensured the efficient transmission of an electrical signal under mechanical action. The composite hydrogel demonstrated high stretchability, improved mechanical properties and reliability under cyclic loads, and high signal reproducibility, indicating the stability of the sensor function of the hydrogel. The transfer of electrons explains the mechanism of operation of the strain gauge, confirmed by interface modeling, from the WS2 layers to the highly conductive Ti3C2Tx surface, accompanied by the formation of a built-in electric field. Thus, PAM-Ag-g/WS2/Ti3C2Tx hydrogel is a promising material for creating flexible strain gauges applicable in various applications such as motion and chemical sensing, optics and catalysis.

Author Contributions

Conceptualization, A.A. and A.U.; methodology, Software, Writing—review and editing, D.W.B., A.S. and E.D.; software, A.K.; validation, E.B. and L.M.; formal analysis, A.A. and A.K.; investigation, A.U. and L.M.; resources, A.A.; data curation, A.S., A.K. and E.B.; writing—original draft preparation, A.A.; writing—review and editing, A.A. and E.D.; visualization, E.B. and A.S.; supervision, E.D.; project administration, E.D.; funding acquisition, E.D. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Science Committee of the Ministry of Science and Higher Education of the Republic of Kazakhstan (Grant No. BR21881954 “Development of technologies for the synthesis of nanostructured materials for efficient photocatalytic electrodes, photo- and gas-sensors”).

Data Availability Statement

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

Conflicts of Interest

Author Laura Mustafa was employed by the company National center of space research and technology JSC, Almaty, 050010, Kazakhstan. The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

References

  1. Pan, J.; Zhou, X.; Gong, H.; Lin, Z.; Xiang, H.; Liu, X.; Chen, X.; Li, H.; Liu, T.; Liu, S. Covalently Functionalized MoS2 Initiated Gelation of Hydrogels for Flexible Strain Sensing. ACS Appl. Mater. Interfaces 2023, 15, 36636–36646. [Google Scholar] [CrossRef] [PubMed]
  2. Wang, Q.H.; Kalantar-Zadeh, K.; Kis, A.; Coleman, J.N.; Strano, M.S. Electronics and Optoelectronics of Two-Dimensional Transition Metal Dichalcogenides. Nat. Nanotechnol. 2012, 7, 699–712. [Google Scholar] [CrossRef] [PubMed]
  3. Randall, J.N.; Luscombe, J.H.; Bate, R.T. Heterostructures and Quantum Devices; Elsevier: London, UK, 1994; Volume 24, ISBN 9780122341243. [Google Scholar]
  4. Feng, S.; Wang, X.; Wang, M.; Bai, C.; Cao, S.; Kong, D. Crumpled MXene Electrodes for Ultrastretchable and High-Area-Capacitance Supercapacitors. Nano Lett. 2021, 21, 7561–7568. [Google Scholar] [CrossRef] [PubMed]
  5. Wu, J.; Li, Q.; Shuck, C.E.; Maleski, K.; Alshareef, H.N.; Zhou, J.; Gogotsi, Y.; Huang, L. An Aqueous 2.1 V Pseudocapacitor with MXene and V–MnO2 Electrodes. Nano Res. 2022, 15, 535–541. [Google Scholar] [CrossRef]
  6. Zhu, M.; Du, X.; Liu, S.; Li, J.; Wang, Z.; Ono, T. A Review of Strain Sensors Based on Two-Dimensional Molybdenum Disulfide. J. Mater. Chem. C 2021, 9, 9083–9101. [Google Scholar] [CrossRef]
  7. Wang, J.; Liu, Y.; Wang, S.; Liu, X.; Chen, Y.; Qi, P.; Liu, X. Molybdenum Disulfide Enhanced Polyacrylamide-Acrylic Acid-Fe3+ Ionic Conductive Hydrogel with High Mechanical Properties and Anti-Fatigue Abilities as Strain Sensors. Colloids Surf. A 2021, 610, 125692. [Google Scholar] [CrossRef]
  8. Zhang, J.; Su, E.; Li, C.; Xu, S.; Tang, W.; Cao, L.N.Y.; Li, D.; Wang, Z.L. Enhancing Artifact Protection in Smart Transportation Monitoring Systems via a Porous Structural Triboelectric Nanogenerator. Electronics 2023, 12, 3031. [Google Scholar] [CrossRef]
  9. Rasool, F.; Pirzada, B.M.; Talib, S.H.; Alkhidir, T.; Anjum, D.H.; Mohamed, S.; Qurashi, A. In Situ Growth of Interfacially Nanoengineered 2D-2D WS2/Ti3C2Tx MXene for the Enhanced Performance of Hydrogen Evolution Reactions. ACS Appl. Mater. Interfaces 2024, 16, 14229–14242. [Google Scholar] [CrossRef]
  10. Han, F.; Chen, S.; Wang, F.; Liu, M.; Li, J.; Liu, H.; Yang, Y.; Zhang, H.; Liu, D.; He, R.; et al. High-Conductivity, Self-Healing, and Adhesive Ionic Hydrogels for Health Monitoring and Human-Machine Interactions Under Extreme Cold Conditions. Adv. Sci. 2025, 12, 2412726. [Google Scholar] [CrossRef]
  11. Lei, H.; Zhao, J.; Ma, X.; Li, H.; Fan, D. Antibacterial Dual Network Hydrogels for Sensing and Human Health Monitoring. Adv. Healthcare Mater. 2021, 10, 2101089. [Google Scholar] [CrossRef]
  12. Hou, W.; Sheng, N.; Zhang, X.; Luan, Z.; Qi, P.; Lin, M.; Tan, Y.; Xia, Y.; Li, Y.; Su, K. Design of Injectable Agar/NaCl/Polyacrylamide Ionic Hydrogels for High Performance Strain Sensors. Carbohydr. Polym. 2019, 211, 322–328. [Google Scholar] [CrossRef]
  13. Wang, Y.; Chen, F.; Liu, Z.; Tang, Z.; Yang, Q.; Zhao, Y.; Du, S.; Chen, Q.; Zhi, C. A Highly Elastic and Reversibly Stretchable All-Polymer Supercapacitor. Angew. Chem. Int. Ed. 2019, 58, 15707–15711. [Google Scholar] [CrossRef]
  14. Giannozzi, P.; Baroni, S.; Bonini, N.; Calandra, M.; Car, R.; Cavazzoni, C.; Ceresoli, D.; Chiarotti, G.L.; Cococcioni, M.; Dabo, I.; et al. Quantum Espresso: A modular and open-source software project for quantum simulations of materials. J. Phys. Condens. Matter 2009, 21, 395502. [Google Scholar] [CrossRef] [PubMed]
  15. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 1996, 77, 3865. [Google Scholar] [CrossRef] [PubMed]
  16. Barone, V.; Casarin, M.; Forrer, D.; Pavone, M.; Sambi, M.; Vittadini, A. Role and effective treatment of dispersive forces in materials: Polyethylene and graphite crystals as test cases. J. Comput. Chem. 2009, 30, 934–939. [Google Scholar] [CrossRef] [PubMed]
  17. Vanderbilt, D. Soft self-consistent pseudopotentials in a generalized eigenvalue formalism. Phys. Rev. B 1990, 41, 7892. [Google Scholar] [CrossRef]
  18. 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, 200048. [Google Scholar] [CrossRef]
  19. Chen, X.; Zhao, Y.; Li, L.; Wang, Y.; Wang, J.; Xiong, J.; Yu, J. MXene/Polymer Nanocomposites: Preparation, Properties, and Applications. Polym. Rev. 2020, 61, 80–115. [Google Scholar] [CrossRef]
  20. Wahba, M.I.; Hassan, M.E. Agar-Carrageenan Hydrogel Blend as a Carrier for the Covalent Immobilization of β-D-Galactosidase. Macromol. Res. 2017, 25, 913–923. [Google Scholar] [CrossRef]
  21. Vattikuti, V.S.V.; Chan, B. Effect of CTAB Surfactant on Textural, Structural, and Photocatalytic Properties of Mesoporous WS2. Sci. Adv. Mater. 2015, 7, 2639–2645. [Google Scholar] [CrossRef]
  22. Yang, J.; Voiry, D.; Ahn, S.J.; Kang, D.; Kim, A.Y.; Chhowalla, M.; Shin, H.S. Two-dimensional hybrid nanosheets of tungsten disulfide and reduced graphene oxide as catalysts for enhanced hydrogen evolution. Angew. Chem. Int. Ed. 2013, 125, 13996–13999. [Google Scholar] [CrossRef]
  23. Chen, R.; Cheng, Y.; Wang, P.; Wang, Y.; Wang, Q.; Yang, Z.; Tang, C.; Xiang, S.; Luo, S.; Huang, S.; et al. Facile synthesis of a sandwiched Ti3C2Tx MXene/nZVI/fungal hypha nanofiber hybrid membrane for enhanced removal of Be(II) from Be(NH2)2 complexing solutions. Chem. Eng. J. 2021, 421, 129682. [Google Scholar] [CrossRef]
  24. Miah, M.Y.; Halder, S.; Saikat, S.P.; Dewanjee, S.; Ashaduzzaman, M.; Bhowmik, S. Microwave-assisted one-step synthesis of polyacrylamide/NiO nanocomposite for biomedical applications. RSC Adv. 2025, 15, 18971–18985. [Google Scholar] [CrossRef] [PubMed]
  25. Wang, Q.; Liu, A.; Qiao, S.; Zhang, Q.; Huang, C.; Lei, D.; Shi, X.; He, G.; Zhang, F. Mott-Schottky MXene@WS2 Heterostructure: Structural and Thermodynamic Insights and Application in Ultra Stable Lithium−Sulfur Batteries. ChemSusChem 2023, 16, e202300507. [Google Scholar] [CrossRef]
  26. Zhang, B.; Luo, C.; Deng, Y.; Huang, Z.; Zhou, G.; Lv, W.; He, Y.B.; Wan, Y.; Kang, F.; Yang, Q.H. Optimized Catalytic WS2–WO3 Heterostructure Design for Accelerated Polysulfide Conversion in Lithium–Sulfur Batteries. Adv. Energy Mater. 2020, 10, 2000091. [Google Scholar] [CrossRef]
  27. Yuan, Z.; Wang, L.; Li, D.; Cao, J.; Han, W. Carbon-Reinforced Nb2CTx MXene/MoS2 Nanosheets as a Superior Rate and High-Capacity Anode for Sodium-Ion Batteries. ACS Nano 2021, 15, 7439–7450. [Google Scholar] [CrossRef]
  28. Wang, X.; Luo, D.; Wang, J.; Sun, Z.; Cui, G.; Chen, Y.; Wang, T.; Zheng, L.; Zhao, Y.; Shui, L.; et al. Recent Advances in Layered Transition Metal Dichalcogenides for Hydrogen Evolution Reaction. Angew. Chem. Int. Ed. 2021, 60, 2371–2378. [Google Scholar] [CrossRef]
  29. Biswas, M.C.; Chakraborty, S.; Bhattacharjee, A.; Mohammed, Z. 4D Printing of Shape Memory Materials for Textiles: Mechanism, Mathematical Modeling, and Challenges. Adv. Funct. Mater. 2021, 31, 2100257. [Google Scholar] [CrossRef]
  30. Leong, W.X.R.; Al-Dhahebi, A.M.; Ahmad, M.R.; Saheed, M.S.M. Ti3C2Tx MXene-Polymeric Strain Sensor with Huge Gauge Factor for Body Movement Detection. Micromachines 2022, 13, 1302. [Google Scholar] [CrossRef]
  31. Boukhvalov, D.W.; Chuchvaga, N.A.; Rakhimzhanov, M.K.; Shongalova, A.; Serikkanov, A.S.; Osipov, V.Y. The Effect of the Metal Impurities on the Stability, Chemical, and Sensing Properties of MoSe2. Surfaces 2025, 8, 56. [Google Scholar] [CrossRef]
Figure 2. (a) FTIR spectra of pure WS2, Ti3C2Tx, and hybrid structure WS2/Ti3C2Tx; (b) XRD patterns of WS2, Mxene and, hybrid structure WS2/Ti3C2Tx; (c) Raman spectra of WS2 and WS2/Ti3C2Tx; (d) XPS survey spectrum; (e) Ti 2p; (f) XPS spectra of C 1s; (g) XPS spectra of O 1s; (h) XPS Spectra of W 4f; (i) XPS Spectra of S 2p.
Figure 2. (a) FTIR spectra of pure WS2, Ti3C2Tx, and hybrid structure WS2/Ti3C2Tx; (b) XRD patterns of WS2, Mxene and, hybrid structure WS2/Ti3C2Tx; (c) Raman spectra of WS2 and WS2/Ti3C2Tx; (d) XPS survey spectrum; (e) Ti 2p; (f) XPS spectra of C 1s; (g) XPS spectra of O 1s; (h) XPS Spectra of W 4f; (i) XPS Spectra of S 2p.
Polymers 17 02588 g002
Figure 3. SEM images of: (a) WS2 nanoflowers, (b) Ti3C2Tx, (c,d) WS2/Ti3C2Tx nanocomposite.
Figure 3. SEM images of: (a) WS2 nanoflowers, (b) Ti3C2Tx, (c,d) WS2/Ti3C2Tx nanocomposite.
Polymers 17 02588 g003
Figure 4. TEM image of WS2/Ti3C2Tx nanocomposite at different magnifications: (a) 500 nm, (b) 100 nm, (c) 10 nm.
Figure 4. TEM image of WS2/Ti3C2Tx nanocomposite at different magnifications: (a) 500 nm, (b) 100 nm, (c) 10 nm.
Polymers 17 02588 g004
Figure 5. (a) Stress–strain curves of PAM-Ag-g и PAM-Ag-g/WS2/Ti3C2Tx; (b) Visual demonstration of the strength of PAM-Ag-g/WS2/Ti3C2Tx.
Figure 5. (a) Stress–strain curves of PAM-Ag-g и PAM-Ag-g/WS2/Ti3C2Tx; (b) Visual demonstration of the strength of PAM-Ag-g/WS2/Ti3C2Tx.
Polymers 17 02588 g005
Figure 6. Electromechanical performances of the PAM-Ag-g/WS2/Ti3C2Tx hydrogel: (a) Conductivity of PAM-Ag-g/WS2/Ti3C2Tx with different MXene/WS contents; (b) Gauge factor of PAM-Ag-g/WS2/Ti3C2Tx; (c) Variation in ΔR/R0 of the PAM-Ag-g/WS2/Ti3C2Tx at 10%, 30%, 60%, 80%, and 100%; (d) The response time and recovery time of the hydrogel sensor.
Figure 6. Electromechanical performances of the PAM-Ag-g/WS2/Ti3C2Tx hydrogel: (a) Conductivity of PAM-Ag-g/WS2/Ti3C2Tx with different MXene/WS contents; (b) Gauge factor of PAM-Ag-g/WS2/Ti3C2Tx; (c) Variation in ΔR/R0 of the PAM-Ag-g/WS2/Ti3C2Tx at 10%, 30%, 60%, 80%, and 100%; (d) The response time and recovery time of the hydrogel sensor.
Polymers 17 02588 g006
Figure 7. The changes in charge density distribution between the WS2 bilayer and the MXene substrate for no applied strain (a), and 5% (b) and (10%) of in-plane uniaxial stretching. The yellow and cyan “clouds” correspond with an increase and a decrease in eletron density, respectively. Tungsten atoms are shown in black, sulfur in light brown, oxygen in red, titanium in gray, carbon in dark brown, and hydrogen in pale pink. The isosurface level is the same on three panels (ac). Total density of states of WS2/Ti3C2Tx interface (d) and W 5d partial (e) densities of states before and after applying in-plane stretching. Fermi energy set to zero. The blue and orange arrows in panel (e) indicate transitions corresponding to adsorption and emission of light, respectively.
Figure 7. The changes in charge density distribution between the WS2 bilayer and the MXene substrate for no applied strain (a), and 5% (b) and (10%) of in-plane uniaxial stretching. The yellow and cyan “clouds” correspond with an increase and a decrease in eletron density, respectively. Tungsten atoms are shown in black, sulfur in light brown, oxygen in red, titanium in gray, carbon in dark brown, and hydrogen in pale pink. The isosurface level is the same on three panels (ac). Total density of states of WS2/Ti3C2Tx interface (d) and W 5d partial (e) densities of states before and after applying in-plane stretching. Fermi energy set to zero. The blue and orange arrows in panel (e) indicate transitions corresponding to adsorption and emission of light, respectively.
Polymers 17 02588 g007
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Arinova, A.; Boukhvalov, D.W.; Umirzakov, A.; Bondar, E.; Shongalova, A.; Mustafa, L.; Kemelbekova, A.; Dmitriyeva, E. Synthesis and Studies of PAM-Ag-g/WS2/Ti3C2Tx Hydrogel and Its Possible Applications. Polymers 2025, 17, 2588. https://doi.org/10.3390/polym17192588

AMA Style

Arinova A, Boukhvalov DW, Umirzakov A, Bondar E, Shongalova A, Mustafa L, Kemelbekova A, Dmitriyeva E. Synthesis and Studies of PAM-Ag-g/WS2/Ti3C2Tx Hydrogel and Its Possible Applications. Polymers. 2025; 17(19):2588. https://doi.org/10.3390/polym17192588

Chicago/Turabian Style

Arinova, Anar, Danil W. Boukhvalov, Arman Umirzakov, Ekaterina Bondar, Aigul Shongalova, Laura Mustafa, Ainagul Kemelbekova, and Elena Dmitriyeva. 2025. "Synthesis and Studies of PAM-Ag-g/WS2/Ti3C2Tx Hydrogel and Its Possible Applications" Polymers 17, no. 19: 2588. https://doi.org/10.3390/polym17192588

APA Style

Arinova, A., Boukhvalov, D. W., Umirzakov, A., Bondar, E., Shongalova, A., Mustafa, L., Kemelbekova, A., & Dmitriyeva, E. (2025). Synthesis and Studies of PAM-Ag-g/WS2/Ti3C2Tx Hydrogel and Its Possible Applications. Polymers, 17(19), 2588. https://doi.org/10.3390/polym17192588

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