Abstract
The detection of biomarkers in exhaled breath offers an efficient approach for the early-stage identification of liver disease. In this work, first-principles calculations were employed to investigate the adsorption and sensing properties of Rh- and Ir-decorated PtS2 monolayers toward four liver disease biomarkers (LDBs: C2H6O, C3H8O, C3H6O, and C5H8). The results reveal that pristine PtS2 exhibits a low affinity for these biomarkers, whereas single-atom decoration with Rh or Ir significantly enhances both adsorption energy and charge transfer. These interactions were further elucidated through analyses of projected density of states, total electron density, charge density difference, and charge transfer. Furthermore, the adsorption of all four LDBs results in |ΔE g| exceeding 16.57%, except for the C3H6O/Rh@PtS2 system. Notably, the adsorption of C5H8 and C3H8O induces a pronounced semiconductor-to-metal transition in Rh@PtS2 and Ir@PtS2 systems, respectively. Crucially, both Rh@PtS2 and Ir@PtS2 show excellent selectivity, exhibiting significantly higher adsorption strengths for the target LDBs compared to common interfering molecules present in exhaled breath (H2O, N2, CO2, and CH4). Additionally, Rh@PtS2 exhibits a suitable recovery time (τ) of 22.7 s at 298 K for C2H6O, along with moderate τ values of 0.27 s (C2H6O) and 3.94 s (C3H6O) at 348 K. Consequently, Rh@PtS2 emerges as a promising reversible sensor material for the detection of C2H6O and C3H6O. This study provides a strategic blueprint for developing PtS2-based gas sensor applications for the medical field.


1. Introduction
Chronic liver disease constitutes a significant and growing global health concern, characterized by a silent progression that frequently leads to severe and irreversible conditions such as cirrhosis and hepatocellular carcinoma. , The late-stage diagnosis of these diseases severely limits therapeutic options and contributes to high mortality rates. Traditional diagnostic methods, including liver biopsies and serological tests, are often associated with high costs, lengthy procedures, and operational risks. In recent years, the noninvasive and real-time detection of volatile organic compounds (VOCs) in the exhaled breath of individuals with liver disease has emerged as a promising approach for effective early-stage screening. Previous studies have identified specific VOCs, including ethanol (C2H6O), isopropanol (C3H8O), acetone (C3H6O), and isoprene (C5H8), as key biomarkers whose concentrations are significantly altered during liver dysfunction. − Consequently, researchers worldwide are actively pursuing the development of gas sensors to monitor these VOCs. ,, However, current gas-sensitive materials still face challenges, including weak adsorption strength, poor sensitivity, and slow response times. Therefore, the exploration of novel, rapid, and highly sensitive gas-sensing materials is essential for the early diagnosis of liver disease.
Among various sensing materials, two-dimensional (2D) transition metal dichalcogenides (TMDs) have attracted significant interest due to their large specific surface area and unique electronic properties. , In particular, recent studies have highlighted the substantial potential of the PtS2 monolayer, a semiconducting TMD characterized by inherent stability and high electron mobility, for gas-sensing applications. , However, the relatively inert surface of pristine PtS2 results in weak physisorption interactions with the VOCs. For instance, first-principles calculations by Kanungo et al. demonstrated that the pristine PtS2 monolayer exhibits poor affinity for ethanol, acetone, and isoprene. The introduction of sulfur vacancies, however, enhances the chemical activity of PtS2 monolayers, leading to increased adsorption energies and charge transfer for these VOCs. Additionally, single-atom metal decoration has emerged as another highly effective strategy to further improve sensing performance. , This approach introduces catalytically active sites that modulate local charge distribution and facilitate stronger, more specific analyte binding. For instance, functionalization with Nb or Ni atoms markedly improves the adsorption capacity and sensing properties of PtS2 monolayers toward dissolved gases in transformer oil, such as CO, C2H2, and C2H4. ,
In addition to non-noble metals, noble metals like Rh and Ir are also highly attractive candidates due to their exceptional catalytic activity and distinctive electronic structures. , Their partially filled d-orbitals facilitate strong orbital hybridization and efficient charge transfer with the functional groups of the liver disease biomarkers. Thus, decorating PtS2 with Rh or Ir atoms is expected to create high-affinity binding sites specifically tailored for certain VOCs, transforming the inert substrate into a highly responsive sensing material. To confirm this hypothesis, density functional theory (DFT) calculations were conducted to systematically investigate the adsorption and sensing mechanisms of Rh- and Ir-decorated PtS2 monolayers (Rh@PtS2 and Ir@PtS2) toward four liver disease biomarkers (LDBs: C2H6O, C3H8O, C3H6O, and C5H8). The adsorption strength and corresponding electronic properties for each system, including the density of states, charge density difference, and band structure, were thoroughly analyzed. Subsequently, the sensing characteristics were evaluated by analyzing the changes in conductivity and work function of the Rh@PtS2 and Ir@PtS2 systems upon gas adsorption, as well as theoretical recovery time. This work aims to assess the potential of these materials for high-performance sensing and to provide a theoretical foundation for developing novel PtS2-based chemiresistive sensors for the rapid, real-time, and noninvasive diagnosis of liver diseases.
2. Computational Details
Structural optimizations and electronic property calculations for all adsorption systems were conducted within the DMol3 module of Materials Studio (MS) software. The generalized gradient approximation with the Perdew–Burke–Ernzerhof functional (GGA-PBE) , was adopted to describe electron exchange-correlation effects. The integration of the DFT semicore pseudopotential (DSPP) approach with the double numerical plus polarization (DNP) orbital basis was utilized to significantly accelerate computational processes. Long-range interactions between VOCs and the PtS2 monolayer were rigorously evaluated using Grimme’s dispersion correction (DFT-D2). Geometric optimizations were converged until the energy, maximum force, and maximum displacement were less than 10–5 Ha, 2 × 10–3 Ha/Å, and 5 × 10–3 Å, respectively. The self-consistent field (SCF) energy convergence threshold was set to 10– 6 Ha, using the DIIS algorithm for acceleration. The Brillouin zone was sampled with a 4 × 4 × 1 k-point mesh for structural relaxations and an 8 × 8 × 1 mesh for electronic property calculations. A 20 Å vacuum layer was applied to eliminate periodic interlayer interactions with an orbital cutoff radius of 5.0 Å and a thermal smearing of 0.005 Ha.
Binding energy (E bin) is a fundamental metric for assessing the stability of TM@PtS2 systems and is defined as follows:
| 1 |
where E TM@PtS2 , E PtS2 , and E TM denote the total energies of the TM@PtS2 complex, pristine PtS2, and a single TM atom, respectively. The interaction strength between VOCs and the TM@PtS2 monolayer is evaluated by calculating the adsorption energy (E ads) and charge transfer (Q T):
| 2 |
| 3 |
Here, E VOC/TM@PtS2 and E VOC represent the energies of the VOC-adsorbed TM@PtS2 system and the isolated VOC molecule, respectively, while the Q adsorbed‑VOC and Q free‑VOC correspond to the charge of the adsorbed VOC and its free-state counterpart. A more negative Q T value indicates stronger interactions between the VOCs and the TM@PtS2 systems.
3. Results and Discussion
3.1. Stability Analysis of TM@PtS2 (TM = Rh, Ir) and Adsorption of LDBs on Pristine PtS2
A single-layer PtS2 consists of a single plane of Pt atoms sandwiched between two planes of S atoms in an S–Pt–S trilayer arrangement, as shown in Figure a. Herein, a 4 × 4 × 1 supercell of PtS2, containing 32 S atoms and 16 Pt atoms, was employed for all calculations. Three potential adsorption sites, including Pttop, Stop, and hollow sites, were considered for the modification of Rh and Ir atoms. The calculated binding energies of Rh and Ir atoms at these sites are listed in Figure a. It is evident that both dopants preferentially adsorb on the Pttop site, followed by the Hollow and Stop sites. The large negative binding energies (−3.66 eV for Rh and −3.72 eV for Ir) indicate that Rh and Ir atoms can be stably embedded in the PtS2 monolayer, as shown in Figure b,c. As depicted in Figure e,f, abundant electron depletion regions are observed around the Rh and Ir atoms, while significant electron accumulation occurs on the PtS2 surface. Consequently, the Rh and Ir atoms donate approximately 0.104 and 0.102 e, respectively, to the substrate. This substantial charge transfer further demonstrates the high stability of the Rh@PtS2 and Ir@PtS2 systems.
1.
Top and side views of the optimized structures for (a) pristine PtS2, (b) Ir@PtS2, (c) Rh@PtS2, and (d) four LDBs molecules, along with the charge density difference (CDD) of (e) Ir@PtS2, and (f) Rh@PtS2. The green (red) regions represent the charge accumulation (depletion).
2.
(a) Comparison of binding energy for the TM@ PtS2, (b) projected density of states (PDOSs) of (b) Rh@PtS2 and (c) Ir@PtS2, along with the band structures of (d) pristine PtS2, (e) Ir@PtS2, and (f) Rh@PtS2.
Figure b,c presents the projected densities of states (PDOSs) for the Rh@PtS2 and Ir@PtS2 monolayers. For the Rh@PtS2, strong hybridization occurs between the S-2p orbital and the Rh-4d orbital within the energy range of −7.60–1.58 eV, with distinct hybridization peaks appearing at −6.05, −4.96, −3.86, −1.30, and −0.28 eV. This hybridization facilitates the formation of Rh–S bonds. In contrast, the interactions between the S-2s orbital and the Rh-5s, Rh-4p orbitals are significantly weaker. For Ir@PtS2 (Figure c), the orbital hybridization between Ir and S atoms exhibits a pattern similar to that of the Rh–S system, confirming the formation of Ir–S covalent bonds between the Ir atom and its three neighboring S atoms. These findings account for the remarkable stability of the TM@PtS2 structures. Additionally, the calculated bandgap of the pristine PtS2 monolayer is 1.794 eV, which aligns closely with the reported value of 1.738 eV. However, doping with Rh and Ir atoms substantially reduces the bandgap to 0.623 and 0.356 eV, respectively. This suggests that Rh and Ir doping can greatly enhance the electrical conductivity of the PtS2 monolayer, thereby improving its potential for sensing applications.
The adsorption behaviors of four LDBs on the pristine PtS2 surface were systematically investigated. To identify the most energetically favorable structures, various adsorption sites and molecular orientations were considered, with the lowest-energy configurations shown in Figure a1–d1. The calculated adsorption strengths of these LDBs follow the order: C5H8 > C3H6O > C3H8O > C2H6O, with adsorption energies ranging from −0.27 to −0.57 eV. The corresponding equilibrium adsorption distances (d) fall within the range of 4.235 to 4.743 Å. These results indicate that the LDBs exhibit physisorption characteristics, governed primarily by van der Waals forces. This conclusion is further supported by the total electron density (TED) plots in Figure a2–d2, which reveal minimal electron sharing at the adsorption interfaces. Additionally, as shown in Figure a3–d3, the C2H6O, C3H6O, C3H8O, and C5H8 all function as electron acceptors, gaining approximately 0.058, 0.069, 0.057, and 0.060 e, respectively, from the PtS2 substrate. The small charge transfer further confirms the weak affinity between PtS2 and these LDBs. As shown in Figure , adsorption of C2H6O and C3H8O has a negligible effect on the PtS2 bandgap, whereas the adsorption of C3H6O and C5H8 reduces the bandgap to 1.310 and 1.333 eV, respectively. Nevertheless, the inherently weak adsorption capacity of PtS2 for LDBs limits its direct application as a gas sensor.
3.
Optimized adsorption structures, total electron density (TED), and CDD of (a1–a3) C2H6O, (b1–b3) C3H6O, (c1–c3) C3H8O, and (d1–d3) C5H8 adsorbed on the pristine PtS2 monolayer. The green (red) regions represent the charge accumulation (depletion).
4.
Band structures of the adsorbed PtS2 monolayer with (a) C2H6O, (b) C3H6O, (c) C3H8O, and (d) C5H8.
3.2. Adsorption Analysis of LDBs on the Rh@PtS2 Monolayer
The adsorption behaviors of the four LDBs (C2H6O, C3H8O, C3H6O, and C5H8) on the Rh@PtS2 surface were investigated, with the lowest-energy configurations depicted in Figure . The calculated adsorption energies for C2H6O and C3H8O are −0.79 and −0.97 eV, respectively, with corresponding adsorption distances of 2.282 and 2.263 Å, as shown in Figure a,c. Because these distances exceed the sum of the covalent radii of Rh and O atoms, these interactions are characteristic of physisorption. By contrast, the Rh atom forms direct chemical bonds with both C3H6O and C5H8, resulting in substantially higher adsorption energies of −0.87 and −1.72 eV, respectively. These findings demonstrate that C3H6O and C5H8 undergo chemisorption on the Rh@PtS2 surface. To further analyze these interactions, the TED and CDD were calculated for each adsorption configuration (Figure ). As shown in Figure a, c, e, and g, all four LDBs exhibit substantial electron sharing with the PtS2 substrate, as indicated by the blue regions at the adsorption interfaces. Consequently, the Rh dopant atom substantially enhances the adsorption capacity of the PtS2 monolayer.
5.
Top and side views of the optimized structures for (a) C2H6O, (b) C3H6O, (c) C3H8O, and C5H8 adsorbed on the Rh@PtS2 monolayer.
6.

Comparisons of TED and CDD plots for different adsorption systems: (a, b) C2H6O/Rh@PtS2, (c, d) C3H6O/Rh@PtS2, (e, f) C3H8O/Rh@PtS2, and (g, h) C5H8/Rh@PtS2.
As shown in Figure b, f, h, significant charge accumulation is observed around the Rh atoms, while minimal charge is distributed around the C2H6O, C3H8O, and C5H8 molecules. Consequently, these molecules donate approximately 0.163, 0.119, and 0.140 e, respectively, to the Rh@PtS2 monolayer. In contrast, the C3H6O molecule acts as an electron acceptor, gaining approximately 0.091 e from the Rh@PtS2 substrate. The large charge transfer between these LDBs and Rh@PtS2 further indicates their strong interactions. To further analyze the bonding mechanism, the PDOS for each adsorption system was calculated and is presented in Figure a–d. For the C2H6O and C3H8O adsorption systems (Figure a,c), the adsorption is primarily attributed to the hybridization of O-2p and Rh-4d orbitals within the energy range of −8.00 to 1.50 eV. Furthermore, three resonance peaks are observed at about −6.62, −4.94, and −3.33 eV for the C2H6O/Rh@PtS2 system, whereas four peaks appear at −6.80, −5.20, −4.64, and −3.28 eV for the C3H8O/Rh@PtS2 system. Therefore, the adsorption strength of C3H8O is greater than that of C2H6O, which matches well with the previously calculated adsorption energies.
7.
Comparisons of PDOSs and band structures for different adsorption systems: (a, e) C2H6O/Rh@PtS2, (b, f) C3H6O/Rh@PtS2, (c, g) C3H8O/Rh@PtS2, and (d, h) C5H8/Rh@PtS2.
In contrast, the orbital hybridization between O-2p and Rh-4d in the C3H6O adsorption system is significantly stronger than that observed in the C2H6O and C3H8O systems (Figure b). Furthermore, the C-2p orbital of C3H6O exhibits pronounced overlap with the Rh-4d orbital, which underpins the chemical adsorption mechanism of C3H6O on the Rh@PtS2 surface. Similarly, the chemisorption of C5H8 involves the bonding of two carbon atoms to the active Rh atom, facilitated by a strong interaction between their C-2p orbitals and the Rh-4d orbital. Additionally, the adsorption of these molecules induces notable changes in the electronic properties of the Rh@PtS2 substrate. Specifically, C2H6O adsorption increases the bandgap from 0.623 to 0.727 eV, while C3H6O and C3H8O adsorption decrease it to 0.586 and 0.409 eV, respectively, as shown in Figure e–g. Most notably, the adsorption of C5H8 induces a semiconductor-to-metal transition in Rh@PtS2 (Figure h). These pronounced changes in the bandgap clearly demonstrate the high sensitivity of Rh@PtS2 toward these specific gases.
3.3. Adsorption Analysis of LDBs on the Ir@PtS2 Monolayer
This section explores the adsorption characteristics of four LDBs on the Ir@PtS2 surface. The most stable adsorption configurations, along with their TED and CDD plots, are illustrated in Figures and . As shown in Figure a, c, both C2H6O and C3H8O adsorb on the Ir@PtS2 via their O atoms approaching the active Ir sites, exhibiting the E ads values of −1.19 and −1.35 eV, respectively, and adsorption distances of 2.252 and 2.263 Å. Obviously, the two gases do not form chemical bonds with Ir@PtS2 and, therefore, are classified as physisorption. In contrast, one C and one O atom from C3H6O, as well as two C atoms from C5H8, form chemical bonds with the Ir atom. This results in significantly more negative E ads values (−1.19 eV ∼ −2.24 eV) and shorter adsorption distances (2.011 Å ∼ 2.251 Å), indicating that C3H6O and C5H8 are chemisorbed on the surface. Additionally, the substantial electron sharing between these LDBs and Ir@PtS2, as evidenced by the TED and CDD analyses, suggests their strong interactions, which is further supported by the significant charge transfer illustrated in Figure . Specifically, C2H6O and C3H8O act as electron donors, whereas C3H6O and C5H8 function as electron acceptors.
8.
Top and side views of the optimized structures for (a) C2H6O, (b) C3H6O, (c) C3H8O, and C5H8 adsorbed on the Ir@PtS2 monolayer.
9.
Comparisons of TED and CDD plots for different adsorption systems: (a, b) C2H6O/Ir@PtS2, (c, d) C3H6O/Ir@PtS2, (e, f) C3H8O/Ir@PtS2, and (g, h) C5H8/Ir@PtS2.
Figure presents the PDOSs and band structures for various Ir@PtS2 adsorption systems. As depicted in Figure a–d, the affinity of Ir@PtS2 for these gases primarily stems from orbital hybridization between Ir-5d and O-2p (or C-2p) orbitals across the entire energy range. Specifically, for C2H6O and C3H8O, weak hybridization near the Fermi level and strong hybridization within the range of −7.40 to −2.20 eV are observed between Ir-5d and O-2p orbitals. In contrast, C3H6O and C5H8 exhibit significant orbital overlap at the Fermi level and pronounced hybridization spanning −8.13 to 1.50 eV, involving the Ir-5d and either the O-2p orbital of C3H6O or the C-2p orbital of C5H8, as illustrated in Figure b,d. Thus, C3H6O and C5H8 exhibit obviously stronger chemisorption on Ir@PtS2 compared to C2H6O and C3H8O. Additionally, Figure e–h reveals that the adsorption of C2H6O, C3H6O, and C5H8 increases the bandgap of Ir@PtS2 from its initial value of 0.356 eV to 0.415, 0.457, and 0.525 eV, respectively. In contrast, the adsorption of C3H8O induces a semiconductor-to-metal transition in the Ir@PtS2 monolayer.
10.
Comparisons of PDOSs and band structures for different adsorption systems: (a, e) C2H6O/Ir@PtS2, (b, f) C3H6O/Ir@PtS2, (c, g) C3H8O/Ir@PtS2, and (d, h) C5H8/Ir@PtS2.
3.4. Effect of Humidity and Interfering Gases
Exhaled breath is a complex matrix, typically comprising water vapor, nitrogen, carbon dioxide, and trace amounts of other species like CH4. A highly effective sensor must exhibit preferential affinity for the target LDBs over these potential interferents. Consequently, this section systematically investigates the adsorption behavior of H2O, N2, CO2, and CH4 on both Rh@PtS2 and Ir@PtS2 surfaces, with the main results summarized in Table .
1. Adsorption Energy (E ads), Charge Transfer (Q T), and Adsorption Distance (d) of a Single H2O Molecule and Various Interfering Gases Adsorbed on the Rh@PtS2 and Ir@PtS2 Monolayers, Where the R sum Represents the Sum of the Theoretical Radii of Two Atoms.
| materials | gases | E ads, eV | Q T, e | d, Å | R sum, Å |
|---|---|---|---|---|---|
| Rh-PtS2 | H2O | –0.53 | 0.179 | 2.343 | 2.21 (Rh–O) |
| N2 | –0.69 | –0.064 | 1.985 | 2.29 (Rh–N) | |
| CH4 | –0.23 | 0.048 | 2.659 | 2.40 (Rh–C) | |
| CO2 | –0.15 | –0.009 | 2.437 | 2.40 (Rh–C) | |
| Ir-PtS2 | H2O | –0.93 | 0.218 | 2.330 | 2.28 (Ir–O) |
| N2 | –1.06 | –0.066 | 1.914 | 2.36 (Ir–N) | |
| CH4 | –0.26 | 0.094 | 2.527 | 2.47 (Ir–C) | |
| CO2 | –0.49 | –0.011 | 2.476 | 2.47 (Ir–C) |
For the Rh-PtS2 system, H2O, CH4, and CO2 exhibit weak physisorption, with adsorption energies of −0.53, −0.23, and −0.15 eV, respectively. These values are substantially lower than the E ads values observed for all LDBs, indicating that the target analytes can competitively adsorb onto the Rh-PtS2 surface, even in the presence of these interfering species. Although N2 shows a slightly stronger interaction, with a relatively large E ads of −0.69 eV and a short adsorption distance of 1.985 Å, its adsorption remains weaker than that of all LDBs. Therefore, Rh-PtS2 shows good selectivity toward the target LDBs. For the Ir-PtS2 system, the adsorption energies of H2O, CH4, and CO2 are −0.93, −0.26, and −0.49 eV, with corresponding adsorption distances of 2.330 2.527, and 2.476 Å, respectively. All these adsorption distances exceed the sum of the theoretical radii of the two interacting atoms, confirming that these gases are physisorbed. In contrast, N2 exhibits chemisorption on Ir-PtS2, as indicated by its large E ads of −1.06 eV and a short adsorption distance of 1.914 Å. Nevertheless, the significantly stronger binding of all four LDBs compared to H2O, N2, CO2, and CH4 underscores the high selectivity of the Ir@PtS2 monolayer.
Figure gives the most stable structures of Rh@PtS2 and Ir@PtS2 adsorbed with different numbers of H2O molecules to simulate the effect of high humidity on their selectivity. For the Rh@PtS2 system, the average adsorption energies for two and three H2O molecules are −0.63 and −0.51 eV, respectively. A similar trend is observed for the Ir@PtS2 system, where the average adsorption energies are −0.92 eV for two H2O molecules and −0.87 eV for three H2O molecules. This trend indicates that as surface coverage increases, the binding affinity for additional water molecules weakens. This is likely because the first H2O molecule strongly interacts with the active metal dopant, while subsequent molecules primarily form weaker hydrogen bonds with already adsorbed water. It is noteworthy that in both cases, the average adsorption energy is weaker than the adsorption energies calculated for all four target LDBs. This suggests that the LDBs can competitively displace adsorbed water molecules or preferentially bind to the active sites, even in a high-humidity environment. Therefore, both Rh@PtS2 and Ir@PtS2 are predicted to maintain high selectivity against water vapor, which is crucial for their practical application in breath analysis.
11.
Lowest-energy configurations of different numbers of H2O molecules adsorbed on (a,b) Rh@PtS2 and (c,d) Ir@PtS2 monolayers, along with their corresponding average adsorption energies.
To further elucidate the microscopic interactions, H2O and N2 adsorption systemsboth exhibiting relatively high adsorption strengthsare selected for detailed electronic structure analysis, as shown in Figures and . From Figure a1–a4, H2O is obliquely adsorbed on both Rh@PtS2 and Ir@PtS2, with adsorption distances of 2.343 and 2.330 Å, respectively. These distances exceed the sum of the theoretical radii of Rh–O or Ir–O, indicating that these adsorptions are clarified as physisorption. In contrast, N2 is vertically adsorbed on both Rh@PtS2 and Ir@PtS2, with adsorption distances of 1.985 and 1.914 Å, which are smaller than the sum of the theoretical radii of Rh–N or Ir–N, suggesting the chemisorption feature for the N2. Therefore, both materials exhibit physisorption toward H2O and chemisorption toward N2. Furthermore, as depicted in Figure b1–b4, the H2O and N2 molecules share some charges with Rh@PtS2 and Ir@PtS2, indicating their strong interactions. The CDD plots reveal significant charge depletion around H2O and charge accumulation around N2. Consequently, H2O acts as an electron donor, while N2 serves as an electron acceptor in these adsorption systems.
12.
(a1–a4) Top and side views of atomic structures, (b1–b4) TED, and (c1–c4) CDD plots of the (a1–c1) H2O/Rh@PtS2, (a2–c2) H2O/Ir@PtS2, (a3–c3) N2/Rh@PtS2 and (a4–c4) N2/Ir@PtS2 systems.
13.
PDOS plots of different adsorption systems: (a) H2O/Rh@PtS2, (b) N2/Rh@PtS2, (c) H2O/Ir@PtS2, and (d) N2/Ir@PtS2.
Figure presents the PDOS values for different adsorption systems. In the H2O/Rh@PtS2 system (Figure a), weak orbital hybridization is observed between the O-2p and Rh-4d states within the energy range of −7.50 to −3.35 eV, resulting in a weak affinity of Rh@PtS2 for H2O. In contrast, the orbital overlap in the H2O/Ir@PtS2 system is significantly stronger than that in H2O/Rh@PtS2, indicating a greater adsorption strength of Ir@PtS2 for H2O. For the N2/Rh@PtS2 system (Figure c), broad orbital hybridization between N-2p and Rh-4d states is observed within the energy range of −10.94 to −5.00 eV, with two resonance peaks at approximately −10.26 and −6.65 eV, indicating the formation of a Rh–N covalent bond. An even broader orbital overlap is found in the N2/Ir@PtS2 system, where hybridization occurs between the 2s and 2p orbitals of N and the 6s, 5p, and 5d orbitals of Ir around −11.20 eV, along with strong hybridization between N-2p and Ir-5d states in the range of −7.60 to −5.00 eV. This suggests a strong affinity of Ir@PtS2 for N2. Nevertheless, the interaction strengths with H2O and N2 remain lower than those observed for all four target biomarkers. Therefore, this electronic-level analysis confirms the high selectivity of the Rh@PtS2 and Ir@PtS2 monolayers, reinforcing their potential for reliable biomarker detection in complex, humid environments.
3.5. Evaluation of Sensing Behavior
Sensitivity is a critical parameter in evaluating the gas sensing performance of a material, as it directly determines its practical applicability. Gas adsorption induces changes in surface charge distribution, leading to dynamic shifts in the bandgap, which in turn affect carrier concentration and electrical conductivity. Therefore, the change in the band gap (ΔE g) of a material upon gas adsorption can be used as an indicator of its sensitivity to a specific gas. Alternatively, the change in the work function (ΔΦ) resulting from gas adsorption can also serve as a metric for assessing the intrinsic sensitivity of a material. The conductivity (σ) and work function (Φ) can be calculated as follows , :
| 4 |
| 5 |
Here, λ, k B, and T represent a material-specific constant, the Boltzmann constant, and the absolute temperature, respectively, while Φvacuum and ΦFermi represent the vacuum energy level and Fermi level of the material, respectively.
Figure illustrates the ΔE g and ΔΦ of pristine PtS2 and TM@PtS2 monolayers before and after LDB adsorption. As shown in Figure a, the PtS2 monolayer exhibits high sensitivity to C3H6O and C5H8 with ΔE g values of −25.70 and −27.00%. However, its poor adsorption capacity restricts its real-world applicability. In contrast, the adsorption of C2H6O, C3H8O, and C5H8 on the Rh@PtS2 monolayer induces ΔE g values of 16.69, −34.35, and −100.00%, respectively, demonstrating its excellent sensitivity to these three LDBs. More remarkably, the Ir@PtS2 monolayer displays high responsiveness to all four LDBs, with |ΔE g| ranging from 16.57 to 100%, indicating its potential as a resistive gas sensor for these analytes. Regarding work function variation (Figure b), the adsorption of C2H6O, C3H8O, and C5H8 yields ΔΦ values of −16.92, −18.45, and −14.36% for Rh@PtS2, and 9.24, −19.49, and −18.45% for Ir@PtS2, respectively. These results indicate that both Rh@PtS2 and Ir@PtS2 monolayers are promising candidates for work function-based sensors for the detection of C2H6O, C3H8O, and C5H8 molecules.
14.
Changes in the (a) bandgap and (b) work functions for various materials induced by the adsorption of the LDBs.
In addition to its high sensitivity, the practical utility of a gas sensor is fundamentally determined by its reusability. An appropriate recovery time is essential for real-time continuous monitoring. According to transition state theory, the recovery time (τ) can be calculated using the van’t Hoff-Arrhenius equation , :
| 6 |
where ν0 represents the attempt frequency (1012 s– 1), E ads is the adsorption energy, k B is the Boltzmann constant, and T is the absolute temperature. The calculated recovery times for LDBs desorption from Rh@PtS2 and Ir@PtS2 surfaces at various temperatures are presented in Figure .
15.
Recovery times for the desorption of various LDBs from (a) Rh@PtS2 and (b) Ir@PtS2 surfaces.
As illustrated in Figure a, C2H6O exhibits a reasonable recovery time of 22.7 s at 298 K. Furthermore, both C2H6O and C3H6O molecules can rapidly desorb from the Rh@PtS2 surface at 348 K, with τ values of 0.27 and 3.94 s, respectively. These results confirm the suitability of Rh@PtS2 for the reversible detection of these two biomarkers. In contrast, C3H8O and C5H8 molecules are bound so strongly that their theoretical recovery times extend to 2.5 × 104 s and 1.2 × 1017 s at 298 K, respectively, rendering desorption practically impossible at room temperature. When the temperature is increased to 398 K, the recovery time for C3H8O desorption from the Rh@PtS2 surface becomes acceptable (1.9 s). However, the desorption kinetics for the Ir@PtS2 monolayer, as shown in Figure b, present a more significant challenge. At 298 K, the Ir@PtS2 exhibits prohibitively long recovery times for all the four (1.3 × 108 s ∼ 7.4 × 1025 s). Even at 398 K, the recovery times remain excessively long, indicating that effective recovery is still impractical for these gases. Consequently, the application of Ir@PtS2 would likely be limited to single-use detection.
Based on the above discussions, Rh@PtS2 exhibits a good balance of sensitivity, selectivity, and recovery, making it a strong candidate for a reusable sensor. In contrast, Ir@PtS2 is better suited for use as a high-performance, disposable sensor. While it displays superior sensitivity, evidenced by the higher adsorption energies and more significant electronic changes, its strong binding to the analytes results in prohibitively long recovery times, rendering it effectively irreversible under practical conditions. In summary, both Rh- and Ir-modified PtS2 monolayers are highly promising for detecting LDBs. The choice between them depends on the specific application requirements: Rh@PtS2 for reusable sensing applications, and Ir@PtS2 for single-use diagnostic tools.
Finally, the adsorption and sensing performance of C2H6O and C3H6O on the Rh@PtS2 surface are systematically compared with previously reported 2D materials (Table ). For the C2H6O, materials such as Pt-MoSe2 and Au-MoSe2 exhibit excessively strong adsorption strength, with E ads values of −1.56 and −1.52 eV, respectively, resulting in theoretical recovery times on the order of 1010 s at 348 K. Such prolonged recovery durations render these materials impractical for reversible sensing applications. Conversely, Ag-WTe2 exhibits overly rapid recovery due to its weaker adsorption energy (−0.57 eV). In contrast, the Rh@PtS2 monolayer achieves a balanced adsorption energy (−0.79 eV), which induces substantial bandgap modulation (ΔE g = 16.69%) while enabling rapid recovery (0.27 s at 348 K). For C3H6O, Pd-SnS2 and Ag-WTe2 similarly suffer from insufficient adsorption strength, resulting in undetectable signals due to excessively short recovery times. Meanwhile, Ru-PtTe2 and Al-MoSe2 exhibit overly high affinity with C3H6O, severely compromising their reusability. Overall, the favorable adsorption energy and moderate recovery of Rh@PtS2 position it as a highly promising material for the development of reliable, reusable C2H6O and C3H6O sensors.
2. Comparative Analysis of Ethanol (C2H6O) and Acetone (C3H6O) Adsorption and Sensing on Rh@PtS2 Versus Other 2D Materials.
| target gases | materials | E ads, eV | △E g, % | △Φ, % | τ, s (348 K) | ref. |
|---|---|---|---|---|---|---|
| C2H6O | Ag-WTe2 | –0.57 | 100 | –0.79 | 1.79 × 10–4 | |
| Pt-MoSe2 | –1.56 | 3.85 × 1010 | ||||
| Au-MoSe2 | –1.52 | 1.01 × 1010 | ||||
| Rh@PtS2 | –0.79 | 16.69 | –16.92 | 0.27 | this work | |
| C3H6O | Pd-SnS2 | –0.53 | –25.96 | 4.71 × 10–5 | ||
| Ag-WTe2 | –0.40 | –5.88 | 6.18 × 10–7 | |||
| Ru-PtTe2 | –1.12 | 0.16 | 1.64 × 104 | |||
| Al-MoSe2 | –1.79 | –23.51 | 8.22 × 1013 | |||
| Rh@PtS2 | –0.87 | –5.94 | 7.18 | 3.94 | this work |
4. Conclusions
Employing DFT calculations, this work systemically investigates the adsorption properties of four LDBs on the Rh@PtS2 and Ir@PtS2 monolayers, with a detailed analysis of their electronic structures through TED, CDD, and DOS. The main conclusions are as follows:
-
1.
The pristine PtS2 monolayer exhibits weak physisorption interactions with the four LDBs, characterized by low adsorption energies (−0.27 to −0.57 eV) and minimal charge transfer. However, decoration with a single Rh or Ir atom creates highly active sites that dramatically enhance its adsorption capabilities.
-
2.
The enhanced adsorption capacity of TM@PtS2 (TM = Rh, Ir) systems is attributed to strong orbital hybridization between the Rh-4d/Ir-5d orbitals of the dopant atoms and the O-2p/C-2p orbitals of the LDBs. As a result, C2H6O and C3H8O are physisorbed onto both substrates, whereas C3H6O and C5H8 undergo chemisorption.
-
3.
Both Rh@PtS2 and Ir@PtS2 exhibit remarkable sensitivity, as evidenced by substantial changes in their band structures and work functions upon LDBs adsorption. Notably, the adsorption of C5H8 on Rh@PtS2 and C3H8O on Ir@PtS2 induces a semiconductor-to-metal transition.
-
4.
The evaluation of selectivity against common exhaled breath components (H2O, N2, CO2, and CH4) demonstrates that both Rh@PtS2 and Ir@PtS2 possess a pronounced preferential affinity for the target LDBs. This high selectivity is crucial for reliable detection in realistic and complex gaseous environments.
-
5.
Rh@PtS2 exhibits a reasonable recovery time (22.7 s) for the C2H6O at 298 K. It also exhibits suitable recovery time for C3H6O at 348 K and for C3H8O at 398 K, suggesting its potential as a reusable sensor for these three gases. In contrast, while Ir@PtS2 displays superior sensitivity, it binds the analytes too strongly, making it more suitable as a disposable sensor.
Acknowledgments
This work was carried out in National Supercomputing Center in Shenzhen.
The authors declare no competing financial interest.
References
- Moon A. M., Singal A. G., Tapper E. B.. Contemporary Epidemiology of Chronic Liver Disease and Cirrhosis. Clin. Gastroenterol. H. 2020;18:2650–2666. doi: 10.1016/j.cgh.2019.07.060. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Berasain C., Arechederra M., Argemí J., Fernández-Barrena M. G., Avila M. A.. Loss of liver function in chronic liver disease: An identity crisis. J. Hepatol. 2023;78:401–414. doi: 10.1016/j.jhep.2022.09.001. [DOI] [PubMed] [Google Scholar]
- Shan T., Ran X., Li H., Feng G., Zhang S., Zhang X., Zhang L., Lu L., An L., Fu R., Sun K., Wang S., Chen R., Li L., Chen W., Wei W., Zeng H., He J.. Disparities in stage at diagnosis for liver cancer in China. J. Nation. Cancer Cen.J. 2023;3:7–13. doi: 10.1016/j.jncc.2022.12.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Adams L. A., Sanderson S., Lindor K. D., Angulo P.. The histological course of nonalcoholic fatty liver disease: a longitudinal study of 103 patients with sequential liver biopsies. J. Hepatol. 2005;42:132–138. doi: 10.1016/j.jhep.2004.09.012. [DOI] [PubMed] [Google Scholar]
- Zhang W. X., Yan H. M., He C.. g-C6N6 monolayer: A highly sensitive molecule sensor for biomarker volatiles of liver cirrhosis. Appl. Surf. Sci. 2021;566:150716. doi: 10.1016/j.apsusc.2021.150716. [DOI] [Google Scholar]
- Mojumder S., Das T., Das S., Chakraborty N., Saha D., Pal M.. Y and Al co-doped ZnO-nanopowder based ultrasensitive trace ethanol sensor: A potential breath analyzer for fatty liver disease and drunken driving detection. Sens. Actuators B Chem. 2022;372:132611. doi: 10.1016/j.snb.2022.132611. [DOI] [Google Scholar]
- Hanouneh I. A., Zein N. N., Cikach F., Dababneh L., Grove D., Alkhouri N., Lopez R., Dweik R. A.. The Breathprints in Patients With Liver Disease Identify Novel Breath Biomarkers in Alcoholic Hepatitis. Clin. Gastroenterol. H. 2014;12:516–523. doi: 10.1016/j.cgh.2013.08.048. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Alkhouri N., Cikach F., Eng K., Moses J., Patel N., Yan C., Hanouneh I., Grove D., Lopez R., Dweik R.. Analysis of breath volatile organic compounds as a noninvasive tool to diagnose nonalcoholic fatty liver disease in children. Eur. J. Gastroen. Hepat. 2014;26:82–87. doi: 10.1097/MEG.0b013e3283650669. [DOI] [PubMed] [Google Scholar]
- Dadamio J., Van den Velde S., Laleman W., Van Hee P., Coucke W., Nevens F., Quirynen M.. Breath biomarkers of liver cirrhosis. J. Chromatogr. B. 2012;905:17–22. doi: 10.1016/j.jchromb.2012.07.025. [DOI] [PubMed] [Google Scholar]
- Solga S. F., Alkhuraishe A., Cope K., Tabesh A., Clark J. M., Torbenson M., Schwartz P., Magnuson T., Diehl A. M., Risby T. H.. Breath biomarkers and non-alcoholic fatty liver disease: Preliminary observations. Biomarkers. 2006;11:174–183. doi: 10.1080/13547500500421070. [DOI] [PubMed] [Google Scholar]
- Sehnert S. S., Long J., Burdick J. F., Risby T. H.. Breath biomarkers for detection of human liver diseases: preliminary study. Biomarkers. 2002;7:174–187. doi: 10.1080/13547500110118184. [DOI] [PubMed] [Google Scholar]
- Saito N., Haneda H., Watanabe K., Shimanoe K., Sakaguchi I.. Highly sensitive isoprene gas sensor using Au-loaded pyramid-shaped ZnO particles. Sens. Actuators B Chem. 2021;326:128999. doi: 10.1016/j.snb.2020.128999. [DOI] [Google Scholar]
- Wu P., Huang M.. Mechanism of adsorption and gas-sensing of hazardous gases by MoS2 monolayer decorated by Pdn (n = 1–4) clusters. Colloid. Surface. A. 2024;695:134200. doi: 10.1016/j.colsurfa.2024.134200. [DOI] [Google Scholar]
- Sanyal G., Kandasamy M., Mondal B., Seetharaman A., Chakraborty B.. Pd-anchored VSe2 for glucose sensing: Prediction from first principle simulations. Sur. Interfaces. 2024;54:105235. doi: 10.1016/j.surfin.2024.105235. [DOI] [Google Scholar]
- Yang X., Duan Y., Qin Z., Wang J.. A DFT study of gas molecules adsorption on TM-doped PtS2 gas nanosensors. Phys. Scr. 2024;99:115401–115415. doi: 10.1088/1402-4896/ad7d44. [DOI] [Google Scholar]
- Cui H., Feng Z., Wang W., Peng X., Hu J.. Adsorption Behavior of Pd-Doped PtS2 Monolayer Upon SF6 Decomposed Species and the Effect of Applied Electric Field. IEEE Sens. J. 2022;22:6764–6771. doi: 10.1109/JSEN.2022.3154119. [DOI] [Google Scholar]
- S A. G., Kong I., Kanungo S.. Comparative study of different volatile organic compound sensing in pristine and defective PtS2 monolayer-An Ab-initio calculation-based theoretical perspective. Sur. Interfaces. 2025;58:105828. doi: 10.1016/j.surfin.2025.105828. [DOI] [Google Scholar]
- Yang M., Xiong H., Ma Y., Yang L.. Theoretical investigation of Ag and Au modified CSiN monolayer as a potential gas sensor for air decomposition components detection. J. Mol. Liq. 2024;410:125648. doi: 10.1016/j.molliq.2024.125648. [DOI] [Google Scholar]
- Ma Y., Yang M., Deng G., Xiong H.. Theoretical insights into gas-sensitive properties of B, Ga and In doped WS2 monolayer towards oxygen-containing toxic gases. Appl. Surf. Sci. 2024;670:160725. doi: 10.1016/j.apsusc.2024.160725. [DOI] [Google Scholar]
- Zeng Q., Wang M., Zhang Y., Liu J., Xu M., Chen D., Jia P.. Machine learning-assisted development of TMDs-type gas-sensitive materials for dissolved gases in oil-immersed transformer oils. Mater. Today Chem. 2025;44:102583. doi: 10.1016/j.mtchem.2025.102583. [DOI] [Google Scholar]
- Xu Z.. Ni-Decorated PtS2 Monolayer as a Strain-Modulated and Outstanding Sensor upon Dissolved Gases in Transformer Oil: A First-Principles Study. ACS Omega. 2023;8:6090–6098. doi: 10.1021/acsomega.3c00022. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yao W., Guan H., Zhang K., Wang G., Wu X., Jia Z.. Nb-doped PtS2 monolayer for detection of C2H2 and C2H4 in on-load tap-changer of the oil-immersed transformers: A first-principles study. Chem. Phys. Lett. 2022;802:139755. doi: 10.1016/j.cplett.2022.139755. [DOI] [Google Scholar]
- Xiong H., Zhang S., Ma Y., Zhang Y., Huang H., Li J., Sun C., Zhong X.. First-principles study of WS2 monolayer decorated with noble metals (Pt, Rh, Ir) as the promising candidates to detect nitrogenous poisonous gases. Sur. Interfaces. 2024;55:105363. doi: 10.1016/j.surfin.2024.105363. [DOI] [Google Scholar]
- Xu H., Li L., Xia J., Zeng W., Zhou Q.. Innovative doping strategies in ZrSe2: Co and Rh for enhanced agricultural greenhouse gas adsorption. Sur. Interfaces. 2025;60:105988. doi: 10.1016/j.surfin.2025.105988. [DOI] [Google Scholar]
- Jiang W., Zeng Q., Zhang Y., Wang G., Sun K., Xu M., Chen D., Jia P.. Adsorption and sensing characteristics of transition metal (Au, Fe, Pt, Rh and Cu) doped WS2 on dissolved gases in transformer oil: A DFT perspective. Comput. Theor. Chem. 2025;1251:115327. doi: 10.1016/j.comptc.2025.115327. [DOI] [Google Scholar]
- Delley B.. From molecules to solids with the DMol3 approach. J. Chem. Phys. 2000;113:7756–7764. doi: 10.1063/1.1316015. [DOI] [Google Scholar]
- Hammer B., Hansen L. B., No̷rskov J. K.. Improved adsorption energetics within density-functional theory using revised Perdew-Burke-Ernzerhof functionals. Phys. Rev. B. 1999;59:7413–7421. doi: 10.1103/PhysRevB.59.7413. [DOI] [Google Scholar]
- Perdew J. P., Burke K., Ernzerhof M.. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996;77:3865–3868. doi: 10.1103/PhysRevLett.77.3865. [DOI] [PubMed] [Google Scholar]
- Vozzi C., Negro M., Calegari F., Sansone G., Nisoli M., De Silvestri S., Stagira S.. Generalized molecular orbital tomography. Nat. Phys. 2011;7:822–826. doi: 10.1038/nphys2029. [DOI] [Google Scholar]
- Inada Y., Orita H.. Efficiency of numerical basis sets for predicting the binding energies of hydrogen bonded complexes: Evidence of small basis set superposition error compared to Gaussian basis sets. J. Comput. Chem. 2008;29:225–232. doi: 10.1002/jcc.20782. [DOI] [PubMed] [Google Scholar]
- Grimme S.. Semiempirical GGA-type density functional constructed with a long-range dispersion correction. J. Comput. Chem. 2006;27:1787–1799. doi: 10.1002/jcc.20495. [DOI] [PubMed] [Google Scholar]
- Sellers H.. The C2-DIIS convergence acceleration algorithm. Int. J. Quantum Chem. 1993;45:31–41. doi: 10.1002/qua.560450106. [DOI] [Google Scholar]
- Monkhorst H. J., Pack J. D.. Special points for Brillouin-zone integrations. Phys. Rev. B. 1976;13:5188–5192. doi: 10.1103/PhysRevB.13.5188. [DOI] [Google Scholar]
- Cui H., Wu H., He D., Ma S.. Noble metal (Pd, Pt)-functionalized WSe2 monolayer for adsorbing and sensing thermal runaway gases in LIBs: A first-principles investigation. Environ. Res. 2025;269:120847. doi: 10.1016/j.envres.2025.120847. [DOI] [PubMed] [Google Scholar]
- Chen Y., Jiang W., Zhang Y., Chen D., Xu M., Liu J., Jia P.. Gas sensitivity study of early diagnostic markers for lung cancer using MoTe2 single molecular membranes doped with different TM atoms: Based on density functional theory. Colloid. Surface. A. 2025;722:137289. doi: 10.1016/j.colsurfa.2025.137289. [DOI] [Google Scholar]
- Ma Y., Deng G., Xiong H.. Density function theory investigation of bimetallic phthalocyanine as a potential sensor and scavenger for nitrogen-containing toxic gases. J. Environ. Chem. Eng. 2024;12:113687. doi: 10.1016/j.jece.2024.113687. [DOI] [Google Scholar]
- Wu Y., Li X., Duan Y., Pan Y., Yuan J., Zhang X.. Adsorption and sensing of SF6 decomposition gas molecules by Ni-InN monolayer: A first-principles study. Mater. Sci. Semicond. Process. 2025;188:109137. doi: 10.1016/j.mssp.2024.109137. [DOI] [Google Scholar]
- Zhao W., Wang J., Shen C., Xie B., Li G., An T.. Mechanistic insights into the adsorption of different types of VOCs on monolayer MoS2 via first-principles approaches. Environ. Scie. Nano. 2025;12:1230–1240. doi: 10.1039/D4EN00953C. [DOI] [Google Scholar]
- Liu H., Luo X.. Au- and Pd-Doped SnS2 Monolayers for Lung Cancer Biomarkers (C3H6O, C6H6, and C5H8) Detection: A Density Functional Theory Investigation. ACS Omega. 2024;9:7658–7667. doi: 10.1021/acsomega.3c06346. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gilani R., Alarfaji S. S., Nadeem K., Saeed A., Isa Khan M.. Pristine and Aurum-decorated tungsten ditellurides as sensing materials for VOCs detection in exhaled human breath: DFT analysis. RSC Adv. 2024;14:26788–26800. doi: 10.1039/D4RA04569F. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhou L., Yu S., Yang Y., Li Q., Li T., Zhang D.. Adsorption of gas molecules (NH3, C2H6O, C3H6O, CO, H2S) on a noble metal (Ag, Au, Pt, Pd, Ru)-doped MoSe2 monolayer: a first-principles study. New J. Chem. 2021;45:12367–12376. doi: 10.1039/D1NJ01705E. [DOI] [Google Scholar]
- Hafeez J., Usama Islam M., Ali S. M., Khalid S., Ashraf N., khan M. I.. Computational exploring the potential of pure and Ag-decorated WTe2 for detecting volatile organic compounds (VOCs) Mater. Sci. Semicond. Process. 2024;182:108710. doi: 10.1016/j.mssp.2024.108710. [DOI] [Google Scholar]
- Wan Q., Chen X., Xiao S.. Ru-Doped PtTe2 Monolayer as a Promising Exhaled Breath Sensor for Early Diagnosis of Lung Cancer: A First-Principles Study. Chemosensors. 2022;10:428–440. doi: 10.3390/chemosensors10100428. [DOI] [Google Scholar]
- Liu T., Cui Z., Li X., Cui H., Liu Y.. Al-Doped MoSe2 Monolayer as a Promising Biosensor for Exhaled Breath Analysis: A DFT Study. ACS Omega. 2021;6:988–995. doi: 10.1021/acsomega.0c05654. [DOI] [PMC free article] [PubMed] [Google Scholar]














