Open Access Article
A. S. M. Jannatul Islam
a,
Md. Sherajul Islam
*a,
Ashraful G. Bhuiyana,
Catherine Stampfl
b and
Jeongwon Park
c
aDepartment of Electrical and Electronic Engineering, Khulna University of Engineering &Technology, Khulna 9203, Bangladesh. E-mail: sheraj_kuet@eee.kuet.ac.bd
bSchool of Physics, The University of Sydney, New South Wales 2006, Australia
cDepartment of Electrical and Biomedical Engineering, University of Nevada, Reno, NV 89557, USA
First published on 11th December 2025
The pronounced electronegativity difference and exceptional non-centrosymmetricity of two-dimensional (2D) cadmium sulfide (CdS) offer an exceptional piezoelectric effect, positioning it as a promising material for next-generation energy-harvesting devices. This study examines the piezoelectric properties of 2D CdS bilayers driven by tribological effects and assesses their feasibility for nano energy harvesting applications. Two types of tribological effects, namely in-plane sliding motion and out-of-plane compression, are applied to a CdS bilayer to investigate its piezoelectric properties, including potential energy, polarization deviations, charge density, shear strength, induced voltage, and power density. The results indicate that in-plane sliding, which transitions the bilayer structure from an A–B to an A–A stacking pattern, leads to a significant increase in out-of-plane piezoelectricity. Upon vertical compression, we find that the range of maximal energy corrugation is 724–1365 meV, and the range of shear strength is 49.79–99.20 GPa. Additionally, due to vertical compression, an induced voltage of 0.35 V is found for bilayer 2D CdS. Finally, a compressive-sliding motion-based nanogenerator model is proposed for a 2D CdS bilayer structure that can engender an output power density of up to 51.72 mW cm−2. These outcomes pave the way for new opportunities in nanoenergy extraction from the tribo-piezoelectric effect of CdS structures, advancing the development of wearable electronics, self-powered devices, and wireless technologies.
In this context, several piezoelectric materials have been investigated, each presenting distinct benefits and obstacles. Conventional piezoelectric materials, including lead zirconate titanate (PZT) nanofibers, gallium nitride (GaN), aluminum nitride (AlN), and quartz, have historically prevailed in the domain owing to their robust piezoelectric responses.8,13,16–19 Nonetheless, these materials have intrinsic obstacles. PZT, despite its superior performance, is plagued by environmental issues associated with its lead concentration. GaN and AlN have exceptional thermal stability; yet, their high production costs and stiffness restrict their use in flexible and wearable electronics. Emerging materials such as zinc oxide (ZnO) nanostructures and polyvinylidene fluoride (PVDF) have broadened the potential for energy harvesting. ZnO nanowires, especially in lateral configurations, provide significant piezoelectric sensitivity and scalability; however, need meticulous alignment and structural integrity for best performance. Conversely, PVDF offers flexibility and simplicity of production, but its inferior piezoelectric coefficients relative to inorganic materials limit its use in high-output systems.
Amid this landscape, two-dimensional (2D) materials have of late gained immense interest for advancing nanoscale ambient energy harvesting devices, owing to their exceptional properties such as remarkable quantum Hall effects, ultra-thin structure, mechanical stability, biocompatibility, flexibility, and high carrier mobility.20–27 It has been revealed that due to their characteristics of high piezoelectric constants, and breakdown of inversion symmetry, layered 2D materials—for instance, WSe2,28 MoS2,29 SnS,30 and MXenes31 can reveal strong in-plane piezoelectricity. The out-of-plane response is often limited to odd-numbered layers and gradually declines with increasing thickness. Several 2D systems, such as CuInP2S6, LiNbO3, In2Se3, and NbOI2, have shown significant out-of-plane piezoelectricity.32–37 This is due to twist-angle-dependent symmetry destroying, interface-engineered nanostructures, and ferroelectric composite interactions. However, these techniques frequently rely on extensive surface modification steps and have limited tunability, resulting in significant variation in piezoelectric performance.
Tribological phenomena, including lateral sliding, out-of-plane compression, and interfacial friction-driven variation of polarization, have recently emerged as a possible alternative method for increasing piezoelectricity. Such tribo-piezoelectric phenomena have been observed in nitride bilayers and Janus TMD bilayers;38–42 however, the induced voltage, power density, shear strength, and out-of-plane polarization are all moderate. These limitations highlight the need for new bilayer systems that can generate stronger and more robust tribo-piezoelectric responses. Inspired by this issue, the current investigation explores group-II sulfides, with a specific emphasis on bilayer CdS, to investigate their potential as a next-generation 2D energy-harvesting material. 2D CdS exhibits excellent thermodynamic stability, an impressive electronic bandgap (2.5 eV), a low work function, high mechanical strength, strong excitonic binding energy, and remarkable optical properties.43–46 In addition, CdS belongs to the family of 2D II–VI chalcogenides that can be synthesized through well-established techniques such as chemical vapor deposition and solution-mediated exfoliation, ensuring the practical representativeness of the system.47–50 Applications of 1D and 2D CdS nanostructures have been proven in many different domains, including waveguides, dye-sensitized solar cells, photoconductors, thermo-electronics, field emitters, and logic gates.43–46 Although 2D CdS is considered a promising piezoelectric material owing to the substantial electronegativity difference between Cd (1.69) and S (2.58) atoms, a comprehensive understanding of tribology-induced piezoelectric responses in bilayer 2D CdS is still missing.
This work uses a comprehensive first-principles density functional theory (DFT) analysis to study the tribo-piezoelectricity of 2D CdS bilayers. We show that bilayer CdS might be stimulated to undergo vertical polarization by in-plane sliding along a specific route of two CdS monolayers. The polarization amplification is connected to greater vertical piezoelectricity and interfacial charge transmission once the CdS bilayer overcomes the in-plane sliding obstacle and transforms into the A–A stacking configuration. The potential energy variation, polarization deviation, and shear strength alteration with interlayer gap change reveal tunable piezoelectric properties. Both the vertical charge polarization and the horizontal sliding energy resistance of a CdS bilayer are improved with decreasing interlayer spacing. In-plane sliding and associated opposition are tribological activities. Nevertheless, there is a remarkable strengthening of out-of-plane polarization, and no triboelectric charges are produced during interlayer sliding in the transition from A–B to A–A assembling pattern with a given interlayer spacing. This suggests that stronger vertical polarization has been triggered by tribological energy, resulting in piezoelectricity instead of triboelectricity. It should be noted that, because interlayer sliding does not produce any triboelectric charges, this type of tribo-piezoelectricity differs from traditional piezo-triboelectricity. Since a variation in the interlayer movement can produce vertical piezoelectricity, a compressive slipping nanoenergy harvester with a CdS bilayer arrangement is suggested based on the tribo-piezoelectric influence and can deliver a generated voltage as high as ∼0.35 V and output power density of up to 51.72 mW cm−2.
| EPES(x,y) = E(x,y) − Elowest | (1) |
![]() | (2) |
| ΔP(x,y) = Pd(x,y) − PFn | (3) |
![]() | (4) |
In this study, DFT is used for all computations, and the plane wave self-consistent field (PWscf) suite of Quantum Espresso is utilized.51,52 The exchange-correlation functional is taken to be the PBE functional for the generalized gradient approximation.53 The electron–ion interactions in the PWscf program are described by Kresse–Joubert projector augmented wave (KJPAW) potentials.54 The weak intermolecular van-der Waals (vdW) relations are represented by the optB86b-vdW functional.55,56 The initial PBE vdW-DF is altered by optB86b-vdW relations to guarantee correct replication for the interatomic distances and energies at equilibrium.55–58 A 2D-centered Monkhrost–Pack k-mesh grid with dimensions of 15 × 15 × 1 is used to carry out the Brillouin zone integration. An energy cut-off value of 10−8 a.u. is utilized to ensure correct convergence of the ground state electron density. The charge densities and wave function have kinetic energy cut-offs of 350 Ry and 45 Ry, respectively. Furthermore, for precise stress and force computations, a force convergence cutoff of 10−3 eV nm−1 is employed. To prevent direct contact between the monolayers and bilayers, a 20 Å vacuum space is applied.
The two distinct interlayer distances (3.14 Å and 2.84 Å) are used to examine the impact of transverse sliding. Fig. 2 presents the calculated PES and PDS distributions at various sliding positions. As would be expected from the equilibrium geometry of the flat honeycomb constructed CdS, the PES and PDS maps exhibit periodic repetition.59 As seen in Fig. 2, very strong vertical polarization is shown when the original arrangement slides to resemble the A–A stacking arrangement. Moreover, with a smaller interlayer separation, this strong vertical polarization increases as well. However, when the interlayer spacing changes, the A–B stacking arrangement yields comparatively little vertical polarization. In addition, the A–B stacking responds to interlayer spacing far less strongly than the A–C stacking. Consequently, the A–B stacking arrangement is considered the starting state, and the A–A stacking arrangement will achieve the greatest polarization improvement. Nevertheless, the A–A stacking is not energetically favorable in the absence of an outside mechanical input or constraint. Only an in-plane pushing force stronger than the shear strength and interlayer resistive force may cause a switch to the A–A arrangement during the lateral slipping of the A–B configuration.
The resistance that accompanies interlayer sliding is a tribological phenomenon, but vertical polarization is greatly enhanced when A–A stacking is used with equilibrium interlayer spacing. Therefore, transverse interlayer slipping in the CdS bilayer may be considered as the cause of the vertical polarization amplification, which can be explained as a tribo-piezoelectric phenomenon. Through tribological energy transformation, the energy obstruction of sliding resistance is overcome, leading to piezoelectricity. Outside external force and compression are needed to push the CdS bilayer to the A–A state, which is equivalent to transforming tribological energy into electrical power. The triboelectric effect is maintained at a given interlayer spacing through the procedure of overwhelming the energy barrier associated with interlayer shear strength and resistive force. Therefore, the tribo-piezoelectric action will be the cause of the vertical plane slipping that resulted in vertical polarization. In our work, the tribological energy generates piezoelectricity, in contrast to earlier research where the triboelectric effect results in static charges.9,13 Cai et al. have also discovered analogous tribo-piezoelectric behaviors in Janus TMD bilayers.69 Thus, out-of-plane interlayer resistive frictions determine the corrugated energy (ΔE) in the PES of the CdS bilayer; greater interlayer resistance and friction would result in a larger energy corrugation.70
The energy corrugation (ΔE) determines how much tribo-piezoelectricity is generated. The compressive fluctuation of the interlayer separation from 3.14 Å to 2.74 Å is shown in Fig. 3(a), which shows the greatest polarization deviation (ΔPmax) and maximum energy corrugation (ΔEmax). Fig. 3(a) shows that when the maximum energy corrugation increases, the maximum polarization deviation escalates linearly. The stronger the magnitude of vertical polarization, the more energy will be needed to overwhelm the interlayer sliding friction. Relative shear strength can be obtained from eqn (4) by using the greatest static resistance force acting on the sliding progression. The shear strength at an interlayer spacing between 2.74 Å and 3.14 Å is revealed in Fig. 3(b). Throughout the vertical compressive displacement of the CdS layer, the highest energy barrier variation (724–1670 meV) and shear strength (49–125 GPa) are greater than those attained with the Janus TMD bilayers (180–480 meV) and (2.5–7.2 GPa), correspondingly.69 The CdS bilayer may also yield a larger vertical polarization value (1.73–2.44 pC m−1) than the Janus TMD bilayers (0.6–2.3 pC m−1).69 A decrease in the interlayer spacing results in a quadratic rise in the values of maximum corrugated energy and shear strength. To obtain a greater triboelectric to electrical energy conversion, vertical compressive pressure is therefore an effective method. The discrepancies are principally due to CdS's highly ionic Cd–S bonds with strong polarity, which induce significant interlayer electrostatic attraction and charge redistribution during sliding. This results in a conspicuous corrugated potential-energy landscape and high resistance to lateral motion, particularly under compression, where Cd-5s/S-3p orbital overlap improves interlayer coupling. Janus TMDs are characterized by weak vdW interactions and limited interlayer charge transfer that provide low-energy sliding routes. Moreover, because of the impact of in-plane strain brought on by vertical tension, a structural variation has been seen throughout this procedure. The bilayer's atoms' interaction is affected by the change in interlayer distance; this impact is greater when the bilayer is positioned at a smaller interlayer gap than the equilibrium one. At the optimal interlayer spacing of 3.14 Å, the lattice constant was found to be 4.251 Å. However, reducing the interlayer distance lowers the lattice constant. The lattice constant decreases to 4.225 Å and 4.219 Å, respectively, with spacings of 2.84 Å and 2.74 Å. Upon applying vertical tension, the system is relaxed each time, and the computations account for variations in the lattice constant.
The charge density differences (CDD) in the CdS bilayer are studied to illustrate the charge involvement in the observed tribo-piezoelectric characteristic. The CDD at two interlayer distances (3.14 Å and 2.84 Å) between the CdS monolayers is shown in Fig. 4. The CDD may be expressed as follows:
| Δρ = ρtotal − ρtop − ρbottom | (5) |
![]() | (6) |
![]() | (7) |
and
with zero average forces. The typical fluctuations in charge density with interlayer distances ranging from 3.14 to 2.74 Å are depicted in Fig. 5(a) and (b). It is observed that an increase in the mean charge density inside the layers is caused by a decrease in the interlayer spacing. In the A–B stacking pattern, a larger charge density is seen at the bottom layer than at the top layer due to vertical compressive movement concerning the layers. In the A–A stacking arrangement, the charge density of the upper layer predominates at the same time. The charge allocation in the CdS upper and lower monolayers is in good accord with the CDD plot (Fig. 4).
Unlike conventional triboelectric nanogenerators—where electrical output relies on interaction between two different materials with dissimilar electron affinities—the bilayer CdS system is composed of two identical CdS sheets. Because the two surfaces possess the same electron affinity, significant interfacial charge transmission through the triboelectric effect is not demanded. As a substitute, the measured output originates solely from piezoelectric polarization, which is generated when lattice asymmetry is introduced during mechanical deformation (sliding or compression). To verify this mechanism, we performed Bader charge analysis39 for various stacking configurations and interlayer distances. The results confirm that the net charge transfer between the two layers is negligible, indicating that triboelectric contributions are minimal. The polarization response is therefore dominated by intralayer electron-density redistribution induced by mechanical perturbation, fully consistent with a piezoelectric rather than triboelectric origin. A summary of the Bader charge values for bilayer CdS under different structural conditions is provided in Table 1.
| Stacking pattern | Interlayer distance (Å) | Net Bader charge transfer (e) |
|---|---|---|
| A–A | 3.14 | −0.046855 |
| A–B | 3.14 | +0.040976 |
| A–A | 2.84 | −0.006575 |
| A–B | 2.84 | +0.071818 |
Compressive-sliding nanogenerators can be developed by taking into consideration the charge reallocation between the atoms and layers. Since most of the charge builds up around the S-atoms, the electron density surrounding the S-surface will be larger than the electron density at the interface between the layers. As depicted in Fig. 5(c), assigning electrodes to the upper and lower surfaces can provide a little voltage. As seen in Fig. 5(d), a greater concentration of inductive charges is observed close to the interface for the A–A stacking pattern. As a result, creating a nanogenerator with the A–A stacking arrangement will enable the extraction of a higher voltage. Moreover, Fig. 5(e) shows the charge gain in the upper and lower layers during the out-of-plane compression procedure. The charge gain is determined by detracting the single layer charges between the unit cells of the A–A and A–B assembling arrangements.
Subsequently, taking the A–A assembling arrangement, the electrostatic potential variations (ΔU) between the S atoms in the upper and lower layers are analyzed to calculate the maximum generated voltage. At the monolayer surfaces, a variation in the electrostatic potential differences with interlayer change can be determined by:
| ΔU = ΔUT − ΔUB | (8) |
![]() | (9) |
![]() | (10) |
and
denote the conforming cases at zero average forces. As seen in Fig. 6, a reduction in the interlayer spacing results in a rise in out-of-plane polarization and a rise in the electrostatic potential at both the upper and lower layers. During the procedure, an induced voltage between 0.23 and 0.35 V can be obtained by using the formula,
, where q denotes the unit charge and φ represents the induced voltage, to account for the electrostatic potential and unit charges.
To utilize the suggested bilayer CdS configuration as a nanogenerator, Fig. 7 presents the construction of a compression-sliding motion model. Using the nanoscale probing technique, force or rotation on a single layer or many layers of material at a certain route and trend may be incorporated to experimentally build the suggested nanogenerator.69,71–73 In this instance, a tiny CdS flake moves on a large CdS substrate in response to transverse and vertical motion imparted by means of the probe tip. The bilayer CdS is connected to two electrodes at its edges, with the upper electrode linked to a probe for controlling the mobility of the flake. The probe is set up to perform as a conductor, joining the lower electrode.74 In this concept, Au electrodes that are frequently utilized can be employed. Our further studies show that the adhesion force between the Au electrode and the CdS monolayer surface is greater than the force between two CdS monolayers. Moreover, the Au contact has minimal effect on the tribo-piezoelectric properties of the CdS bilayer, as shown in Fig. 8. The compression-sliding method is durable and strong with the CdS bilayer, as shown by our findings and previous research on the Au/2D materials contact.
![]() | ||
| Fig. 7 Diagrammatic representation of a prototype CdS bilayer nanogenerator based on compression-sliding motion. The probe is represented in blue, and the electrodes are shown in yellow. | ||
Taking a relaxed Au-supported CdS structure, as seen in Fig. 8(a), the impact of applying a metallic Au electrode on the CdS monolayer is investigated. The computations are performed in the same manner as described in Section 2. To align with the lattice constant of the Au substrate, the CdS's lattice constant was compressed by approximately 1.9%. The details of the lattice mismatch between CdS and Au, as well as the supercell matching strategy, are provided in the SI. Between the Au and CdS layers, an optimal interlayer gap of 2.74 Å is found. The Au–CdS contact will thus exhibit physical adsorption. For the Au/CdS and CdS bilayers, the calculated binding energies are −2.935 eV and −2.713 eV, respectively. The adhesion force between the Au electrode and the CdS monolayer surface is considerably greater than the force generated between two CdS monolayers. An analogous outcome was also discovered at the Au–TMD interface.69 Additionally, we have illustrated the projected density of states of the Cd and S atoms for the CdS monolayer and Au/CdS interface, as seen in Fig. 8(c) and (d). At the Au/CdS interface, the monolayer CdS's Fermi level is pinned close to the conduction band, suggesting that the interface is an n-type contact. Charge carrier transport between the Au and CdS interfaces is therefore restricted as a result of the creation of an interlayer Schottky barrier with a superior contact resistance between the Au/CdS contact. Moreover, we have extended the contact analysis beyond the PDOS by including a band structure diagram (Fig. S1) in the SI, which explicitly shows the relative positions of the Fermi level and the conduction- and valence-band edges at the Au/CdS interface. This diagram clearly confirms the formation of an n-type Schottky barrier, consistent with the high work function of Au and its tendency to pin the Fermi level near the CdS conduction band. Our results are consistent with previous theoretical investigations of the metal–2D material contact.75,76 The tribo-piezoelectricity process demonstrated in Fig. 5 states that inductive charges are generated when the electrode comes into contact with the CdS layer and that no charge is transferred from the CdS to the electrode. The Schottky barrier at the interface will obstruct the charge flow between the CdS layer and the metal electrode. Consequently, the tribo-piezoelectricity of the CdS bilayer is mostly insensitive to the metal contact. Additionally, recent Au/TMD tests have shown similar electrical characteristics, suggesting that the Au electrode has a minimal impact on the TMD layer.77,78
Mechanical energy is transformed into electrical energy by the interlayer resistive and sliding forces. Consequently, the majority of the electrical energy (Eelec) produced during the sliding operation will come from tribological energy (Etribo). The relation can be used to determine the quantity of Eelec generation if a conversion coefficient (δ) is taken,
| Eelec = δEtribo ≈ Pt | (11) |
![]() | (12) |
![]() | (13) |
715 µW cm−2 for sliding velocities between 1 and 100 nm s−1, and ΔEmax for the case where the interlayer gap is decreased to 2.84 Å, supposing complete conversion of tribological energy to electrical energy. Experimental findings13 indicate that a tribo-piezoelectric generator has a power density of 7 µW cm−2. According to earlier studies, the power density output of triboelectric nanogenerators82 and piezoelectric nanogenerators ranges79 from 400 to 50
000 µW cm−2 and 4.41 to 5.92 µW cm−2, respectively. A reasonable supreme power density and a significant induced electrical output voltage are seen by the predicted CdS bilayer nanogenerator. This is comparable to the triboelectric nanogenerators of GaN nanowire,83 Janus TMD bilayers,69 h-BN polypropylene,84 monolayer MoS2 flake,29 PVDF–TrFE, and graphene oxide,79 which are presented in Table 2.
| Offered material | Highest power density (mW cm−2) | Highest induced output voltage (V) | ||
|---|---|---|---|---|
| Theoretical | Experimental | Theoretical | Experimental | |
| Planned bilayer 2D CdS | ∼52 | 0.35 | ||
| Double layer of 2D BeO85 | ∼30 | 0.22 | ||
| Double layer of 2D h-BN42 | ∼25 | 0.172 | ||
| Double layer of 2D InN86 | ∼74 | 0.20 | ||
| Nanowire array of GaN83 | — | — | 0.15–0.35 | |
| Janus TMD double layers69 | ∼29.64 | 0.25 | ||
| Monolayer MoS2 flake29 | — | ∼2 × 10−4 | 0.01825 | 0.015 |
| h-BN polypropylene84 | — | ∼29 × 10−4 | 2.3 | |
| Bilayer PVDF–TrFE and graphene oxide79 | 4.41 × 10−3 | 4 | ||
Furthermore, we expanded our investigation by taking into account a 3 × 3 CdS supercell bilayer structure, which yields a power density of around 519.16 to 51
916 µW cm−2, negligibly higher than the values obtained with the CdS unit cells. The supercell structure has an induced voltage of 0.233 V and a polarization of 1.735 pC m−1, whereas the unit cell structure exhibits values of 0.231 V and 1.73 pC m−1, respectively. Hence, the consistency between the unit cell and supercell results highlights the periodicity effects and affirms the reliability of our DFT first-principles calculations. Regarded as a potential choice for energy harvesting applications, the CdS nanogenerator shows promise, particularly for low-power devices and self-powered sensors.87 Recently, there have been enormous advancements and research efforts to create CdS-based devices, such as water-splitters,43,44,63 photovoltaics,45 nanowires, and thin-film-based piezo-electric transducers,46 photodetectors,88–90 and nano thin-film transistors.91 Additionally, CdS-based nanowires and other structures are constantly improving.12 Therefore, it is anticipated that in the near future, the experimental construction of a 2D CdS-based tribo-piezoelectric nanogenerator will be viable and sustainable.
715 µW cm−2 by modulating the tension on the probe tip. The results of this study indicate that CdS bilayers have the potential to be employed in the development of a highly efficient, innovative, and stable nanogenerator that can translate and harvest energy at the nanoscale.
| This journal is © The Royal Society of Chemistry 2025 |