Next Article in Journal
Investigation on the Interfaces in Organic Devices by Photoemission Spectroscopy
Previous Article in Journal
Polarization-Independent Broadband Infrared Selective Absorber Based on Multilayer Thin Film
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Strain Engineering of Anisotropic Electronic, Transport, and Photoelectric Properties in Monolayer Sn2Se2P4

School of Microelectronics and Control Engineering, Changzhou University, Changzhou 213164, China
*
Author to whom correspondence should be addressed.
Nanomaterials 2025, 15(9), 679; https://doi.org/10.3390/nano15090679
Submission received: 24 March 2025 / Revised: 22 April 2025 / Accepted: 24 April 2025 / Published: 30 April 2025
(This article belongs to the Section Nanoelectronics, Nanosensors and Devices)

Abstract

:
In this study, we demonstrate that the Sn2Se2P4 monolayer exhibits intrinsic anisotropic electronic characteristics with the strain-synergistic modulation of carrier transport and optoelectronic properties, as revealed by various levels of density functional theory calculations combined with the non-equilibrium Green’s function method. The calculations reveal that a-axis uniaxial compression of the Sn2Se2P4 monolayer induces an indirect-to-direct bandgap transition (from 1.73 eV to 0.97 eV, as calculated by HSE06), reduces the hole effective mass by ≥70%, and amplifies current density by 684%. Conversely, a-axis uniaxial expansion (+8%) boosts ballistic transport (a/b-axis current ratio > 105), rivaling black phosphorus. Notably, a striking negative differential conductance arises with the maximum Ipeak/Ivalley in the order of 105 under the 2% uniaxial compression along the b-axis of the Sn2Se2P4 monolayer. Visible-range anisotropic absorption coefficients (~105 cm−1) are achieved, where −4% a-axis strain elevates the photocurrent density (6.27 μA mm−2 at 2.45 eV) and external quantum efficiency (39.2%) beyond many 2D materials benchmarks. Non-monotonic strain-dependent photocurrent density peaks at 2.00 eV correlate with hole effective mass reduction patterns, confirming the carrier mobility of the Sn2Se2P4 monolayer as the governing parameter for photogenerated charge separation. These results establish Sn2Se2P4 as a multifunctional material enabling strain-tailored anisotropy for logic transistors, negative differential resistors, and photovoltaic devices, while guiding future investigations on environmental stabilization and heterostructure integration toward practical applications.

1. Introduction

The rapid development of integrated circuits has driven the continuous demand for performance enhancement in electronic devices, positioning two-dimensional (2D) semiconductors as promising candidates for applications in field-effect transistors (FETs) and photoelectric conversion systems [1,2,3,4,5,6,7,8]. However, critical limitations persist in existing 2D materials: transition metal dichalcogenides (TMDs) like MoS2 and ReS2 suffer from restricted carrier mobility [9,10,11,12], while high-mobility phosphorenes (e.g., black phosphorus and violet phosphorus) exhibit poor environmental stability due to oxidation susceptibility [13,14,15]. Furthermore, the narrow band gap of materials such as silicene and germanene increases heat loss and reduces energy conversion efficiency [16,17]. To address these challenges, it is essential not only to explore novel 2D materials through density functional theory (DFT) computational simulations but also to develop property-modification strategies like strain engineering, doping, and van der Waals heterostructure construction [18,19].
As a particularly effective approach, strain engineering enables the precise modulation of electronic structures in 2D materials owing to their exceptional mechanical flexibility. Appropriate mechanical strain enables the precise modulation of bandgap magnitude, bandgap type, and band curvature in two-dimensional materials. This strain engineering strategy allows the bandgap range to be tuned to 0.77–1.19 eV for optimal channel material performance in electronic devices or 1.1–1.6 eV for photovoltaic applications [20,21,22]. Hence, the controlled manipulation of band structures holds promise for enhancing electronic transport and optoelectronic characteristics in functional devices. This capability has been successfully demonstrated in graphene, hexagonal boron nitride (hBN), TMDs, and MXenes, opening new avenues for designing high-performance nanodevices [23,24,25,26,27].
Recent attention has been focused on a novel family of ternary 2D materials with puckered honeycomb structures—triphosphides and triarsenides (general formula AX3, where A = P/As and X = group II/XIII/XIV/XV elements) [28,29,30,31]. Two-dimensional triarsenides (CaAs3, GeAs3) demonstrate exceptional carrier mobility exceeding ~3 × 104 cm2 V−1 s−1 [28,32], while their triphosphide counterparts (GeP3, SnP3) are renowned for their high optical absorption coefficients (~105 cm−1) across the visible-to-infrared spectrum [30,33]. To expand this material family beyond conventional stoichiometries, recent theoretical studies have proposed structural derivatives through unit cell doubling and selective chalcogen substitution, leading to the prediction of Ge2S2P4 and Sn2S2P4 monolayers with exceptional photocatalytic water-splitting capabilities [34,35].
Particularly noteworthy is the high-performance two-dimensional monolayer (ML) Sn2Se2P4 predicted by Trung et al., exhibiting phonon-dispersion-validated dynamic stability and ultra-high anisotropic mobility properties [36]. An ab initio molecular dynamics (AIMD) simulation was employed to heat the Sn2Se2P4 monolayer at 500 K for 10 ps, showing sufficient thermal stability. As determined by the Perdew–Burke–Ernzerhof (PBE) calculations, the Sn2Se2P4 monolayer exhibits an indirect bandgap of 1.04 eV, which decreases and transitions to a direct bandgap under biaxial compressive strain. Remarkably, this material demonstrates an outstanding carrier transport performance, with electron and hole mobilities reaching 5.4 × 103 cm2 V−1 s−1 and an extraordinary 7.4 × 104 cm2 V−1 s−1, respectively. These superior properties position Sn2Se2P4 as a promising candidate for next-generation electronic and optoelectronic devices. While existing studies have primarily investigated the effects of biaxial strain on this material, the comprehensive modulation of electrical transport and optoelectronic properties in Sn2Se2P4 through uniaxial strain—a more flexible and controllable tuning strategy—remain systematically underexplored. Notably, whether uniaxial strain can effectively modulate the bandgap of Sn2Se2P4 into the optimal range for optoelectronic devices while preserving or even enhancing its exceptional carrier mobility remains an open question. Motivated by the pivotal role of strain engineering in optimizing the performance of two-dimensional (2D) materials, we systematically investigate the electronic structure, carrier transport properties, and optoelectronics characteristics of Sn2Se2P4 monolayers under uniaxial strain, aiming to establish a strain-dependent framework for multifunctional device design.
In this work, we systematically investigate the strain-tunable anisotropic properties of orthorhombic Sn2Se2P4 monolayers through first-principles calculations. Our density functional theory (DFT) analysis reveals the following: (1) direction-dependent bandgap engineering from indirect to direct transitions under uniaxial compression along the a-axis; (2) strain-modulated effective mass reductions (holes under a-axis compression) correlated with enhanced transport properties; (3) non-equilibrium Green’s function (NEGF) calculations demonstrating anisotropic current–voltage characteristics with current ratios exceeding 105 and strain-enhanced negative differential conductance (NDC) effects; and (4) optoelectronic property optimization under specific strain conditions, achieving an external quantum efficiency (39.2%) and photocurrent density (6.27 μA mm−2) surpassing conventional photovoltaic materials. These findings establish Sn2Se2P4 as a versatile platform for developing strain-engineered nanoelectronic and optoelectronic devices.

2. Calculational Methods

2.1. Details of Structural Optimization and Electronic Structure

The geometric optimization and electronic properties of the Sn2Se2P4 monolayer were calculated using density functional theory (DFT) implemented in the QuantumATK software (ver. S-2021.06) [37,38]. The Perdew–Burke–Ernzerhof (PBE) formulation of the generalized gradient approximation (GGA) was employed in conjunction with the linear combination of atomic orbitals (LCAO) method [39]. The PseuDodojo pseudopotential was adopted to replace the atomic all-electron potential [40], while the medium-precision numerical basis was used for the wave function expansion. A real-space density mesh cutoff of 105 Hartree was applied, and the first Brillouin zone of Sn2Se2P4 was sampled with a 4 × 3 × 1 k-point grid. Structural optimization convergence was achieved when the maximum force on each atom decreased below 0.001 eV Å−1. A vacuum spacing of 25 Å was implemented along the non-periodic direction to eliminate interlayer interactions. To enhance the accuracy of bandgap results, the Heyd–Scuseria–Ernzerhof (HSE06) hybrid functional was subsequently employed [41].

2.2. The Calculational Method for Carrier Transport Properties

We employed the NEGF method based on DFT to calculate the carrier transport properties [42,43]. In DFT calculations, again, the PBE functional with the LCAO method and PseuDodojo pseudopotentials were utilized for calculating the carrier transport. For current–voltage (IV) characteristic calculations, Brillouin zone sampling was implemented with k-grid configurations of 1 × 3 × 132 and 4 × 1 × 76 along the a-axis and b-axis directions, respectively. The IV characteristics of the two-probe system were subsequently derived through Landauer–Büttiker formalism [44]:
I V b i a s = 2 e h T E , ε L , ε R × f R E , ε R f L E , ε L d E
where V b i a s represents the bias voltage added to both sides of the electrodes and is defined as e V b i a s = ε R ε L . The Fermi energy levels of the left and right electrodes are denoted as ε L and ε R , respectively. E represents the electrons. T is the carrier transport coefficient, and the Fermi Dirac distribution of the left and right electrodes are denoted as f L ( E , ε L ) and f R ( E , ε R ) , respectively.

2.3. Calculational Method for Photocurrent

The photocurrent density (Jph) was calculated through the integration of NEGF methodology within the first-order perturbation theory framework, based on the first-born approximation [45,46]. The perturbation induced by electron–photon interactions was defined through the Hamiltonian, expressed as:
H ^ = H ^ 0 + e m 0 A · p ^
where H ^ 0 is the Hamiltonian of the two-probe device, e denotes the electron charge, m 0 represents the free electron mass, p ^ is the momentum operator, and A is defined as the electromagnetic vector potential. The transmission coefficients can be obtained using [47]:
T α E = T r i Γ α 1 f α G p h < + f α G p h >
Tr denotes the trace operator. G p h < and G p h > are the lesser and greater Keldysh Green’s functions approximate to the first order, while Γ α and f α are the line width function and Fermi distribution for the α (left or right) electrode. Jph is formulated as [48,49]:
J p h = e S d E 2 π α T α E
Furthermore, the functional, basis, pseudopotentials, and k-point grids employed in the photocurrent calculations remained identical to those adopted in carrier transport simulations.

3. Results and Discussion

3.1. Structural and Electronic Properties of Sn2Se2P4

The Sn2Se2P4 monolayer was constructed by substituting two P atoms in the SnP3 hexagonal unit cell with Se atoms [33]. Subsequent geometric optimization using the Vienna ab initio simulation package (VASP) refined the atomic coordinates with an energy convergence threshold of 1 × 10−6 eV. Figure 1a,b present the top and side views of the optimized 2D Sn2Se2P4 monolayer, respectively. To facilitate carrier mobility calculations, the hexagonal primitive cell was transformed into an orthogonal unit cell, demarcated by bronze and purple dashed lines. The optimized monolayer exhibited a thickness of d = 2.30 Å and lattice constants of a = 7.17 Å and b = 12.42 Å, aligning closely with the theoretical predictions by Trung et al. Prior computational evaluations established its thermal, dynamic, and mechanical stability [36]. The electron localization function (ELF) analysis elucidated the bonding characteristics through the spatial mapping of the electron pair distribution. As depicted in Figure 1c, the electrons are localized around the P and Se atoms, thereby revealing that Se–P and P–P bonds exhibit typical covalent features, while Sn–P and Sn–Se bonds demonstrate ionic character.
The electronic band structure of the Sn2Se2P4 monolayer calculated using the PBE functional exhibited indirect bandgap semiconductor characteristics with a computed bandgap of 1.02 eV (Figure 1d). To address the well-known bandgap underestimation inherent in the PBE method, we employed the HSE06 hybrid functional, resulting in a significant increase in the bandgap to 1.73 eV (Figure 1e), which aligned with previously reported data [36]. Spin–orbit coupling (SOC) effects were investigated through PBE + SOC and HSE06 + SOC band structure calculations (Figure S1), revealing a neglectable difference in the band gap values. Projected density of states (PDOS) analysis (Figure 1d,e) demonstrated that the conduction band minimum (CBM) arose from the hybridized contributions of Se, Sn, and P atomic orbitals, while the valence band maximum (VBM) was predominantly governed by P orbital states.
To systematically investigate the strain-dependent modulation of electronic structures, uniaxial strains ranging from −8% to +8% were applied along the a- and b-axis of the Sn2Se2P4 monolayer (Figure 2a). HSE06-calculated band evolution under uniaxial strains (Figure 2b,c) revealed that the indirect bandgap characteristic persisted under b-axis strain, whereas compressive a-axis strain induced CBM-VBM degeneracy at the G-point of the Brillouin zone, triggering an indirect-to-direct bandgap transition. Concurrently, the bandgap decreased monotonically from 1.73 eV to 0.97 eV with increasing compressive strain. The strain-dependent shifts in band edge positions and HSE06-calculated band gaps along both axes are illustrated in Figure 2d,e. The redox potentials given by the Nernst equation (ESI†) for photocatalytic water splitting at pH = 0 (purple dashed) and pH = 7 (green dashed) are annotated [50], with ECBM lying below the reduction potential and EVBM above the oxidation potential, satisfying photocatalytic water splitting requirements. Notably, structures with b-axis strains of −4% to +8% optimally meet the pH = 0 criteria, while those between −4% and −8% align with pH = 7 conditions (see ESI†), demonstrating strain engineering’s precision in tailoring photocatalytic performance.
Figure 2f further reveals the strain-dependent modulation of carrier effective masses. The electron effective mass exhibits weak strain dependence along both the a-axis (0.25–0.36 m e * ) and b-axis (0.46–0.71 m e * ). In contrast, the hole effective masses show prominent anisotropic responses: along the a-axis, the hole effective mass abruptly increases to 1.98 m e * under +2% tensile strain but sharply decreases to 0.14 m e * at −2% compressive strain. Conversely, b-axis hole effective masses remain within 1.23–2.51 m e * across all strain conditions. Notably, the significant reduction in hole effective mass (<1 m e * ) within the a-axis compressive strain range of −2% to −8% predicts an order-of-magnitude enhancement in hole mobility, providing a theoretical foundation for optimizing carrier transport properties via strain engineering.

3.2. Anisotropic I–V Characteristics

The pronounced anisotropic carrier mobility in the Sn2Se2P4 monolayer—with a-axis hole mobility ( μ h a   = 7.4 × 104 cm2 V−1·s−1, ~300 times higher than electron mobility) and b-axis electron mobility ( μ e b = 5.4 × 103 cm2 V−1·s−1, ~62 times greater than hole mobility)—motivated our design of the two-probe configurations to reveal the intrinsic transport behavior (Figure 3a,b). To eliminate metal–semiconductor contact barriers, homogeneous heavily doped electrodes were implemented. The hole transport channel along the a-axis was engineered by elevating the electrode’s VBM 0.15 eV above the Fermi level, achieving a hole density of 1014 cm−2. Conversely, the electron transport channel along the b-axis was configured by lowering the CBM 0.15 eV below the Fermi level, corresponding to an electron density of 1014 cm−2.
The unstrained-state (ε = 0%) transport properties of Sn2Se2P4 monolayer exhibited prominent directional dependence (Figure 4a). In the |V| ≤ 1.0 V bias range, the a-axis hole transport demonstrated a linear current–voltage response, achieving Ia = 1.57 × 102 nA at V = 1.0 V. In contrast, the b-axis electron transport showed abnormal negative differential conductance (NDC) near a 0.4 V bias (Figure 4b), where the current abruptly dropped to 4.0 × 10−5 nA, creating a current ratio of 10⁵ versus the a-axis transport. This phenomenon arose from the b-direction effective barrier height elevation (Φb = 0.33 eV, 0.12 eV higher than the a-direction effective barrier height Φa) that impeded electron transport, according to the projected local density of states (PLDOS) plots (Figure 4c,d). Consequently, focusing on this distinctive −0.6 V to 0.6 V bias window, we explored strain-modulated IV characteristics under uniaxial strain engineering.

3.3. Uniaxial Strain-Modulated I–V Characteristics

3.3.1. a-Axis Strain-Modulated I–V Characteristics

By applying uniaxial a-axis strain (−8% ≤ εa ≤ +8%) to the Sn2Se2P4 monolayer, we observed a strong correlation between the magnitude of a-axis strain (tensile/compressive) and the anisotropic carrier transport properties (Figure 5 and Figure 6). Under a-axis tensile strain (+2% ≤ εa ≤ +8%), the current in the low-bias regime (|V| < 0.4 V) decreased by 90% compared to the unstrained structure, while a strain-enhanced effect emerged at 0.4 V < |V| < 0.6 V: the current Ia increased by 421% as εa rose from +2% to +8% (Figure 5a). According to Figure 5c–e, PLDOS analysis revealed that tensile strain reduced the effective barrier height of the a-axis transport channel to Φa = 0.12 eV at εa = +8%, marking a 20% reduction from the unstrained value (Φa = 0.15 eV). Notably, strain simultaneously enhanced transport anisotropy, achieving an a/b-axis current ratio of 105 at V = 0.6 V for εa = +8% (Figure 5b). This ratio matched the switching performance of black phosphorus (105), fulfilling the ON/OFF ratio requirements for field-effect transistors (FETs) [51]. The phenomenon aligned with an 80% reduction in the a-axis effective barrier height (Φa = 0.12 eV versus Φb = 0.62 eV) at V = 0.6 V under εa = +8% (Figure 5f).
In contrast, compressive a-axis strain (εa = −6% and −8%) significantly enhanced the carrier transport efficiency of the Sn2Se2P4 monolayer (Figure 6a). At εa = −8%, the current density Ia reached 3.53 × 103 nA, representing a two-order-of-magnitude enhancement over the unstrained structure. First-principles calculations revealed that this nonlinear response originated from strain-induced band reconstruction: compressive strain triggered a transition from an indirect to a direct bandgap (Eg decreased from 1.73 eV to 0.97 eV), accompanied by a 0.43 eV downward shift in the CBM and a 0.38 eV upward shift in the VBM toward the Fermi level (Figure 2b). PLDOS calculations further confirmed that, compared to the unstrained structure (Figure 5c), compressive strains of εa = −6% and −8% reduced the a-axis effective barrier height to Φa < 0.1 eV (Figure 6b,c), thereby markedly facilitating hole tunneling transport.

3.3.2. b-Axis Strain-Modulated I–V Characteristics and NDC Effects

The transport properties of the Sn2Se2P4 monolayer under b-axis strain exhibit unique NDC modulation in 0.1–0.4 V (−0.4 to −0.1 V) bias regimes. Under b-axis strain (εb = 0% and −2%), the transport current initially rises within 0–0.4 V bias, abruptly decreases, and subsequently recovers beyond 0.4 V (Figure 7a,c). The calculated logarithmic-scale IV characteristics (Figure 7b,d) reveal the pronounced NDC behavior in both the unstrained structure and εb = −2% strained case within 0.1–0.4 V, achieving a peak-to-valley current ratio (Ipeak/Ivalley) in the order of 105 (Figure 8a,d). This ratio surpasses the reported NDC effects in graphene (50–200) [52], phosphorene (25) [53], phosphorene/ReS2 heterostructures (4.2–6.9) [15], and MoS2/WSe2 heterostructures (103) [54]. PLDOS analysis demonstrates that the effective barrier height Φb at V = 0.4 V (Figure 8c,f) is significantly higher than at V = 0.1 V (Figure 8b,e) for both cases, leading to suppressed carrier transport. This disparity in carrier transport capacity between the peak and valley states directly induces the pronounced NDC effect.

3.4. Photocurrent Transport Properties

To investigate the optoelectronic performance of the Sn2Se2P4 monolayer in practical device environments, we construct optical nanodevices along the a- and b-axes, leveraging its intrinsic anisotropy. As shown in Figure 9a,b, the device features an intrinsic Sn2Se2P4 scattering region (length set as 50 Å) for linearly polarized light irradiation, connected to heavily doped p-type (source) and n-type (drain) Sn2Se2P4 semi-infinite electrodes. While the PBE functional inherently underestimates band gaps, high-concentration doping strategies align the effective band gaps with theoretical predictions, ensuring computational reliability. Building on our prior discovery that a-axis compressive strain (−8% ≤ εa ≤ −2%) induces indirect-to-direct bandgap transitions, we further investigate its modulation effects on optoelectronic performance.
As shown in Figure 10a, the optical absorption coefficient of the Sn2Se2P4 monolayer exhibits pronounced anisotropy across the visible spectrum (1.61–3.11 eV). The unstrained structure (εa = 0%) achieves an a-axis absorption coefficient of ~105 cm−1, comparable to that of the same family of ternary 2D monolayers, including Sn2Te2P(As)4 (104–105 cm−1) [55], Ge2S2P(As)4 (about 105 cm−1) [34], SnGeS2As4 (105–106 cm−1) [56], and Sn2S2P4 (about 105 cm−1) [35]. Under a-axis compressive strain (−8% ≤ εa ≤ −2%), absorption is further enhanced, accompanied by distinct spectral separation between the a-axis and b-axis absorption profiles, indicating strain-induced anisotropy enhancement. This phenomenon aligns with strain-modulated carrier transport properties, confirming the synergistic control of optoelectronic and transport properties.
Photocurrent density (Jph), as a critical metric for optoelectronic device performance, quantifies the photogenerated current per unit area under illumination, reflecting the material’s light absorption and conversion efficiency. By computing the Jph, we demonstrate that strain engineering enhances the photovoltaic conversion efficiency of the Sn2Se2P4 monolayer. At εa = −4%, the a-axis device achieves a Jph peak of 6.27 μA mm−2 at 2.45 eV (Figure 10b), representing a 684% enhancement over the unstrained state (0.80 μA mm−2) and surpassing the InSe (0.018 μA mm−2) and NaCuTe (1.68 μA mm−2) counterparts [57,58]. In contrast, the b-axis device under the same strain conditions only yields 2.51 μA mm−2 (Figure 10c), highlighting the a-axis as the preferred photoelectric channel. Notably, the Jph peak around 2.00 eV initially rises sharply from 0.80 μA mm−2a = 0%) to 5.59 μA mm−2a = −2%), then declines to 0.89 μA mm−2a = −8%). This trend aligns with strain-dependent hole effective mass variations (Figure 2f), suggesting carrier transport limitations under extreme strains. Overall, the εa = −2% and εa = −4% strains significantly enhance Jph while preserving the anisotropy of photocurrent density.
Furthermore, photoresponsivity (Rph) and external quantum efficiency (EQE) also serve as critical metrics for evaluating optoelectronic device performance. Their expressions are defined as follows [59]:
R p h = J p h I ω E
E Q E = R p h h c e λ
where E is the photon energy, Iω represents the photon flux, h denotes Planck’s constant, c is the speed of light, e is the electron charge, and λ represents wavelength.
We investigate the Rph and EQE of the a-axis and b-axis devices in the Sn2Se2P4 monolayer, applying the same methodology as used for Jph (Figures S2 and S3). Given the deterministic influence of photocurrent density on Rph and EQE, their strain-dependent trends under a-axis compressive strain align with the photocurrent density behavior. Specifically, an optimization threshold exists within the strain modulation window: Rph and EQE peak at εa = −2% and εa = −4%, with performance gains diminishing outside this range. Thus, our analysis focuses on devices under εa = −2% and εa = −4%.
As shown in Figure S2a, the calculated Rph reaches 0.16 A/W at 2.45 eV (εa = −4%) and 0.17 A/W at 2.00 eV (εa = −2%), surpassing the reported values for MoS2 (0.016 A/W) and graphene (5 × 10−4 A/W) [60]. Similarly, Figure S3a shows EQE values of 39.2% at 2.45 eV (εa = −4%) and 34.9% at 2.00 eV (εa = −2%), exceeding those of SnS (22.01%) [61], NaCuTe (34.3%) [58], NaCuSe (8%) [58], and KAgSe (17.9%) [62]. Consistent with the anisotropy of photocurrent density, the Rph and EQE retain pronounced anisotropy (Figure 10e,f), with significant enhancements over the unstrained counterpart.
In this part, we demonstrate that the Sn2Se2P4 monolayer under moderate a-axis compressive strains (εa = −2% and εa = −4%) simultaneously achieves enhanced light absorption, tunable optoelectronic properties, and optimized carrier transport. Its integrated performance surpasses benchmark 2D optoelectronic materials (e.g., MoS2, SnS) and establishes a novel material platform for designing spectrally selective and directionally sensitive integrated optoelectronic devices.

4. Conclusions

This study systematically unveils the anisotropic electronic, transport, and optoelectronic properties of Sn2Se2P4 monolayers under strain engineering through DFT and NEGF methodologies. The pristine Sn2Se2P4 monolayer exhibits an indirect bandgap of 1.73 eV (HSE06). Under −8% a-axis compressive strain, the bandgap monotonically decreases to 0.97 eV with concurrent indirect-to-direct transition and drastic hole effective mass reduction ( m h * : 1.79→0.13 m e * ), thereby surmounting the intrinsic mobility limitations caused by the high effective mass in conventional TMDs. Notably, ballistic transport simulations (30 Å channel length) of the Sn2Se2P4 monolayer reveal an a/b-axis current ratio reaching ~105 at 0.4 V under ±1 V bias, comparable to black phosphorus, with sustained performance under 6–8% a-axis tensile strain. b-axis compression (0–2%) induces pronounced NDC effects (Ipeak/Ivalley ≈ 105). a-axis strain engineering for the Sn2Se2P4 monolayer induces strong anisotropy in optoelectronic responses, including optical absorption coefficients (>105 cm−1), photocurrent density (Jph), responsivity (Rph), and external quantum efficiency (EQE). At −4% a-axis strain under 2.45 eV photon energy, Jph peak reaches 6.27 μA mm−2 (684% enhancement versus unstrained state) with an EQE of 39.2%, significantly surpassing benchmark materials like NaCuTe (1.68 μA mm−2, 39.4%). Most compellingly, the non-monotonic Jph peak evolution at 2.00 eV under a-axis compression (−8%→0%) correlates strongly with strain-dependent m h * variations, demonstrating carrier mobility as the dominant factor in photocarrier extraction for Sn2Se2P4 monolayers. These findings demonstrate that the Sn2Se2P4 monolayer achieves the synergistic modulation of electronic–optoelectronic properties through uniaxial strain engineering, with its tailorable anisotropy establishing an ideal platform for developing multifunctional nano-devices in the post-Moore era. Future investigations should prioritize environmental stability enhancement and 2D heterostructure integration to expedite experimental validation and device implementation.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/nano15090679/s1, Figure S1: The calculated electronic band structure of the Sn2Se2P4 monolayer with spin–orbital coupling (SOC) (orange dash curves), and without SOC (blue curves), extracted from (a) the PBE method and (b) HSE method; Figure S2: The calculated Rph of the Sn2Se2P4 monolayer along (a) the a-axis and (b) b-axis under different compressive strains; Figure S3: The calculated EQE of the Sn2Se2P4 monolayer along (a) the a-axis and (b) b-axis under different compressive strains. Ref. [50] is cited in the Supplementary Materials.

Author Contributions

Conceptualization, H.X. and Y.X.; Methodology, H.X. and Y.X.; Validation, H.X. and Y.X.; Formal analysis, H.X. and Y.X.; Investigation, H.X. and Y.X.; Data curation, H.X.; Writing—original draft, H.X.; Writing—review and editing, H.X. and Y.X.; Visualization, H.X.; Supervision, Y.X.; Project administration, Y.X. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Acknowledgments

Calculations were carried out in the High-Performance Computing Laboratory of Changzhou University and Hefei Advanced Computing Center.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Feng, X.; Cheng, R.; Yin, L.; Wen, Y.; Jiang, J.; He, J. Two-Dimensional Oxide Crystals for Device Applications: Challenges and Opportunities. Adv. Mater. 2024, 36, 2304708. [Google Scholar] [CrossRef] [PubMed]
  2. Hu, J.; Dong, M. Recent advances in two-dimensional nanomaterials for sustainable wearable electronic devices. J. Nanobiotechnol. 2024, 22, 63. [Google Scholar] [CrossRef] [PubMed]
  3. Jiang, J.; Xu, L.; Qiu, C.; Peng, L.-M. Ballistic two-dimensional InSe transistors. Nature 2023, 616, 470–475. [Google Scholar] [CrossRef] [PubMed]
  4. Joseph, S.; Mohan, J.; Lakshmy, S.; Thomas, S.; Chakraborty, B.; Thomas, S.; Kalarikkal, N. A review of the synthesis, properties, and applications of 2D transition metal dichalcogenides and their heterostructures. Mater. Chem. Phys. 2023, 297, 127332. [Google Scholar] [CrossRef]
  5. Qu, J.; Cheng, H.; Lan, H.; Zheng, B.; Luo, Z.; Yang, X.; Yi, X.; Wu, G.; Chen, S.; Pan, A. Space-Confined Growth of Ultrathin P-Type GeTe Nanosheets for Broadband Photodetectors. Small 2024, 20, 2309391. [Google Scholar] [CrossRef]
  6. Tang, X.; Hao, Q.; Hou, X.; Lan, L.; Li, M.; Yao, L.; Zhao, X.; Ni, Z.; Fan, X.; Qiu, T. Exploring and Engineering 2D Transition Metal Dichalcogenides toward Ultimate SERS Performance. Adv. Mater. 2024, 36, 2312348. [Google Scholar] [CrossRef]
  7. Wu, Y.; Yuan, W.; Xu, M.; Bai, S.; Chen, Y.; Tang, Z.; Wang, C.; Yang, Y.; Zhang, X.; Yuan, Y.; et al. Two-dimensional black phosphorus: Properties, fabrication and application for flexible supercapacitors. Chem. Eng. J. 2021, 412, 128744. [Google Scholar] [CrossRef]
  8. Xu, Y.; Li, D.; Sun, H.; Xu, H.; Li, P. Comprehensive understanding of electron mobility and superior performance in sub-10 nm DG ML tetrahex-GeC2 n-type MOSFETs. Phys. Chem. Chem. Phys. 2024, 26, 4284–4297. [Google Scholar] [CrossRef]
  9. Baglov, A.; Khoroshko, L.; Zhoidzik, A.; Dong, M.; Weng, Q.; Kazi, M.; Khandaker, M.U.; Islam, M.A.; Chowdhury, Z.Z.; Sayyed, M.I.; et al. Evolution of structural and electronic properties standardized description in rhenium disulfide at the bulk-monolayer transition. Heliyon 2024, 10, e28646. [Google Scholar] [CrossRef]
  10. Yamusa, S.A.; Shaari, A.; Alsaif, N.A.M.; Alsalamah, I.M.; Isah, I.; Rekik, N. Elucidating the Structural, Electronic, Elastic, and Optical Properties of Bulk and Monolayer MoS2 Transition-Metal Dichalcogenides: A DFT Approach. ACS Omega 2022, 7, 45719–45731. [Google Scholar] [CrossRef]
  11. Javaid, M.; Drumm, D.W.; Russo, S.P.; Greentree, A.D. A study of size-dependent properties of MoS2 monolayer nanoflakes using density-functional theory. Sci. Rep. 2017, 7, 9775. [Google Scholar] [CrossRef] [PubMed]
  12. Aparna, M.P.; Chatanathodi, R. Density Functional Theory Calculations of the Stability, Electronic Structure, and Magnetism of ReS2 Nanoribbons: An Emerging Material for Electrocatalytic Reactions. ACS Appl. Nano Mater. 2022, 5, 2385–2394. [Google Scholar] [CrossRef]
  13. Xu, Y.; Dai, J.; Zeng, X.C. Electron-Transport Properties of Few-Layer Black Phosphorus. J. Phys. Chem. Lett. 2015, 6, 1996–2002. [Google Scholar] [CrossRef] [PubMed]
  14. Xu, Y.; Long, Q.; Li, D.; Li, P. 2D Violet phosphorene with highly anisotropic mobility and its vdW heterojunction design for device applications. Phys. Chem. Chem. Phys. 2022, 24, 3379–3385. [Google Scholar] [CrossRef] [PubMed]
  15. Shim, J.; Oh, S.; Kang, D.-H.; Jo, S.-H.; Ali, M.H.; Choi, W.-Y.; Heo, K.; Jeon, J.; Lee, S.; Kim, M.; et al. Phosphorene/rhenium disulfide heterojunction-based negative differential resistance device for multi-valued logic. Nat. Commun. 2016, 7, 13413. [Google Scholar] [CrossRef]
  16. Kaneko, S.; Tsuchiya, H.; Kamakura, Y.; Mori, N.; Ogawa, M. Theoretical performance estimation of silicene, germanene, and graphene nanoribbon field-effect transistors under ballistic transport. Appl. Phys. Express 2014, 7, 035102. [Google Scholar] [CrossRef]
  17. Devi, P.K.; Singh, K.K. Electronic and work function-based glucose sensors on graphene, silicene, and germanene sheets—DFT studies. Comput. Theor. Chem. 2023, 1225, 114134. [Google Scholar] [CrossRef]
  18. Gueorguiev, G.K.; Stafström, S.; Hultman, L. Nano-wire formation by self-assembly of silicon–metal cage-like molecules. Chem. Phys. Lett. 2008, 458, 170–174. [Google Scholar] [CrossRef]
  19. Filho, M.A.M.; Farmer, W.; Hsiao, C.-L.; dos Santos, R.B.; Hultman, L.; Birch, J.; Ankit, K.; Gueorguiev, G.K. Density Functional Theory-Fed Phase Field Model for Semiconductor Nanostructures: The Case of Self-Induced Core–Shell InAlN Nanorods. Cryst. Growth Des. 2024, 24, 4717–4727. [Google Scholar] [CrossRef]
  20. Shockley, W.; Queisser, H.J. Detailed Balance Limit of Efficiency of p-n Junction Solar Cells. J. Appl. Phys. 1961, 32, 510–519. [Google Scholar] [CrossRef]
  21. Park, N.-G.; Segawa, H. Research Direction toward Theoretical Efficiency in Perovskite Solar Cells. ACS Photonics 2018, 5, 2970–2977. [Google Scholar] [CrossRef]
  22. Rühle, S. Tabulated values of the Shockley–Queisser limit for single junction solar cells. Sol. Energy 2016, 130, 139–147. [Google Scholar] [CrossRef]
  23. Xu, Z.-H.; Jiang, J.; Wang, Z.; Wei, K.; Liu, G.; Ke, S.-H. Tunable electronic and optoelectronic characteristics of two-dimensional g-GeC monolayer: A first-principles study. Opt. Express 2024, 32, 40539–40553. [Google Scholar] [CrossRef]
  24. Liao, L.; Lin, Y.-C.; Bao, M.; Cheng, R.; Bai, J.; Liu, Y.; Qu, Y.; Wang, K.L.; Huang, Y.; Duan, X. High-speed graphene transistors with a self-aligned nanowire gate. Nature 2010, 467, 305–308. [Google Scholar] [CrossRef]
  25. Ares, P.; Cea, T.; Holwill, M.; Wang, Y.B.; Roldán, R.; Guinea, F.; Andreeva, D.V.; Fumagalli, L.; Novoselov, K.S.; Woods, C.R. Piezoelectricity in Monolayer Hexagonal Boron Nitride. Adv. Mater. 2020, 32, 1905504. [Google Scholar] [CrossRef]
  26. Li, P.; Wu, W.; Xu, Y.; Liu, J.; Wu, S.; Ye, Y.; Liang, C.; Zeng, X.C. Two-Dimensional IV–V Monolayers with Highly Anisotropic Carrier Mobility and Electric Transport Properties. J. Phys. Chem. Lett. 2021, 12, 1058–1065. [Google Scholar] [CrossRef] [PubMed]
  27. Conley, H.J.; Wang, B.; Ziegler, J.I.; Haglund, R.F., Jr.; Pantelides, S.T.; Bolotin, K.I. Bandgap Engineering of Strained Monolayer and Bilayer MoS2. Nano Lett. 2013, 13, 3626–3630. [Google Scholar] [CrossRef]
  28. Laurien, M.; Saini, H.; Rubel, O. Band alignment of monolayer CaP3, CaAs3, BaAs3 and the role of p–d orbital interactions in the formation of conduction band minima. Phys. Chem. Chem. Phys. 2021, 23, 7418–7425. [Google Scholar] [CrossRef] [PubMed]
  29. Jalil, A.; Zhuo, Z.; Sun, Z.; Wu, F.; Wang, C.; Wu, X. A phosphorene-like InP3 monolayer: Structure, stability, and catalytic properties toward the hydrogen evolution reaction. J. Mater. Chem. A 2020, 8, 1307–1314. [Google Scholar] [CrossRef]
  30. Jing, Y.; Ma, Y.; Li, Y.; Heine, T. GeP3: A Small Indirect Band Gap 2D Crystal with High Carrier Mobility and Strong Interlayer Quantum Confinement. Nano Lett. 2017, 17, 1833–1838. [Google Scholar] [CrossRef]
  31. Tang, P.; Yuan, J.-H.; Song, Y.-Q.; Xu, M.; Xue, K.-H.; Miao, X.-S. BaAs3: A narrow gap 2D semiconductor with vacancy-induced semiconductor–metal transition from first principles. J. Mater. Sci. 2019, 54, 12676–12687. [Google Scholar] [CrossRef]
  32. Xie, M.; Li, Y.; Liu, X.; Yang, J.; Li, H.; Li, X.a. Two-dimensional IV–VA3 monolayers with enhanced charge mobility for high-performance solar cells. Phys. Chem. Chem. Phys. 2022, 24, 20694–20700. [Google Scholar] [CrossRef]
  33. Sun, S.; Meng, F.; Wang, H.; Wang, H.; Ni, Y. Novel two-dimensional semiconductor SnP3: High stability, tunable bandgaps and high carrier mobility explored using first-principles calculations. J. Mater. Chem. A 2018, 6, 11890–11897. [Google Scholar] [CrossRef]
  34. Zhu, Y.-L.; Yuan, J.-H.; Song, Y.-Q.; Xue, K.-H.; Wang, S.; Lian, C.; Li, Z.-N.; Xu, M.; Cheng, X.-M.; Miao, X.-S. Promising photocatalysts with high carrier mobility for water splitting in monolayer Ge2P4S2 and Ge2As4S2. Int. J. Hydrogen Energy 2019, 44, 21536–21545. [Google Scholar] [CrossRef]
  35. Liu, Y.-L.; Jiang, X.-X.; Li, B.; Shi, Y.; Liu, D.-S.; Yang, C.-L. First principles study of photoelectrochemical water splitting in monolayer Sn2S2P4 with high solar-to-hydrogen efficiency. Appl. Phys. Lett. 2021, 119, 143102. [Google Scholar] [CrossRef]
  36. Trung, P.D.; Tong, H.D. First principles study of strain effects on prospective 2D photocatalysts Sn2Se2X4 (X = P, As) with ultra-high charge carrier mobility. Phys. Chem. Chem. Phys. 2024, 26, 4437–4446. [Google Scholar] [CrossRef]
  37. Smidstrup, S.; Stradi, D.; Wellendorff, J.; Khomyakov, P.A.; Vej-Hansen, U.G.; Lee, M.-E.; Ghosh, T.; Jónsson, E.; Jónsson, H.; Stokbro, K. First-principles Green's-function method for surface calculations: A pseudopotential localized basis set approach. Phys. Rev. B 2017, 96, 195309. [Google Scholar] [CrossRef]
  38. Smidstrup, S.; Markussen, T.; Vancraeyveld, P.; Wellendorff, J.; Schneider, J.; Gunst, T.; Verstichel, B.; Stradi, D.; Khomyakov, P.A.; Vej-Hansen, U.G.; et al. QuantumATK: An integrated platform of electronic and atomic-scale modelling tools. J. Phys. Condens. Matter 2020, 32, 015901. [Google Scholar] [CrossRef]
  39. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865–3868. [Google Scholar] [CrossRef]
  40. van Setten, M.J.; Giantomassi, M.; Bousquet, E.; Verstraete, M.J.; Hamann, D.R.; Gonze, X.; Rignanese, G.M. The PseudoDojo: Training and grading a 85 element optimized norm-conserving pseudopotential table. Comput. Phys. Commun. 2018, 226, 39–54. [Google Scholar] [CrossRef]
  41. Heyd, J.; Scuseria, G.E.; Ernzerhof, M. Hybrid functionals based on a screened Coulomb potential. J. Chem. Phys. 2003, 118, 8207–8215. [Google Scholar] [CrossRef]
  42. Polanco, C.A. Nonequilibrium Green’s functions (NEGF) in vibrational energy transport: A topical review. Nanoscale Microscale Thermophys. Eng. 2021, 25, 1–24. [Google Scholar] [CrossRef]
  43. Brandbyge, M.; Mozos, J.-L.; Ordejón, P.; Taylor, J.; Stokbro, K. Density-functional method for nonequilibrium electron transport. Phys. Rev. B 2002, 65, 165401. [Google Scholar] [CrossRef]
  44. Datta, S. Electronic Transport in Mesoscopic Systems; Cambridge University Press: Cambridge, UK, 1997. [Google Scholar]
  45. Henrickson, L.E. Nonequilibrium photocurrent modeling in resonant tunneling photodetectors. J. Appl. Phys. 2002, 91, 6273–6281. [Google Scholar] [CrossRef]
  46. Chen, J.; Hu, Y.; Guo, H. First-principles analysis of photocurrent in graphene PN junctions. Phys. Rev. B 2012, 85, 155441. [Google Scholar] [CrossRef]
  47. Zhang, L.; Gong, K.; Chen, J.; Liu, L.; Zhu, Y.; Xiao, D.; Guo, H. Generation and transport of valley-polarized current in transition-metal dichalcogenides. Phys. Rev. B 2014, 90, 195428. [Google Scholar] [CrossRef]
  48. Haug, H.; Jauho, A.P. Quantum kinetics in transport and optics of semiconductors. In Quantum Kinetics in Transport & Optics of Semiconductors; Springer: Berlin/Heidelberg, Germany, 2008. [Google Scholar]
  49. Palsgaard, M.; Markussen, T.; Gunst, T.; Brandbyge, M.; Stokbro, K. Efficient First-Principles Calculation of Phonon-Assisted Photocurrent in Large-Scale Solar-Cell Devices. Phys. Rev. Appl. 2018, 10, 014026. [Google Scholar] [CrossRef]
  50. Walczak, M.M.; Dryer, D.A.; Jacobson, D.D.; Foss, M.G.; Flynn, N.T. pH Dependent Redox Couple: An Illustration of the Nernst Equation. J. Chem. Educ. 1997, 74, 1195. [Google Scholar] [CrossRef]
  51. Ling, X.; Wang, H.; Huang, S.; Xia, F.; Dresselhaus, M.S. The renaissance of black phosphorus. Proc. Natl. Acad. Sci. USA 2015, 112, 4523–4530. [Google Scholar] [CrossRef]
  52. Ren, H.; Li, Q.-X.; Luo, Y.; Yang, J. Graphene nanoribbon as a negative differential resistance device. Appl. Phys. Lett. 2009, 94, 173110. [Google Scholar] [CrossRef]
  53. Nourbakhsh, Z.; Asgari, R. Charge transport in doped zigzag phosphorene nanoribbons. Phys. Rev. B 2018, 97, 235406. [Google Scholar] [CrossRef]
  54. Nourbakhsh, A.; Zubair, A.; Dresselhaus, M.S.; Palacios, T. Transport Properties of a MoS2/WSe2 Heterojunction Transistor and Its Potential for Application. Nano Lett. 2016, 16, 1359–1366. [Google Scholar] [CrossRef]
  55. Vo, D.D.; Vu, T.V.; Kartamyshev, A.I.; Ho, T.H.; Hieu, N.N. Monolayers Sn2Te2X4 (X = P, As) as promising materials for photocatalytic water splitting and flexible devices: A DFT study. Nanoscale Adv. 2024, 6, 6019–6028. [Google Scholar] [CrossRef] [PubMed]
  56. Trung, P.D.; Tong, H.D. Prediction of quaternary SnGeS2As4 monolayer as a promising photocatalyst for water splitting: A DFT study. J. Phys. D Appl. Phys. 2025, 58, 105307. [Google Scholar] [CrossRef]
  57. Jin, H.; Li, J.; Dai, Y.; Wei, Y. Engineering the electronic and optoelectronic properties of InX (X = S, Se, Te) monolayers via strain. Phys. Chem. Chem. Phys. 2017, 19, 4855–4860. [Google Scholar] [CrossRef] [PubMed]
  58. Li, H.; Jiang, X.; Xu, X.; Xu, G.; Li, D.; Li, C.; Cui, B.; Liu, D.-S. High mobility and enhanced photoelectric performance of two-dimensional ternary compounds NaCuX (X = S, Se, and Te). Phys. Chem. Chem. Phys. 2021, 23, 2475–2482. [Google Scholar] [CrossRef]
  59. Wang, F.; Wang, Z.; Xu, K.; Wang, F.; Wang, Q.; Huang, Y.; Yin, L.; He, J. Tunable GaTe-MoS2 van der Waals p–n Junctions with Novel Optoelectronic Performance. Nano Lett. 2015, 15, 7558–7566. [Google Scholar] [CrossRef]
  60. Pospischil, A.; Furchi, M.M.; Mueller, T. Solar-energy conversion and light emission in an atomic monolayer p–n diode. Nat. Nanotechnol. 2014, 9, 257–261. [Google Scholar] [CrossRef]
  61. Zhao, P.; Yang, H.; Li, J.; Jin, H.; Wei, W.; Yu, L.; Huang, B.; Dai, Y. Design of new photovoltaic systems based on two-dimensional group-IV monochalcogenides for high performance solar cells. J. Mater. Chem. A 2017, 5, 24145–24152. [Google Scholar] [CrossRef]
  62. Wang, Q.; Li, J.; Liang, Y.; Nie, Y.; Wang, B. KAgSe: A New Two-Dimensional Efficient Photovoltaic Material with Layer-Independent Behaviors. ACS Appl. Mater. Interfaces 2018, 10, 41670–41677. [Google Scholar] [CrossRef]
Figure 1. (a) Top and (b) side views of the crystal structure of the Sn2Se2P4 monolayer. (c) In-plane electron localization function map of the Sn2Se2P4 monolayer. Band structure and partial density of the state of the Sn2Se2P4 monolayer calculated at the (d) PBE and (e) HSE06 level of theory. The inset in (d) shows the Brillouin zone.
Figure 1. (a) Top and (b) side views of the crystal structure of the Sn2Se2P4 monolayer. (c) In-plane electron localization function map of the Sn2Se2P4 monolayer. Band structure and partial density of the state of the Sn2Se2P4 monolayer calculated at the (d) PBE and (e) HSE06 level of theory. The inset in (d) shows the Brillouin zone.
Nanomaterials 15 00679 g001
Figure 2. (a) Schematic illustration of the Sn2Se2P4 monolayer under uniaxial strain. The directions of tensile and compressive strains are indicated by the arrow orientations. Calculated (b,c) band structures, (d,e) band edges with band gaps, and the (f) effective mass of the Sn2Se2P4 monolayer under uniaxial strain ranging from −8% to 8% along the a-axis and b-axis, respectively.
Figure 2. (a) Schematic illustration of the Sn2Se2P4 monolayer under uniaxial strain. The directions of tensile and compressive strains are indicated by the arrow orientations. Calculated (b,c) band structures, (d,e) band edges with band gaps, and the (f) effective mass of the Sn2Se2P4 monolayer under uniaxial strain ranging from −8% to 8% along the a-axis and b-axis, respectively.
Nanomaterials 15 00679 g002
Figure 3. Schematic diagrams of the two-probe system for computing the transport properties of the Sn2Se2P4 monolayer along the (a) a-direction and (b) b-direction. Channel length was set to 30 Å.
Figure 3. Schematic diagrams of the two-probe system for computing the transport properties of the Sn2Se2P4 monolayer along the (a) a-direction and (b) b-direction. Channel length was set to 30 Å.
Nanomaterials 15 00679 g003
Figure 4. (a) Computed IV characteristic curves of the unstrained Sn2Se2P4 monolayer. (b) To better observe the current changes, the logarithm scale was used to plot the IV characteristic curve. Computed projected local density of states (PLDOS) of the unstrained Sn2Se2P4 monolayer at 0.4 V along (c) the a-axis and (d) the b-axis.
Figure 4. (a) Computed IV characteristic curves of the unstrained Sn2Se2P4 monolayer. (b) To better observe the current changes, the logarithm scale was used to plot the IV characteristic curve. Computed projected local density of states (PLDOS) of the unstrained Sn2Se2P4 monolayer at 0.4 V along (c) the a-axis and (d) the b-axis.
Nanomaterials 15 00679 g004
Figure 5. (a) Calculated IV characteristic curves of the Sn2Se2P4 monolayer at uniaxial strains ranging from 0% to +8% in the a-axis. (b) Comparative results of IV characteristics for a-axis strains (0%, +6%, +8%) calculated along the a-axis versus those obtained along the b-axis. The PLDOS of the Sn2Se2P4 monolayer at 0.6 V under (c) 0%, (d) +6%, and (e) +8% a-axis uniaxial strain computed along the a-axis; and (f) the +8% a-axis uniaxial strain calculated along the b-axis.
Figure 5. (a) Calculated IV characteristic curves of the Sn2Se2P4 monolayer at uniaxial strains ranging from 0% to +8% in the a-axis. (b) Comparative results of IV characteristics for a-axis strains (0%, +6%, +8%) calculated along the a-axis versus those obtained along the b-axis. The PLDOS of the Sn2Se2P4 monolayer at 0.6 V under (c) 0%, (d) +6%, and (e) +8% a-axis uniaxial strain computed along the a-axis; and (f) the +8% a-axis uniaxial strain calculated along the b-axis.
Nanomaterials 15 00679 g005
Figure 6. (a) Computed IV characteristic curves of the Sn2Se2P4 monolayer under uniaxial strains ranging from −8% to 0% in the a-axis. The calculated PLDOS at 0.6 V under (b) −6% and (c) −8% a-axis uniaxial strain.
Figure 6. (a) Computed IV characteristic curves of the Sn2Se2P4 monolayer under uniaxial strains ranging from −8% to 0% in the a-axis. The calculated PLDOS at 0.6 V under (b) −6% and (c) −8% a-axis uniaxial strain.
Nanomaterials 15 00679 g006
Figure 7. (a) b-axis 0–+8% tensile strain and (b) −8–0% compressive strain IV characteristics of the Sn2Se2P4 monolayer, with logarithmic-scale IV plots under (c) tensile and (d) compressive strain along the b-axis.
Figure 7. (a) b-axis 0–+8% tensile strain and (b) −8–0% compressive strain IV characteristics of the Sn2Se2P4 monolayer, with logarithmic-scale IV plots under (c) tensile and (d) compressive strain along the b-axis.
Nanomaterials 15 00679 g007
Figure 8. Computed (a) IV characteristics and PLDOS at (b) 0.1 V and (c) 0.4 V in the b-axis of the unstrained Sn2Se2P4 monolayer along with the calculated (d) IV characteristics and PLDOS at (e) 0.1 V and (f) 0.4 V for the −2% b-axis compressive strain counterpart.
Figure 8. Computed (a) IV characteristics and PLDOS at (b) 0.1 V and (c) 0.4 V in the b-axis of the unstrained Sn2Se2P4 monolayer along with the calculated (d) IV characteristics and PLDOS at (e) 0.1 V and (f) 0.4 V for the −2% b-axis compressive strain counterpart.
Nanomaterials 15 00679 g008
Figure 9. Schematic diagrams of the Sn2Se2P4 monolayer-based optical nanodevices along the (a) a-axis and (b) b-axis, respectively.
Figure 9. Schematic diagrams of the Sn2Se2P4 monolayer-based optical nanodevices along the (a) a-axis and (b) b-axis, respectively.
Nanomaterials 15 00679 g009
Figure 10. (a) Calculated absorption coefficients of Sn2Se2P4 monolayer under a-axis compressive strain (−8% ≤ εa ≤ −2%) along a-axis (blue curves) and b-axis (red curves), respectively. Computed Jph under a-axis compressive strain (−8% ≤ εa ≤ −2%) along (b) a-axis and (c) b-axis, respectively. Comparative results of (d) Jph, (e) Rph, and (f) EQE under 0%, −2%, and −4% a-axis compressive strain computed along both axes.
Figure 10. (a) Calculated absorption coefficients of Sn2Se2P4 monolayer under a-axis compressive strain (−8% ≤ εa ≤ −2%) along a-axis (blue curves) and b-axis (red curves), respectively. Computed Jph under a-axis compressive strain (−8% ≤ εa ≤ −2%) along (b) a-axis and (c) b-axis, respectively. Comparative results of (d) Jph, (e) Rph, and (f) EQE under 0%, −2%, and −4% a-axis compressive strain computed along both axes.
Nanomaterials 15 00679 g010
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

Xu, H.; Xu, Y. Strain Engineering of Anisotropic Electronic, Transport, and Photoelectric Properties in Monolayer Sn2Se2P4. Nanomaterials 2025, 15, 679. https://doi.org/10.3390/nano15090679

AMA Style

Xu H, Xu Y. Strain Engineering of Anisotropic Electronic, Transport, and Photoelectric Properties in Monolayer Sn2Se2P4. Nanomaterials. 2025; 15(9):679. https://doi.org/10.3390/nano15090679

Chicago/Turabian Style

Xu, Haowen, and Yuehua Xu. 2025. "Strain Engineering of Anisotropic Electronic, Transport, and Photoelectric Properties in Monolayer Sn2Se2P4" Nanomaterials 15, no. 9: 679. https://doi.org/10.3390/nano15090679

APA Style

Xu, H., & Xu, Y. (2025). Strain Engineering of Anisotropic Electronic, Transport, and Photoelectric Properties in Monolayer Sn2Se2P4. Nanomaterials, 15(9), 679. https://doi.org/10.3390/nano15090679

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