Next Article in Journal
Einstein Aggregation Operator Technique in Circular Fermatean Fuzzy Environment for MCDM
Previous Article in Journal
Assessment of Measured Mixing Time in a Water Model of Eccentric Gas-Stirred Ladle with a Low Gas Flow Rate: Tendency of Salt Solution Tracer Dispersions
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Novel Flexible Pressure Sensor with Symmetrical Structure Based on 2-D MoS2 Material on Polydimethylsiloxane Substrate

School of Integrated Circuits, Jiangnan University, Wuxi 214122, China
*
Author to whom correspondence should be addressed.
Symmetry 2024, 16(9), 1242; https://doi.org/10.3390/sym16091242
Submission received: 14 August 2024 / Revised: 14 September 2024 / Accepted: 18 September 2024 / Published: 21 September 2024
(This article belongs to the Section Engineering and Materials)

Abstract

:
Flexible pressure sensors can be widely utilized in healthcare, human–computer interaction, and the Internet of Things (IoT). There is an increasing demand for high-precision and high-sensitivity flexible pressure sensors. In response to this demand, a novel flexible pressure sensor with a symmetrical structure composed of MoS2 and PDMS is designed in this paper. Simulation is conducted on the designed flexible pressure sensor. Its piezoresistive effect is analyzed, and the influence of the cavity structure on its sensitivity is investigated. Additionally, a fully symmetrical Wheatstone bridge composed of the flexible pressure sensor is designed and simulated. Its symmetrical structure improves the temperature stability and the sensitivity of the sensor. The structure can be used to convert pressure changes into voltage changes conveniently. It indicates that the sensor achieves a sensitivity of 1.13 kPa−1 in the micro-pressure range of 0–20 kPa, with an output voltage sensitivity of 3.729 V/kPa. The designed flexible pressure sensor exhibits promising potential for applications in wearable devices and related fields, owing to its high sensitivity and precision.

Graphical Abstract

1. Introduction

Since various smart devices are applied deeply in our daily lives, various types of sensors for signal acquisition are highly required. A pressure sensor is an important de-device for front-end signal input with the merits of a simple structure, good stability, and high precision [1,2].
Unlike traditional pressure sensors based on rigid materials, novel flexible pressure sensors can be conformed to various surfaces owing to the applied flexible materials for both conductive components and substrates. This characteristic enhances the convenience of the testing process and provides an advantage in measuring pressure on non-planar surfaces. Moreover, flexible pressure sensors find extensive applications across diverse fields due to their unique properties, including ultra-thinness, a low modulus, light weight, high sensitivity, and scalability. The reported applications include wearable electronics [1], electronic skin [3], and human–computer interaction [2], among others.
During the development process of flexible pressure sensors, the main performance of the sensors has been continuously improved in terms of sensitivity, response time, relaxation time, stability, etc.
In 2014, Zhu et al. [4] introduced a flexible resistive tactile sensor utilizing microstructured graphene arrays. The sensor exhibits a high sensitivity of −5.5 kPa−1 in the low-pressure range (<100 Pa), an ultra-low detection limit of 1.5 Pa, and an ultra-fast response time (0.2 ms). Importantly, the sensitivity of the flexible sensor can be adjusted by selecting microstructures with different parameters to suit various sensing purposes.
In 2017, Kim et al. [5] presented a fabrication process for VACNT/PDMS composite materials, leading to a flexible resistive pressure sensor composed of a VACNT/PDMS composite conductor with irregular surface morphology. The sensor shows high sensitivity (~0.3 kPa−1 to 0.7 kPa−1), a stable steady-state response across various pressures (20 Pa to 5 kPa), and relatively fast response (162 ms), and relaxation times (~108 ms), along with mechanical durability (5000 cycles) and reversible on/off behavior.
In 2017, Luo et al. developed a material processing strategy to produce hollow structured graphene-PDMS composites, ensuring an excellent piezoresistive response. This unique structure enables independent adjustment of the resistance and mechanical modulus, thereby varying the sensitivity and linear range, with the highest sensitivity reaching 15.9 kPa−1 in the range of 0–60 kPa [6].
In 2018, Ma et al. prepared MX/rGO hybrid aerogels using ice-template freezing technology and applied them in piezoresistive sensors. The 3D aerogels demonstrate superior sensor performance due to the synergistic effect between MXene and rGO, featuring high sensitivity (22.56 kPa−1), a rapid response (<200 ms), and excellent stability over more than 10,000 cycles [7].
In 2018, Zhang, YP, et al. designed a MXene/black phosphorus (BP)-based self-powered smart sensor system. With MXene/BP as the sensitive layer in a flexible pressure sensor, the pressure sensitivity of the device can be improved to 77.61 kPa−1 at an optimized elastic modulus of 0.45 MPa, with a fast response time of 10.9 ms [8].
There are numerous options for the conductive components of flexible sensors. Carbon materials (including graphene, GO, rGO, and carbon nanotubes (CNT)), metal nanomaterials, and flexible organic-based piezoelectric materials by supramolecular packing, bestowing macroscopic dipole-driven electro-mechanical and optical functions [9,10], and conductive polymers are all viable choices. Carbon-based polymer composites, in particular, have gained significant attention in research due to their superior electrical and mechanical properties, as well as their unique two-dimensional (2-D) structure [11]. However, challenges such as high hysteresis and poor repeatability still exist [12]. The strong van der Waals interactions between graphene sheets and the high junction contact significantly inhibit the high conductivity and mechanical strength of graphene sheets [11]. In this study, MoS2, which is a two-dimensional transition metal dichalcogenide (TMDC) material, is selected as the flexible pressure-sensitive material.
Bulk MoS2 exhibits a hexagonal layered structure with weak van der Waals forces between the adjacent layers. MoS2 is in a “sandwich” structure, comprising two layers of sulfur atoms enclosing a central layer of molybdenum atoms [13]. Each molybdenum atom is surrounded by six sulfur atoms, forming strong covalent bonds, while each sulfur atom is bonded to three molybdenum atoms. 2D MoS2 can offer several advantages. Firstly, its 2-D electron confinement, especially in monolayer form, makes it an ideal channel material for high-performance electronic and optoelectronic devices, resulting in competitive electron mobility. Secondly, its strong in-plane covalent bonds and atomic thickness contribute to excellent mechanical strength (fracture strain > 2.2%), flexibility, and optical transparency [14]. Compared to graphene, MoS2 possesses a wider band gap of 1.8 eV, chemical inertness, and superior temperature stability. Lastly, MoS2 exhibits stable operation and excellent sensitivity due to its tunable band gap and high gauge factor [12]. These attributes make MoS2 particularly suitable for flexible piezoresistive sensors.
In 2018, Kim et al. reported on a cracked paddy-shaped MoS2/graphene porous network (GPN) infiltrated into an Ecoflex hybrid nanostructure for a wearable strain pressure sensor. This MoS2/GPN/Ecoflex sensor demonstrates stable characteristics, maintaining device performance even after 4000 cycles of repeated pressing and release. Additionally, it exhibits a high gauge factor (GF) of 24.1 and excellent elasticity in bending and tensile tests [11].
In 2021, Xing Pang et al. designed and fabricated a flexible piezoresistive pressure sensor combining molybdenum disulfide with PDMS. The sensor, with 1510 μm thick PDMS film, shows high sensitivity up to 22.62 MPa−1 in the pressure range of 0–0.23 MPa and even 866.9 MPa−1 above 0.4 MPa pressure [15]. It was successfully employed as a wearable pressure sensor for measuring plantar pressure, demonstrating good repeatability.
In 2022, X.Y. Chen et al. proposed a soaking method to fabricate an MoS2/HEC/PU sponge pressure sensor. The porous structure imparted excellent performance to the sensor, including high sensitivity (0.746 kPa−1 in the 50–250 kPa pressure range), a wide pressure detection range (up to 250 kPa), a fast response/release time (120 ms), and excellent repeatability over 2000 cycles [16].
In 2022, D.D. Xu et al. introduced a novel flexible ionic electronic pressure sensor based on MoS2 material. The sensor exhibits ultra-high sensitivity (S = 89.75 kPa−1) and a wide sensing range (722.2 kPa). Additionally, it shows rapid response and relaxation times (<3 ms), indicating its quick response speed to external pressure stimuli. Long-term cycling stability is also demonstrated, enduring over 5000 cycles without obvious degradation under pressure of 138.9 kPa [17].
In this paper, the integration of a fully symmetrical Wheatstone bridge structure with a flexible piezoresistive pressure sensor based on MoS2 and PDMS substrate is proposed, showing superior sensitivity. The paper is organized as follows: In the first section, the research background and the significance of the PDMS + MoS2 flexible sensor are introduced, along with an overview of current research trends. In the second part, the piezoresistive effect of the designed flexible sensor is analyzed. The sensor’s manufacturing process is reported. In the third section, the piezoresistive effect of the MoS2 film on PDMS is simulated. The impact of the cavity structure on the sensor’s sensitivity is investigated. A circuit-level simulation model of the fully symmetrical Wheatstone bridge structure is developed to verify the zero offset effect resulting from the resistance asymmetry caused by process deviations during preparation. In the fourth part, the conclusions are summarized.

2. Design of a Novel Flexible Pressure Sensor Structure

2.1. Working Principle of a Piezoresistive Pressure Sensor

The operation principle of the piezoresistive pressure sensor is shown in Figure 1. With changing pressure, the resistance and the output voltage can be changed.
The value of the resistance is:
R = ρ L A
It can be seen that the resistance is influenced by the resistivity (ρ), the contact area (A), and the electrical path length (L). When the pressure is applied, the pressure-sensitive membrane undergoes deformation, altering the values of ρ, L, and A, thus changing the resistance (R) and resulting in a corresponding change in voltage (V). The voltage change can be used to quantify the stress variation.
Figure 1a shows the diagram with rigid material, while Figure 1b shows the situation with flexible material. As shown in Figure 1a, as the applied pressure is increased, if the change in resistivity ρ (piezoresistive effect) is not considered, the electrical path length L should be changed accordingly. On the other hand, as shown in Figure 1b, with the flexible material, in addition to the resulting change in L, the contact area with the negative electrode is also increased due to squeezing the material. The combined changes caused by the pressure on the flexible material could result in an obvious improved change in the resistance. In this way, the sensitivity of the sensor based on the flexible material is improved. Based on this idea, various microstructures are proposed to enhance the sensitivity by augmenting the contact area or conduction paths upon pressure application.
The related change of the resistance can be expressed with the following equation:
Δ R R = ( 1 + 2 v ) + Δ ρ ρ
The term (1 + 2ν) denotes the impact of geometric alterations, while the latter signifies the piezoresistive effect intrinsic to the material itself. With the applied metals and metal alloys, the resistance is typically increased by several times. However, the semiconductor materials, such as germanium (Ge) and silicon (Si), exhibit a significant increase in resistance due to changes in their band gap or atomic spacing, often achieving enhanced performance through a substantial coefficient [18].
Figure 2 shows the cross-sectional schematic of the piezoresistive pressure sensor. When pressure is applied, both the piezoresistive film and the piezoresistors are deformed, leading to varying stresses across different locations of the piezoresistors, thereby causing changes in resistance. These resistance changes can be converted into voltage outputs through the Wheatstone bridge structure [19].
The piezoresistive effect indicates the alteration in energy bands within a semiconductor material with the subjected stress, resulting in shifts in energy valleys and subsequent changes in resistivity. The magnitude of this effect is quantified by the piezoresistive coefficient π, defined as the relative change in resistivity per unit stress.
Δ R R = π P
The sensitivity denotes the extent of resistance change in response to variations in the pressure. Higher sensitivity signifies greater measurement accuracy, indicating that the sensor can detect smaller changes in pressure with higher precision.
S = Δ R R / Δ P
Figure 3 shows the schematic diagram of the Wheatstone bridge structure based on the flexible pressure sensors. In the figure, Vdd represents the power supply voltage applied to the bridge, while R1, R2, R3, and R4 denote the resistors positioned on the four arms of the bridge. The output voltage Vout is the measured voltage, which is used to calculate the strain value. If Vout is equal to zero, it means there is no pressure applied to any part of the bridge. With the applied pressure on the resistance, Vout can be expressed as the following equation:
V O U T = ( R 2 R 1 + R 2 R 4 R 3 + R 4 ) V d d
The Wheatstone bridge structure is employed for the symmetric connection, serving to not only mitigate the zero temperature drift of the sensor, but also to convert the challenging-to-measure variations of the varistor into easily detectable current or voltage outputs. This configuration facilitates convenient integration with subsequent detection circuits.

2.2. Structural Design

In the paper, 2D-MoS2 is adopted as the piezoresistive material, while PDMS is the substrate. A symmetrical Wheatstone bridge structure is adopted to construct a flexible piezoresistive pressure sensor.
Polydimethylsiloxane (PDMS) is widely utilized in the fabrication of electronic skin and other stretchable electronic devices due to its commercial availability and well-established properties. Its advantages include chemical inertness, stability across a broad temperature range, transparency, tunable mechanical properties, and the capability to define adhesive and non-adhesive regions through UV irradiation, which is crucial for bonding electronic materials to its surface [20].
To fabricate the sensor, firstly, PDMS is cast onto a mold to create a cavity structure. ITO/Al electrodes are deposited onto the PDMS surface. A monolayer of MoS2 is grown on another Si/SiO2 surface and subsequently transferred to the PDMS surface using a PLLA rapid transfer method. The MoS2 layer is then photo-lithographically etched using Cl radicals and Ar+ ion beams before being covered with another layer of PDMS.
The schematic of the fabricated flexible pressure sensor is shown in Figure 4a, including the upper PDMS, lower PDMS, MoS2, and electrodes. The facilities for the measurement of the flexible pressure sensor are shown in Figure 4b. The measured flexible pressure sensor is placed in a test chamber (KOMEG KMT-64-S). A Druck PACE5000 gas pressure controller is used to control the pressure on the sensor. A digital multimeter (DMM) UT60B is used to measure the output of the Wheatstone structure.

2.3. Manufacturing Process

Figure 5 shows the soft lithography process for the PDMS cavity. The silanization treatment is shown in Figure 5a. The mold is placed in a culture dish, and a few drops of TMCS (trimethylchlorosilane) are applied around the mold. The dish is sealed for 15 min. Afterward, the dish is opened, allowing an additional 10 min for all the TMCS to evaporate. The contact angle of the treated mold is measured. The PDMS is mixed with a curing agent with a ratio of 10:1. The bubbles in the PDMS mixture are removed by initiating a vacuum and maintaining it for 30 min. Then, the mixture is continuously baked. The degassed PDMS is poured onto the silanized mold, as shown in Figure 5b. The PDMS is baked at 80 °C for 2 h. It can be demolded after the cooling process (as shown in Figure 5c).
The production process follows the method for preparing MoS2/Au composite electrodes as described by D. D. Xu et al. [17].The PLLA rapid transfer method, proposed by Hai Li et al. [21], is utilized to transfer the layered MoS2 grown on a silicon/silicon dioxide (Si/SiO2) substrate.
Figure 6 shows the fabrication process of the flexible pressure sensor, with 2D MoS2 materials on a PDMS substrate. Initially, a single layer MoS2 is grown on a silicon/silicon dioxide (SiO2) wafer using chemical vapor deposition (CVD) (as shown in Figure 6a). First, the silicon wafer is pretreated, and plasma cleaning is used to remove organic pollutants and oxide layers on the surface of the silicon wafer, increasing the surface energy to improve the adhesion of MoS2. Then, the silicon wafer is placed in a CVD furnace (SENTECH Instruments (Berlin-Adlershof): SENTECH SI 500), heated (650–850 °C), and a precursor nitrogen gas containing sulfur and molybdenum is introduced. The precursor is de-composed. The sulfur and the molybdenum atoms react on the surface of the silicon wafer to form a single layer of MoS2. Post annealing is conducted in a nitrogen atmosphere on the grown 2-D material.
PLLA in dichloromethane solution (3.0 wt %) is applied as a carrier on top of the MoS2 layer by spin coating at 3000 rpms. Subsequently, a polymer strip with a width of 1 mm is scratched along the edge of the SiO2/Si substrate to expose the hydrophilic SiO2 surface. Then, PDMS film is pressed onto the PLLA film. Deionized water is dripped onto the exposed hydrophilic SiO2 area, allowing it to penetrate between the substrate and the hydrophobic PLLA film. After 5–10 min, the PDMS/PLLA film is peeled off from the SiO2/Si substrate, as shown in Figure 6b. At the same time, ITO (indium tin oxide)/Al is deposited onto the PDMS substrate, as shown in Figure 6c. It is patterned as the sensor’s electrode, as shown in Figure 6d. Subsequently, the PDMS/MoS2 composite is transferred onto the ITO/Al electrode (Figure 6e). The PDMS/PLLA/MoS2 film is initially transferred and pressed onto the surface of the target substrate (Figure 6f). Utilizing the differential thermal expansion coefficients of PDMS and PLLA, the PDMS film is heated to 50 °C and then peeled off from the PLLA film’s surface. Finally, the PLLA film is dissolved using dichloromethane (DCM) to facilitate the transfer of MoS2 (Figure 6g).
Following the electrode deposition and MoS2 transfer processes described above (Figure 6g or Figure 7c), the photolithography of MoS2 is conducted. The Atomic Layer Etching (ALET) method, reported by T.Z. Lin et al. [22], is employed in this paper. ALET involves a cyclic process, in which chlorine (Cl) radicals are adsorbed onto the MoS2 surface. The followed desorption process, by using argon (Ar+) ion beams, can remove MoS2 layer by layer. Initially, chlorine radicals are generated using an inductively coupled plasma (ICP) system and are adsorbed onto the surface of MoS2. Subsequently, the MoS2 surface undergoes chlorination using a low-energy Ar+ ion beam, as shown in Figure 7d.
To determine the MoS2 size, we use photolithography to determine the size of the device, and then use a microscope to measure the planar dimensions and an Ellipsometer to measure the thickness.

3. Results and Discussion

COMSOL 6.1 software is used to simulate the piezoresistive effects of MoS2 on PDMS substrate. Table 1 presents various material performance parameters of PDMS and MoS2 used in the simulation.
According to the geometrical dimensions listed in Table 1, the upper layer consists of MoS2 measuring 0.5 × 0.8 × 0.1 microns (colored yellow in Figure 8a), while the lower layer is PDMS of identical dimensions (colored gray in Figure 8a). A pressure of 20 kPa is applied on the MoS2 + PDMS compound materials to observe the stress properties.
The constructed device model features PDMS as the substrate encapsulating MoS2. With the applied external pressure, PDMS undergoes deformation, exerting forces on MoS2 due to this deformation. Consequently, the force acting on MoS2 can be transferred to a vertical pressure p combined with a horizontal force p × 0.49 (where 0.49 represents Poisson’s ratio of PDMS), as illustrated in Figure 8b.
An ITO/Al electrode is deposited onto the PDMS surface. The test voltage ranging from 0 to 10 V is applied. Changes in the current under varying pressures can be observed. The relationship between ΔR/R and p can be established.
The relationships between the current and the voltage with different applied pressures are shown in Figure 9a, while the relationships between the resistance and the pressure are shown in Figure 9b. It is evident that the resistance increases with the increase in the applied pressure. This is mainly attributed to the piezoresistant effect of the MoS2 material. When MoS2 experiences a compressive strain less than 2%, its band gap would be widened, resulting in higher resistance [23]. Figure 9b illustrates the flexible pressure sensor with a sensitivity of 1.13 kPa−1.
To further enhance the sensitivity of PDMS + MoS2, a cavity structure within the PDMS substrate is introduced in the design. This exploration optimizes the stress conditions, strategically placing the piezoresistive material in areas of optimal stress to improve the sensitivity. Furthermore, the ratios of cavity height to PDMS thickness are varied, which are used to investigate their impacts on the stress conditions. The detailed parameters are listed in Table 2.
Figure 10 illustrates how stress varies with the different thickness of PDMS. Firstly, regardless of PDMS thickness, the cavity structure can amplify the pressure at its edges. If MoS2 piezoresistive material were placed at the edges, its piezoresistive performance would be significantly enhanced. Secondly, with increasing PDMS thickness, from 0.12 μm to 0.2 μm, the intensity of the pressure on the edge is gradually decreased. Therefore, to maximize sensitivity without compromising structural integrity under stress, the PDMS thickness with 0.12 μm shows the optimized performance.
In order to further explore the optimal PDMS cavity depth, we set the maximum deformation of the middle PDMS membrane to not exceed 30%.
Based on the results in Figure 11, the stability and the reliability of the sensor can be observed. It can be seen that the optimal PDMS thickness is 0.25 μm, which is 2.5 times the cavity depth.
Based on the results shown in Figure 10, the MoS2 piezoresistive materials are positioned at the point of maximum force to get the optimal result. Constructing the Wheatstone structure, its output voltage can be expressed with the following equation:
V O U T = ( R 2 R 1 + R 2 R 4 R 3 + R 4 ) V d d
The resistors R1~R4 can be expressed according to the piezoresistive effect:
R 1 = R 0 ( 1 + Δ 1 ) ( 1 P s e n s i t i v i t y ) R 2 = R 0 ( 1 + Δ 2 ) ( 1 + P s e n s i t i v i t y ) R 3 = R 0 ( 1 + Δ 3 ) ( 1 + P s e n s i t i v i t y ) R 4 = R 0 ( 1 + Δ 4 ) ( 1 P s e n s i t i v i t y )
Substituting the value of resistor R into the Wheatstone bridge output expression:
V O U T = ( ( 1 + Δ 2 ) ( 1 + P S ) ( 1 + Δ 1 ) ( 1 P S ) + ( 1 + Δ 2 ) ( 1 + P S ) ( 1 + Δ 4 ) ( 1 P S ) ( 1 + Δ 3 ) ( 1 + P S ) + ( 1 + Δ 4 ) ( 1 P S ) ) V d d
If we consider the ideal situation Δ 1 = Δ 2 = Δ 3 = Δ 4 ,
then Formula (8) becomes:
= ( 1 + P S 2 1 P S 2 ) V d d = P S V d d
It can be seen from Formula (9) that under ideal conditions, the final output voltage Vout of the sensor is related to three quantities, namely, the applied pressure p, the sensor sensitivity S, and the sensor operating voltage Vdd. When the operating voltage Vdd and the sensitivity S are constant, the output voltage of the sensor is solely linearly related to the applied pressure P.
Practically, variations in the manufacturing process can lead to unexpected asymmetry biases in the resistance values of the four resistors in the Wheatstone bridge. This can result in a non-zero output signal from the pressure sensor, even in the absence of applied pressure. Additionally, existing asymmetry in the piezoresistive coefficients of these resistors could further influence changes in the resistance values.
In this paper, the piezoresistive coefficient and the manufacturing process deviations are treated as the variables in the model. The simulation results of the Wheatstone bridge based on the designed flexible pressure sensors are shown in Figure 12. The sensitivity can be evaluated based on the results.
Based on the simulation results, the sensitivity S of the flexible pressure sensor is 3.729 V/kPa. The symmetrical Wheatstone bridge structure converts the pressure-induced resistance changes into voltage variations, facilitating the design of a piezoresistive pressure sensor.
To assess the impact of deviations in the piezoresistive coefficients and the manufacturing processes on the output voltage, assuming a 5% deviation, the values of variables Del1 to Del4 are adjusted accordingly. Table 3 lists the changed values of Del1 ~ Del4 in different cases.
Figure 13 shows the influences of the process variations on the sensor’s voltage output. It can be confirmed that variations in manufacturing processes and piezoresistive coefficients lead to resistance asymmetry and zero-point offset in the Wheatstone bridge structure. This offset results in an output voltage even without applied pressure. Moreover, it can also be observed that the trend of the sensor output voltage remains the same. A proportional relationship can be observed, showing good linearity of the sensor’s output.
Similarly, temperature changes also have such an impact on the output curve of the sensor. Their essence is to change the resistance through other factors, making the resistance values of various parts of the Wheatstone bridge inconsistent, resulting in zero drift. In addition to zero drift, temperature changes may affect the sensitivity of the bridge, that is, the response degree of the output voltage to input changes. This may lead to degraded accuracy of the measurement results, especially in environments with large temperature changes.
Theoretical analysis on other important characteristics of the sensor is discussed in the following section.
Response time refers to the speed at which a sensor reacts to a change in pressure, that is, the period starting from the applied pressure and ending when the sensor output reaches 90% of the stable value. It is related to the electron or ion mobility of the sensor material and the geometric characteristics of the signal transmission path in the sensor structure. MoS2 has a unique two-dimensional electron confinement property and high electron mobility [14], which makes a great contribution to expedite the response time of the sensor.
The relaxation time is the time required for the sensor resistance signal to be recovered from a stable value to its initial state after the pressure is removed. It is related to the viscoelasticity of the material and the contact conditions of the electrodes.
τ = η G
where τ is the relaxation time, η is the viscosity coefficient of the material, and G is the elastic modulus of the material, which can be replaced by Young’s modulus in this case.
Studies have shown that MoS2 interfacial adhesion energy is increased with decreasing MoS2 thickness and is gradually decreased with increasing temperature [24]. The shorter the relaxation time, the shorter the time it takes for the sensor to recover to its normal state. Based on its response time, the designed flexible sensor can be applied to environments requiring for faster pressure measurement.
From the Figure 14, by adjusting the frequency of force variation over time and observing the output signal, it can be found that the response time of the sensor is 53 ms.
The mechanical durability of a pressure sensor refers to the ability of the sensor to withstand repeated loading and unloading during its service life, as well as the degree to which it maintains stable performance under specific conditions. The fatigue life of the material is an important consideration. MoS2 has strong in-plane covalent bonds and atomic thickness, which gives it excellent mechanical strength (fracture strain > 2.2%) and flexibility [14]. These characteristics make MoS2-based pressure sensors have relatively strong mechanical durability. After 3000 pressure loading and unloading cycles, the PDMS cavity structure shows a 3% offset error.

4. Conclusions

In this paper, a novel flexible pressure sensor based on 2D MoS2 material with PDMS substrate is designed. Its detailed fabrication process is presented. Utilizing soft lithography, PDMS substrates with cavities are fabricated, which employ the PLLA rapid transfer method to integrate MoS2 onto the PDMS substrate. During the fabrication, Ar+ ion beam etching is utilized for MoS2 patterning. The cavity structure shows benefit in enhancing sensitivity, while the symmetrical Wheatstone bridge configuration facilitates accurate measurement of external pressure and improved temperature stability.
As shown in Table 4, a comparison of the sensitivity of various pressure sensors is made. Compared with the piezoresistive pressure sensor made of the same MoS2 material, the sensor designed in this paper has a slight advantage in terms of sensitivity. On the other hand, it can be found that the piezoresistive pressure sensor cannot be compared with the ionic electronic pressure sensor in terms of sensitivity. However, the proposed flexible sensor in this paper shows high stability and a simple structure for easy mass-fabrication. Compared with the sensor made of graphene-PDMS composites, the proposed sensor in this paper shows potential for improvement, since it can also adopt other structures, such as microstructures and arrays, to improve sensitivity.
The pressure sensor exhibits a relatively outstanding sensitivity of 1.13 kPa−1. Furthermore, the sensitivity could be further enhanced by increasing the proportion of cavity to substrate. The piezoresistive characteristics of a sensor employing a MoS2 + PDMS Wheatstone bridge structure is investigated in this paper. Overall, the results highlight the promising application potential of this new flexible pressure sensor with excellent sensitivity in fields such as medical treatment, sports monitoring, and wearable devices.

Author Contributions

Conceptualization, S.D. and F.L.; data curation, S.D.; formal analysis, Y.J.; investigation, S.D.; methodology, S.D. and F.L.; resources, Y.J. and M.C.; software, Y.J.; supervision, Y.J. and M.C.; validation, S.D.; writing—original draft, S.D.; writing—review & editing, Y.J. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported in part by Leading Technology Pre-Research Project (No. XD23008, Wuxi Industrial Innovation Research Institute and Jiangsu JITRI IC Application Technology Innovation Center), National Natural Science Foundation of China under Grant No. 61774078 and Natural Science Foundation of Jiangsu Province (BK20210453).

Data Availability Statement

Data available on request from the authors.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Wang, H.; Li, Z.; Liu, Z.; Fu, J.; Shan, T.; Yang, X.; Lei, Q.; Yang, Y.; Li, D. Flexible capacitive pressure sensors for wearable electronics. J. Mater. Chem. C 2022, 10, 1594–1605. [Google Scholar] [CrossRef]
  2. An, B.W.; Heo, S.; Ji, S.; Bien, F.; Park, J.U. Transparent and flexible fingerprint sensor array with multiplexed detection of tactile pressure and skin temperature. Nat. Commun. 2018, 9, 2458. [Google Scholar] [CrossRef]
  3. Wang, M.; Luo, Y.; Wang, T.; Wan, C.; Pan, L.; Pan, S.; He, K.; Neo, A.; Chen, X. Artificial Skin Perception. Adv. Mater. 2021, 33, e2003014. [Google Scholar] [CrossRef]
  4. Zhu, B.; Niu, Z.; Wang, H.; Leow, W.R.; Wang, H.; Li, Y.; Zheng, L.; Wei, J.; Huo, F.; Chen, X. Microstructured Graphene Arrays for Highly Sensitive Flexible Tactile Sensors. Small 2014, 10, 3625–3631. [Google Scholar] [CrossRef]
  5. Kim, K.-H.; Hong, S.K.; Jang, N.-S.; Ha, S.-H.; Lee, H.W.; Kim, J.-M. Wearable Resistive Pressure Sensor Based on Highly Flexible Carbon Composite Conductors with Irregular Surface Morphology. ACS Appl. Mater. Interfaces 2017, 9, 17500–17508. [Google Scholar] [CrossRef]
  6. Luo, N.; Huang, Y.; Liu, J.; Chen, S.C.; Wong, C.P.; Zhao, N. Hollow-Structured Graphene-Silicone-Composite-Based Piezoresistive Sensors: Decoupled Property Tuning and Bending Reliability. Adv. Mater. 2017, 29, 1702675. [Google Scholar] [CrossRef]
  7. Ma, Y.; Yue, Y.; Zhang, H.; Cheng, F.; Zhao, W.; Rao, J.; Luo, S.; Wang, J.; Jiang, X.; Liu, Z.; et al. 3D Synergistical MXene/Reduced Graphene Oxide Aerogel for a Piezoresistive Sensor. ACS Nano 2018, 12, 3209–3216. [Google Scholar] [CrossRef]
  8. Zhang, Y.; Wang, L.; Zhao, L.; Wang, K.; Zheng, Y.; Yuan, Z.; Wang, D.; Fu, X.; Shen, G.; Han, W. Flexible Self-Powered Integrated Sensing System with 3D Periodic Ordered Black Phosphorus@MXene Thin-Films. Adv. Mater. 2021, 33, e2007890. [Google Scholar] [CrossRef]
  9. Guerin, S.; Stapleton, A.; Chovan, D.; Mouras, R.; Gleeson, M.; McKeown, C.; Noor, M.R.; Silien, C.; Rhen, F.M.F.; Kholkin, A.L.; et al. Control of piezoelectricity in amino acids by supramolecular packing. Nat. Mater. 2018, 17, 180–186. [Google Scholar] [CrossRef]
  10. De, S.; Asthana, D.; Thirmal, C.; Keshri, S.K.; Ghosh, R.K.; Hundal, G.; Kumar, R.; Singh, S.; Chatterjee, R.; Mukhopadhyay, P. A folded π-system with supramolecularly oriented dipoles: Single-component piezoelectric relaxor with NLO activity. Chem. Sci. 2023, 14, 2547–2552. [Google Scholar] [CrossRef]
  11. Kim, S.J.; Mondal, S.; Min, B.K.; Choi, C.G. Highly Sensitive and Flexible Strain-Pressure Sensors with Cracked Paddy-Shaped MoS2/Graphene Foam/Ecoflex Hybrid Nanostructures. ACS Appl. Mater. Interfaces 2018, 10, 36377–36384. [Google Scholar] [CrossRef]
  12. Lu, W.; Yu, P.; Jian, M.; Wang, H.; Wang, H.; Liang, X.; Zhang, Y. Molybdenum Disulfide Nanosheets Aligned Vertically on Carbonized Silk Fabric as Smart Textile for Wearable Pressure-Sensing and Energy Devices. ACS Appl. Mater. Interfaces 2020, 12, 11825–11832. [Google Scholar] [CrossRef]
  13. Gu, P.-C.; Zhang, K.-L.; Feng, Y.-L.; Wang, F.; Miao, Y.-P.; Han, Y.-M.; Zhang, H.-X. Recent progress of two-dimensional layered molybdenum disulfide. Acta Phys. Sin. 2016, 65, 4832975. [Google Scholar]
  14. Chen, X.; Park, Y.J.; Kang, M.; Kang, S.-K.; Koo, J.; Shinde, S.M.; Shin, J.; Jeon, S.; Park, G.; Yan, Y.; et al. CVD-grown monolayer MoS2 in bioabsorbable electronics and biosensors. Nat. Commun. 2018, 9, 1690. [Google Scholar] [CrossRef]
  15. Pang, X.; Zhang, Q.; Shao, Y.; Liu, M.; Zhang, D.; Zhao, Y. A Flexible Pressure Sensor Based on Magnetron Sputtered MoS2. Sensors 2021, 21, 1130. [Google Scholar] [CrossRef]
  16. Chen, X.; Zhang, D.; Luan, H.; Yang, C.; Yan, W.; Liu, W. Flexible Pressure Sensors Based on Molybdenum Disulfide/Hydroxyethyl Cellulose/Polyurethane Sponge for Motion Detection and Speech Recognition Using Machine Learning. ACS Appl. Mater. Interfaces 2023, 15, 2043–2053. [Google Scholar] [CrossRef]
  17. Xu, D.; Duan, L.; Yan, S.; Wang, Y.; Cao, K.; Wang, W.; Xu, H.; Wang, Y.; Hu, L.; Gao, L. Monolayer MoS2-Based Flexible and Highly Sensitive Pressure Sensor with Wide Sensing Range. Micromachines 2022, 13, 660. [Google Scholar] [CrossRef]
  18. Chen, H.; Zhuo, F.; Zhou, J.; Liu, Y.; Zhang, J.; Dong, S.; Liu, X.; Elmarakbi, A.; Duan, H.; Fu, Y. Advances in graphene-based flexible and wearable strain sensors. Chem. Eng. J. 2023, 464, 142576. [Google Scholar] [CrossRef]
  19. Meng, Q.; Lu, Y.; Wang, J.; Chen, D.; Chen, J. A Piezoresistive Pressure Sensor with Optimized Positions and Thickness of Piezoresistors. Micromachines 2021, 12, 1095. [Google Scholar] [CrossRef]
  20. Hammock, M.L.; Chortos, A.; Tee, B.C.K.; Tok, J.B.H.; Bao, Z. 25th Anniversary Article: The Evolution of Electronic Skin (E-Skin): A Brief History, Design Considerations, and Recent Progress. Adv. Mater. 2013, 25, 5997–6037. [Google Scholar] [CrossRef]
  21. Li, H.; Wu, J.; Huang, X.; Yin, Z.; Liu, J.; Zhang, H. A Universal, Rapid Method for Clean Transfer of Nanostructures onto Various Substrates. ACS Nano 2014, 8, 6563–6570. [Google Scholar] [CrossRef] [PubMed]
  22. Lin, T.; Kang, B.; Jeon, M.; Huffman, C.; Jeon, J.; Lee, S.; Han, W.; Lee, J.; Lee, S.; Yeom, G.; et al. Controlled Layer-by-Layer Etching of MoS2. ACS Appl. Mater. Interfaces 2015, 7, 15892–15897. [Google Scholar] [CrossRef] [PubMed]
  23. Lu, P.; Wu, X.; Guo, W.; Zeng, X.C. Strain-dependent electronic and magnetic properties of MoS2 monolayer, bilayer, nanoribbons and nanotubes. Phys. Chem. Chem. Phys. 2012, 14, 13035–13040. [Google Scholar] [CrossRef]
  24. Duan, C.; Liu, J.; Chen, Y.; Zuo, H.; Dong, J.; Ouyang, G. Adhesion properties of MoS2/SiO2 interface: Size and temperature effects. Acta Phys. Sin. 2024, 73, 056801. [Google Scholar] [CrossRef]
Figure 1. Operation principle diagram of piezoresistive pressure sensor. (a) Schematic of piezoresistive pressure sensor based on rigid material. (b) Schematic of piezoresistive pressure sensor on flexible material.
Figure 1. Operation principle diagram of piezoresistive pressure sensor. (a) Schematic of piezoresistive pressure sensor based on rigid material. (b) Schematic of piezoresistive pressure sensor on flexible material.
Symmetry 16 01242 g001
Figure 2. Cross-sectional schematic of the piezoresistive pressure sensor.
Figure 2. Cross-sectional schematic of the piezoresistive pressure sensor.
Symmetry 16 01242 g002
Figure 3. Schematic diagram of Wheatstone bridge structure based on flexible pressure sensors.
Figure 3. Schematic diagram of Wheatstone bridge structure based on flexible pressure sensors.
Symmetry 16 01242 g003
Figure 4. (a) Schematic diagram of the structure of PDMS + MoS2 flexible pressure sensor. (b) Measurement setup for the flexible pressure sensor.
Figure 4. (a) Schematic diagram of the structure of PDMS + MoS2 flexible pressure sensor. (b) Measurement setup for the flexible pressure sensor.
Symmetry 16 01242 g004
Figure 5. Preparation steps of PDMS cavity with soft lithography process. (a) Si/SiO2 mold (b) The degassed PDMS is poured onto the silanized mold (c) The PDMS is removed from the mold.
Figure 5. Preparation steps of PDMS cavity with soft lithography process. (a) Si/SiO2 mold (b) The degassed PDMS is poured onto the silanized mold (c) The PDMS is removed from the mold.
Symmetry 16 01242 g005
Figure 6. Fabrication process of the flexible pressure sensor. (a) MoS2 is grown on a Si/SiO2 wafer (b) PLLA and PDMS are bonded to the MoS2 surface (c) ITO (indium tin oxide)/Al is deposited onto the PDMS substrate (d) Separate the ITO/Al electrode from the PDMS (e) Transfer the PDMS/MoS2 composite material onto the ITO/Al electrodes (f) The image after transferring (g) Remove PLLA and PDMS.
Figure 6. Fabrication process of the flexible pressure sensor. (a) MoS2 is grown on a Si/SiO2 wafer (b) PLLA and PDMS are bonded to the MoS2 surface (c) ITO (indium tin oxide)/Al is deposited onto the PDMS substrate (d) Separate the ITO/Al electrode from the PDMS (e) Transfer the PDMS/MoS2 composite material onto the ITO/Al electrodes (f) The image after transferring (g) Remove PLLA and PDMS.
Symmetry 16 01242 g006
Figure 7. Schematic diagram of the manufacturing process of PDMS + MoS2 flexible pressure sensor. (a) PDMS with cavity structure (b) Deposit electrodes on the PDMS surface (c) Transfer MoS2 onto the electrode (d) Photolithography of MOS2.
Figure 7. Schematic diagram of the manufacturing process of PDMS + MoS2 flexible pressure sensor. (a) PDMS with cavity structure (b) Deposit electrodes on the PDMS surface (c) Transfer MoS2 onto the electrode (d) Photolithography of MOS2.
Symmetry 16 01242 g007
Figure 8. PDMS + MoS2 model construction and stress conditions (a) MoS2 and PDMS structure diagram (b) MoS2 + PDMS stress distribution diagram.
Figure 8. PDMS + MoS2 model construction and stress conditions (a) MoS2 and PDMS structure diagram (b) MoS2 + PDMS stress distribution diagram.
Symmetry 16 01242 g008
Figure 9. (a) Relationship between the current and the voltage of PDMS + MoS2 flexible sensor with different applied pressures. (b) Relationship between the resistance and the pressure of PDMS + MoS2 flexible sensor.
Figure 9. (a) Relationship between the current and the voltage of PDMS + MoS2 flexible sensor with different applied pressures. (b) Relationship between the resistance and the pressure of PDMS + MoS2 flexible sensor.
Symmetry 16 01242 g009
Figure 10. Stress of the cavity structure and its variation with PDMS thickness. (a) H = 0.12 μm (b) H = 0.16 μm (c) H = 0.18 μm (d) H = 0.2 μm.
Figure 10. Stress of the cavity structure and its variation with PDMS thickness. (a) H = 0.12 μm (b) H = 0.16 μm (c) H = 0.18 μm (d) H = 0.2 μm.
Symmetry 16 01242 g010
Figure 11. Variation of the maximum deformation of PDMS membrane with H.
Figure 11. Variation of the maximum deformation of PDMS membrane with H.
Symmetry 16 01242 g011
Figure 12. Simulation results of the output voltage of the Wheatstone bridge with the changing pressure.
Figure 12. Simulation results of the output voltage of the Wheatstone bridge with the changing pressure.
Symmetry 16 01242 g012
Figure 13. Effect of process variation on Wheatstone bridge voltage output.
Figure 13. Effect of process variation on Wheatstone bridge voltage output.
Symmetry 16 01242 g013
Figure 14. Output changes with periodic pressure changes.
Figure 14. Output changes with periodic pressure changes.
Symmetry 16 01242 g014
Table 1. Material modeling parameters of MoS2 and PDMS for simulation.
Table 1. Material modeling parameters of MoS2 and PDMS for simulation.
MaterialPDMSMoS2
Conductivity 0.1 [S/cm]0.5 [S/cm]
Young’s modulus750 [kPa]280 [GPa]
Poisson’s ratio0.490.29
Piezoresistive coupling matrix 22.62 [1/MPa] [15]
MoS2 + PDMS
Size (μm)Upper MoS2: 0.5 × 0.8 × 0.1
Lower layer PDMS: 0.5 × 0.8 × 0.1
Table 2. Dimensions of PDMS and cavity in flexible pressure sensor.
Table 2. Dimensions of PDMS and cavity in flexible pressure sensor.
Size (μm)PDMSCavity
Length × Width1 × 10.5 × 0.5
Height H (From 0.12 μm to 0.2 μm)0.1
Pressure applied20 kPa
Table 3. Changed values of Del1 ~ Del4 in different cases.
Table 3. Changed values of Del1 ~ Del4 in different cases.
ParameterCase1Case2Case3
Del100.05−0.05
Del20−0.050.05
Del30−0.050.05
Del400.05−0.05
Table 4. Comparison of sensitivity of various pressure sensors.
Table 4. Comparison of sensitivity of various pressure sensors.
Pressure SensorSensitivityReference
PDMS + MoS2 (this work)1.13 kPa−1
MoS2/HEC/PU sponge pressure sensor0.746 kPa−1[16]
ionic electronic pressure sensor based on MoS2 material89.75 kPa−1[17]
graphene-PDMS composites15.9 kPa−1[6]
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

Deng, S.; Li, F.; Cai, M.; Jiang, Y. Novel Flexible Pressure Sensor with Symmetrical Structure Based on 2-D MoS2 Material on Polydimethylsiloxane Substrate. Symmetry 2024, 16, 1242. https://doi.org/10.3390/sym16091242

AMA Style

Deng S, Li F, Cai M, Jiang Y. Novel Flexible Pressure Sensor with Symmetrical Structure Based on 2-D MoS2 Material on Polydimethylsiloxane Substrate. Symmetry. 2024; 16(9):1242. https://doi.org/10.3390/sym16091242

Chicago/Turabian Style

Deng, Shaoxiong, Feng Li, Mengye Cai, and Yanfeng Jiang. 2024. "Novel Flexible Pressure Sensor with Symmetrical Structure Based on 2-D MoS2 Material on Polydimethylsiloxane Substrate" Symmetry 16, no. 9: 1242. https://doi.org/10.3390/sym16091242

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