A first principles study of H2S adsorption and decomposition on a Ge(100) surface

Tsung-Fan Tengab, Santhanamoorthi Nachimuthua, Wei-Hsiu Hungb and Jyh-Chiang Jiang*a
aDepartment of Chemical Engineering, National Taiwan University of Science and Technology, Taipei 106, Taiwan. E-mail: jcjiang@mail.ntust.edu.tw; Fax: +886-2-27376644; Tel: +886-2-27376653
bDepartment of Chemistry, National Taiwan Normal University, Taipei 116, Taiwan

Received 19th August 2014 , Accepted 5th December 2014

First published on 5th December 2014


Abstract

We employed density functional theory (DFT) calculations to examine the adsorption configurations and possible reaction paths for H2S on a Ge(100) surface. There are four reaction paths proposed for the decomposition of adsorbed H2S on a Ge(100) surface and the corresponding structural conformations are studied extensively. The present study shows two new possible products and a detailed reaction mechanism for H2S adsorption on a Ge(100) surface and the results are compared with our previous study of H2S adsorption on a Si(100) surface (J. Phy. Chem. C, 115, 2011, 19203). The density of states (DOS) and electron density difference (EDD) analyses are used to illustrate the interaction between S-containing species and surface Ge atoms.


1. Introduction

Over the past few decades, high mobility semiconductors have been investigated as a replacement for the Si channel due to their potential applications in high-performance metal-oxide-semiconductor (MOS) devices.1–3 Recently, Germanium (Ge) has attracted much attention because it gives bulk electron and hole motilities at room temperature which are higher than the conventional silicon based materials.4,5 However, major drawbacks of the application of Ge in these devices are the poor quality of its thermal oxide compared to SiO2 and the Ge-based MOS devices has high interface state density (Dit) which would result in low carrier mobility in the channel. These limitations can be overcome by the passivation of Ge surface prior to the deposition of the gate dielectric which is a key issue for semiconductor surfaces.6,7

Also, the practical use of this semiconductor in the devices requires the passivation to prevent oxidation and maintain the surface order during device processing. Sulfur (S) has been shown to be one of the best passivants of semiconductor surfaces.4 Anderson et al., reported that the sulfur passivation of the germanium surface via both solution and vacuum deposition techniques.8 A sulfur adlayer is deposited on a semiconductor surface through immersion in a solution of (NH4)2S8 or on its exposure to elemental sulfur or H2S.9 On the other hand, Houssa et al. proposed two approaches for the passivation of Ge surface to promote metal-oxide-semiconductor field-effect transistor (MOSFET).10 One of two approaches is the exposure of the Ge surface to H2S, resulting in the formation of S–Ge bonds. Indeed, the Ge energy band gap is free of surface states after H2S exposure, resulting in the electrical passivation of the surface, which is a crucial requirement for the potential use of Ge in the MOSFET.

Similar to Si(100) and C(100) surfaces, the reconstructed Ge(100)-2 × 1 surfaces consist of dimers connected by a strong σ-bond and a weak π-bond.11 The diamond surface has symmetric dimers with a bond length of approximately 1.4 Å, whereas the Si and Ge surfaces both have larger asymmetric, or tilted, dimers with a bond length of 2.3–2.5 Å.12–14 Distortion of the dimer bonds induces a charge transfer from the down (electrophilic) to the upper (nucleophilic) surface atom.15 Such zwitterionic characteristics allow the surface to undergo a nucleophilic/electrophilic reaction, often with direct analogies to the molecular systems. The adsorption of sulfur on Ge(100) surface has been studied both experimentally and theoretically.16–18

Previously, the S-passivation on Ge(100) surface has been investigated using low-energy electron diffraction (LEED),8,19 high-resolution electron-energy loss spectroscopy (EELS),20 ultraviolet photoelectron spectra (UPS),21 X-ray photoelectron spectra (XPS),22 temperature programmed desorption (TPD),23 Near edge X-ray absorption fine structure (NEXAFS)24 and multiple internal reflection-Fourier transform infrared spectrometer (MIR-IR).25 Kuhr and Ranke21 showed in their UPS study that, H2S adsorbs dissociatively at 300 K on a Ge surface; then decomposes completely to form sulfur at 550 K. To understand the sulfidation of the Ge(100) surfaces and its possible passivating reactions, we choose H2S as a passivant to be adsorbed on the Ge(100) surface. In this study, we report ab initio theoretical investigations to provide a complete description of the adsorption effects and thermal decompositions of H2S on Ge(100) surface. We thereby elucidates possible mechanisms of thermal decomposition on Ge(100) that was more conformed to real surface.

2. Computational details

DFT calculations are performed using the Vienna ab initio simulation package (VASP).26–28 The Vanderbilt ultra-soft pseudopotential is used to describe the election–ion interactions29 and the electron–electron exchange and correlation contributions by the generalized gradient approximation (GGA) with energy truncated at 300 eV. The Brillouin zone is sampled with Monkhorst-Pack grid, and the calculations were performed with (4 × 2 × 1) Monkhorst-Pack mesh k-points. We optimized the structures based on the conjugate gradient-minimization scheme taking into consideration of spin polarization. The validities of all the optimized structures and determined transition states (TSs) are checked through normal-mode frequency analysis. For a real minimum structure, all frequency must be positive; TS must have one imaginary frequency corresponding to the reaction coordinate. The clean surfaces are then modeled in the form of a slab. The c(4 × 2) unit cell consisted of nine atomic layers separated by a 20 Å vacuum to prevent interactions between the surface adsorbates and the preceding slab. The optimized geometry of the model Ge(100) surface considered in this study is depicted in Fig. 1.
image file: c4ra08887e-f1.tif
Fig. 1 The optimized structure of Ge(100)-c(4 × 2) surface: (a) front view, (b) side view and (c) top view.

In the structural optimizations, the positions of the last three Ge layers along with the H-passivated layer on the bottom are fixed upon adsorption of H2S and the remaining substrate atoms are allowed to relax with the adsorbates. To economize computing time, only the upper two atomic layers of the surface and adsorbate are relaxed in the vibrational frequency calculations. The adsorption energies are obtained by

Eads = ETotal − (EGas + ESurface)
in which ETotal, EGas and ESurface are the respective calculated electronic energies of the adsorbed species on the surface, a gaseous molecule, and a clean surface. A negative value for Eads indicates an exothermic adsorption. The relative energies (Erel) are defined as
Erel = E(LM or FS)EH2Sad
in which EH2Sad and E(LM or FS) are the calculated adsorption energies of H2S and local minima or final products, respectively. The nudged-elastic-band (NEB) method is applied to locate the transition state structures positioned to interpolate a series of system images between the initial and final states on the potential-energy surface.30,31 A spring force between adjacent images is used to maintain constant spacing between the images, and a true force is applied to impel the images into the minimum energy path (MEP), mimicking an elastic band. Each image is optimized using the NEB algorithm based on a constrained algorithm of molecular dynamics. The highest point on the MEP corresponds to a transitional structure on the proposed reaction path and its energy, relative to that of the initial state, became the activation barrier of the reaction.

The electron density difference (Qdiff) is calculated in a similar manner to the calculation of adsorption energy

Qdiff = QA/Surf − (QSurf + QA)
where Qdiff is the difference at each grid point in the total electron density matrix between that of the adsorbate-bonded surface (QA/Surf) and that of the sum of the surface (QSurf) and the single adsorbate molecule (QA). According to this definition, positive and negative values correspond to increasing and decreasing electron densities, respectively.

3. Results and discussion

3.1. Adsorption of H2S, HS and S on a Ge(100) surface

To ensure the reliability of the computational method, we have calculated the lattice parameters for bulk germanium using various pseudopotentials and the values are given in Table S1 of ESI. As can be seen from those values PAW-PBE and the ultrasoft pseudopotential with the generalized gradient approximation (US-GGA) gives the least discrepancy between the calculated and experimental lattice parameters among the tested pseudopotentials. In order to compare with our previous study of H2S adsorption on a Si(100) surface,32 we considered US-GGA method for further calculations. Also, we have performed the benchmark calculations for the H2S adsorption on the clean Ge(100) surfaces with different functionals and different cutoff energy values and the calculated values are given in Table S2. It has been observed that the trends of the calculated values with different cutoff energies and functionals are almost similar and hence we have chosen the US with 300 eV cutoff energy values for our further study. Table S3 presents the comparison of calculated bond lengths and bond angles of clean Ge(100) surface with other studies33 and it has been found that the calculated bond lengths are in agreement with the previous experimental reports.34–36

Previously, we found that the H2S has three stable adsorption conformations on the Si(100) surfaces with orientation of the S–H bond;32 similarly here we considered the same conformation of H2S adsorbed on Ge(100) surfaces (denoted as H2Sad, H2Sad-1, and H2Sad-2) and the optimized structures are shown in Fig. 2. The calculated adsorption energies and structural parameters for different conformations of H2S are summarized in Table 1. Among these three conformations, H2Sad is found to be the most stable with adsorption energy −0.49 eV and the other two conformations H2Sad-1 and H2Sad-2 adsorbed on the surface with adsorption energies −0.40 and −0.30 eV, respectively. From Table 1, it has been observed that the bond length of the Ge–Ge dimer is increased (∼0.04 Å) and the tilting angle of Ge dimer is decreased to 16° from 20° after the adsorption of H2S on the Ge surface, indicating the decreased zwitterionic and π-bond characters of the Ge dimer upon the adsorption of H2S. The calculated S–Ge bond length value for the most stable conformation (H2Sad) is 2.61 Å, and the two remaining S–H bond lengths are 1.35 and 1.38 Å. The structural parameters of other two conformations, H2Sad-1 and H2Sad-2, are similar to those of H2Sad. Fig. 3 shows the DOS of surface Ge down atom and H2S molecule before and after adsorption on Ge(100) surface, respectively. It is noticed from this figure that the H2S molecule possesses C2v symmetry with four valence states labelled as 4a1, 2b2, 5a1, and 2b1 according to their orbital symmetries. These results are similar to that of our previous study.32 The partial DOS (see Fig. S1 of ESI) indicates that states 4a1 and 2b1 are contributed mainly from 3s and 3pz of S atom and 2b2 and 5a1 states arise from the hybridization of orbitals 3px and 3py of the S atom. All S 3p atomic orbitals (2b2, 5a1, and 2b1 states), especially 3px and 3py of adsorbed H2S have good overlaps with d orbitals of the buckled-down Ge which is bonded to the H2S molecule.


image file: c4ra08887e-f2.tif
Fig. 2 The optimized structures of different conformations of H2S adsorbed on Ge(100) surface: (a) H2Sad, (b) H2Sad-1, and (c) H2Sad-2.
Table 1 The calculated structural parameters and adsorption energies (Eads) for hydrogen sulphide (H2S) adsorption on Ge(100) surface
System d(S–H) (Å) d(S–Ge) (Å) d(Ge–Ge) (Å) ∠HSH (deg) Ge–Ge tilting angle (deg) Eads (eV) Eadsa (eV)
a The calculated adsorption energy for H2S adsorption on Si(100) surface taken from previous study.32
H2Sad 1.35/1.38 2.61 2.58 92.4 16 −0.49 −0.74
H2Sad-1 1.35/1.37 2.64 2.58 91.9 17 −0.40 −0.70
H2Sad-2 1.33/1.39 2.63 2.57 92.5 16 −0.30 −0.65



image file: c4ra08887e-f3.tif
Fig. 3 The DOS of a H2S molecule before (dashed line) and after (black solid line) adsorption on Ge(100) and DOS of Ge down atom (bond black line) bonded with H2S.

3.2. Reaction mechanism of H2S on Ge(100)

We consider the reaction pathway for adsorption of H2S on Ge(100) surface analogous to our previous study of the adsorption of H2S on Si(100).32 The calculated structural parameters and relative energies of intermediates and final products are listed in Table 2. The reaction energies, reaction barriers, and calculated imaginary frequencies of transition state structures are summarized in Table 3. The corresponding structures of intermediates and final products are shown in Fig. 4 and the top views of their corresponding transition state structures are shown in Fig. S2 of ESI. The potential energy diagram (PES) for the adsorption and dissociation of H2S on the Ge(100) surface are shown in Fig. 5. According to the calculation, the possible reaction paths are summarized in the following scheme:
image file: c4ra08887e-u1.tif
Table 2 The calculated structural parameters and relative energies for dissociatively adsorbed H2S on Ge(100) surface
Local minima and final products d(Ge–Ge)a (Å) d(Ge–Ge)b (Å) d(S–Ge) (Å) d(H–Ge) (Å) Ereld (eV)
a Bond length of a Ge(1)–Ge(2) dimer which is bonded either with HS or S.b Bond length of a Ge(3)–Ge(4) or Ge(5)–Ge(6) which is bonded with dissociative H.c Taken from ref. 18.d Relative energy for H2S on Si(100) surface are given in parenthesis.
LM1I 2.55 2.58 2.28 1.57 −0.89 (−0.95)
LM1II 2.51 (2.46c) 2.51 (2.46c) 2.28 (2.18c) 1.55 (1.56c) −1.03 (−1.43)
LM1III 2.60 2.53 2.29 1.54 −0.96 (−1.31)
LM2I 2.52 2.58 2.10 1.55 −0.67 (−1.05)
LM2II 2.52 2.54 2.11 1.55 −0.67 (−1.33)
LM3I 2.56 2.59 2.11 1.55 −0.16 (−1.13)
LM3II 2.62 2.48 2.12 1.54 −1.03 (−1.45)
LM3III 2.54 2.58 2.11 1.56 −1.19 (−1.45)
LM3IV 3.88 2.57 2.24/2.33 1.55 −0.88 (−1.46)
LM3V 3.45 2.55 2.21/2.34 1.55 −1.12 (−1.59)
FSI 2.74 2.73 2.36 1.56 −0.81 (−1.38)
FSII 2.52 2.47 2.17 (2.16c) 1.54 (1.56c) −1.27 (−2.63)
FSIII 2.51 2.53 2.33 1.55 −1.34 (−2.12)
FSIV 3.56 2.25/2.28 1.55 −1.21 (−1.99)


Table 3 The calculated reaction barriers (E, eV), reaction energies (ΔE, eV), and imaginary frequencies (IMF, cm−1) for the transition states of decomposition of H2S on Ge(100) surface
Reaction path E ΔE IMF
H2Sad → TS1I → LM1I 0.53 −0.89 227i
H2Sad → TS1II → LM1II 0.49 −1.03 661i
H2Sad → TS1III → LM1III 0.10 −0.96 300i
LM1I → TS2I-a → LM2I 0.90 0.22 492i
LM1II → TS2I-b → LM2I 0.88 0.36 419i
LM1III → TS2II-a → LM2II 1.89 0.29 1029i
LM1II → TS2II-b → LM2II 1.07 0.36 421i
LM2I → TS3I → LM3I 1.26 0.51 552i
LM2II → TS3II → LM3II 1.71 −0.36 405i
LM2II → TS3III → LM3III 1.24 −0.52 274i
LM2I → TS3IV → LM3IV 0.47 −0.32 128i
LM2II → TS3V → LM3V 0.30 −0.46 197i
LM3I → TS4I → FSI 0.48 −0.65 145i
LM3II → TS4II → FSII 1.20 −0.10 119i
LM3III → TS4III → FSIII 0.58 −0.01 162i
LM3IV → TS4IV → FSIV 1.13 −0.22 919i
LM3V → TS4V → FSIV 1.43 −0.09 347i



image file: c4ra08887e-f4.tif
Fig. 4 Top view of all the local minima's and the final products of the reactions for the adsorption of H2S on Ge(100) surface.

image file: c4ra08887e-f5.tif
Fig. 5 Potential-energy diagram for adsorption and decomposition of H2S on Ge(100) surface. All potential energies of intermediates, transition structures, and final products are referred to the H2S molecule adsorbed on the surface.

The possible reaction pathways described in the above scheme includes first dehydrogenation, second dehydrogenation and sulfur bridged adsorption. The H2Sad species can undergo a first dehydrogenation through three paths. The first path is through TS1I, with a 0.53 eV barrier, to produce LM1I intermediate; i.e., one H atom of H2S dissociated and migrated to the Ge atom of an adjacent dimer along [001] direction. It has been observed that the first dehydrogenation has higher energy barrier due to the H atom being adsorbed on the Ge(3) atom. The second path creates the intermediate LM1II via TS1II with a barrier of 0.49 eV; i.e., the dissociated H and HS separately adsorb on each of a dimer's Ge atom. The third path yields intermediate LM1III via TS1III with the smallest barrier 0.10 eV; the resulting H and HS adsorb on Ge atoms of adjacent dimers along direction [010]. Fig. 6a and b show the EDD contour plots which contain the dimeric S–Ge(1) bond with S–Ha and S–Ge(1) bond with S–Hb bond, respectively. It can be seen from Fig. 6a that the electron density increases between the S and Ge(1) atoms which is bonded to H2S, indicating a strengthening of S–Ge bond. In the meanwhile, the increased electron density is observed between the H and Ge(5) atoms of a neighboring Ge dimer along [010] direction (Fig. 6a) which interacts with each other through a hydrogen bond. This interaction results in the lower energy barrier (0.10 eV) in the first dehydrogenation for H2Sad to form LM1III. As summarized in Table 3, the reaction energies of the first dehydrogenations for the three paths are −0.89, −1.03, and −0.96 eV, respectively. LM1II and LM1III are comparably stable because the first dissociated H preferably adsorbs on the electron rich atoms like Ge(2) and Ge(5) atoms. The S–Ge bond length in LM1I, LM1II, and LM1III are shortened by ∼0.3 Å, compared to that of in the H2Sad. In H2Sad, H2S adsorbed on a Ge atom via a dative bond; whereas in LM1I∼III, it is adsorbed on a Ge atom via a covalent bond.


image file: c4ra08887e-f6.tif
Fig. 6 EDD contours of H2Sad on a plane containing (a) a S–Ge(1) bond and a dimeric S–Ha bond along [010] and (b) a S–Ge(1) bond and a dimeric S–Hb bond along [001]. The red and blue colors represent the increasing and decreasing electron densities, respectively.

Further, the local minima's LM1I, and LM1II can undergo dehydrogenation and form LM2I via the transition states, TS2I-a and TS2I-b with energy barriers of 0.90 and 0.88 eV, respectively. LM2II can be produced from the dehydrogenation of LM1III and LM1II via transition states, TS2II-a, and TS2II-b with barriers of 1.89 and 1.07 eV, respectively. From Table 2, it has been observed that the structural parameters of LM2I and LM2II are almost similar. The S–Ge bond of LM2 species is much shorter than that of LM1, indicating a stronger S–Ge bond in species LM2. The energy barriers of the second dehydrogenation, LM1 → LM2, are higher than that of first dehydrogenation H2Sad → LM1. In the reaction mechanism, there are two types of intermediates between the LM2 and the final products; one is formed through the migration of H atom from LM2I/LM2II (LM3I, LM3II and LM3III) and another one is formed through the bridging of two Ge dimer atoms via S adatom (LM3IV and LM3V). In the first type, the transition states TS3I, TS3II, and TS3III are having the energy barriers of 1.26, 1.71, and 1.24 eV, respectively between the LM2I–II and LM3I–III. Also, the S adatom remains bonded with Ge atom and has a dangling bond in LM3I–III. The final products, FSI∼III are formed from LM3I–III via TS4I–III with respective energy barriers of 0.48, 1.20, and 0.58 eV. LM3I–III can further transform with small barriers to the final products, FSI–III. In FSI–III, the S adatom bridges either two neighboring dimers along [001] direction or two Ge atoms of a dimer or two neighboring dimers along [010] direction, respectively. In LM3IV and LM3V, the S adatom bridges two Ge atoms of a dimer (Fig. 4i and j) and their formation barriers by corresponding transition states, TS3IV and TS3V are 0.47 eV and 0.30 eV, respectively. Further, LM3IV and LM3V isomerized to form the final product FSIV (Fig. 4n), with high barriers (1.13 and 1.43 eV), respectively. From Table 2, it has been noticed that the dimeric Ge–Ge bond length remains nearly at 2.56 ± 0.06 Å in all the intermediates and final products, except LM3IV, LM3V, FSI and FSIV. In the FSI, the Ge(1)–Ge(2) bond length increases to 2.74 Å because the Ge–S–Ge has ring stress. Also, the Ge(1)–Ge(2) bond is cleaved in LM3IV, LM3V and FSIV, so the bond length increases to 3.5–3.8 Å.

There are four reaction paths for the dissociation of H2S on Ge(100) surface are proposed and they are; H2Sad → LM1I/LM1II → LM2I → LM3I → FSI, H2Sad → LM1II/LM1III → LM2II → LM3II → FSII, H2Sad → LM1II/LM1III → LM2II → LM3III → FSIII, and H2Sad → LM1II → LM2I/LM2II → LM3IV/LM3V → FSIV. However from the calculated results of the rate-determining steps (RDS) for the final products (FSI, FSII, FSIII and FSIV) are LM2I → LM3I, LM2II → LM3II, LM2II → LM3III and LM3V → FSIV, respectively. Their barriers at the rate-determining step are 1.26, 1.71, 1.24 and 1.43 eV, respectively. Among the four, the third RDS has the lowest energy barrier. According to these calculations, FSIII is thermodynamically more stable than the others and hence it can be the major product. The more ring strains in FSI and FSIV due to their inter- or intra-dimer lead to break the dimer bonds, so they are unstable products among the others.23 From the above, it has been noticed that the final product FSIII is favorable in terms of both kinetic and thermodynamic points.

From Table 1, it has been found that the calculated adsorption energy values for three possible adsorption conformations of H2S on Ge(100) surface are smaller than that of Si(100) surface. When H2S adsorbs on surface, the 3p orbitals (2b2, 5a1, and 2b1) of S atom have more significant overlap with p orbitals of buckled-down Si atom32 than the d orbitals of buckled-down Ge atom. The strong overlap between the p-orbitals of S and Si atoms result in the large adsorption energies for the adsorption of H2Sad on Si(100) surface.32 From Table 2, it is noticed that the relative energies for all the intermediates and final products of adsorption of H2S on Si(100) surface are more negative than those on Ge(100) surface, which indicates that those are more stable on the Si surface. Based on the present and previous studies,32 we summarized the total reaction energies for final products, energy barriers for RDS and RDS pathways from H2S adsorbed on Ge or Si surface and are shown in Table 4. As can be seen from this table, FSII and FSIII are thermodynamically favourable products for the H2S adsorption on Ge(100) and Si(100), respectively. Their RDSs are from LM2 to LM3 via second dehydrogenation process. In the kinetic point of view, FSI and FSIV are the favorable products for Si surface and FSI and FSIII are the major products for Ge surface. From the above results, it is noted that the final product FSIII is thermodynamically and kinetically favorable product for the H2S adsorption on Ge surface. From the above results it has been observed that the S-passivation products on Si(100) surface are more stable than Ge(100) surface.

Table 4 The calculated total reaction energies (ΔE in eV) for the final products, energy barrier (E in eV) for RDSs and their corresponding pathways for the adsorption of H2S on Ge and Si surface
Final products ΔE E RDS
Ge Si Ge Si Ge Si
FSI −0.81 −1.38 1.26 1.19 LM2I → LM3I LM2I → LM3I
FSII −1.27 −2.63 1.71 1.63 LM2II → LM3II LM2II → LM3II
FSIII −1.34 −2.12 1.24 1.54 LM2II → LM3III LM2II → LM3III
FSIV −1.21 −1.99 1.43 1.22 LM3V → FSIV LM3V → FSIV


3.3. Density of states (DOS)

The electronic structure is intimately related to their fundamental physical and chemical properties. To further understand the adsorption modes of H2S on Ge(100), we analyzed DOS of most possible reaction pathway i.e., H2Sad → LM1II → LM2II → LM3III → FSIII, which includes five different S conformations such as H2S, HS, S adatom and bridged S atoms. The plotted DOS for H2Sad, LM1II, LM2II, LM3III and FSIII are shown in Fig. 7 and it compares the different S conformations adsorbed on Ge(100) surface. In the Fig. 7, the sharp band appears between −10 to −15 eV relative to the Fermi level is attributed from the 3 s orbital of S atoms, which is adsorbed on the surface with different configurations. Initially, when the H2S species absorbs on the Ge(100) surface, this band located at −15.86 eV region (Fig. 7a) and the first and second dehydrogenation shifts this band towards the Fermi level (from −15.86 to −13.5/−11.8 eV, see Fig. 7b and c). This upward shift can be explained that the interaction of H2S on the surface via dative bond, whereas after the first dehydrogenation, LM1II species interact with surface via single bond. Also, the peak at −8.2 eV region (Fig. 7a) is due to the px orbital of S atom shifts close to the Fermi level after the first dehydrogenation and finally it overlaps with the py and pz orbitals after the second dehydrogenation, which is due to the interaction of S atom with the surface by either dative or covalent bond. In LM2II and LM3III, S atom adsorbed in a similar way and hence, the PDOS for adsorbed S atom resembles same (at −11.89 and −12.67 eV respectively in Fig. 7c and d). Finally, the FSIII possesses bridged S conformation on Ge(100) surface, so its 3s orbital band shifted downward and close to the HS conformation. Also, the broad band in the region of 0 to −5 eV, shows the bonding interaction of p orbitals of S and Ge surface atoms. These above results demonstrate that DOS spectra can comprehensively explain the interaction of adsorbed S atom and surface Ge atom.
image file: c4ra08887e-f7.tif
Fig. 7 The total DOS of S adatom in (a) H2Sad, (b) LM1II, (c) LM2II, (d) LM3III, and (e) FSIII structures.

3.4. Vibrational frequency analysis for H2S adsorption on Ge(100)

In order to guide a future surface vibrational spectroscopic study, the vibrational frequencies for adsorbed species on Ge(100) surface are analyzed. The calculated vibrational frequencies for the adsorbed species such as, H2S, LM, and FS species are summarized in Table 5. The bands observed at 2577 and 2159 cm−1 for H2S species are due to asymmetric and symmetric S–H stretching modes. The strong interaction between Hb and Ge(5) atoms (see Fig. 6) leads to weaken the S–Hb bond resulted in redshift of wave number for S–H stretching mode (2159 cm−1). The H–Ge stretching bands are observed in the range of 1900–2024 cm−1. The calculated S–H and H–Ge stretching modes of LM1II are in agreement with the previous EELS experimental values of 2532 cm−1 and 1967 cm−1.37 For S–Ge stretching bands, LM1I∼III belong to the first dissociated group which display similar S–Ge stretching vibration on the surface. There is about 80 cm−1 difference between the H2Sad and LM1I–III for S–Ge stretching mode, which is due to the change in interaction type, i.e. from dative bond to covalent bond between sulfur atom and Ge surface. The other dissociated intermediates (LM2I–II and LM3I–III) exhibits blue shift in the IR spectrum, which is due to the increase in the bond order of S–Ge after the second dehydrogenation. The S–Ge stretching bands for LM3IV–V and FSI–IV are observed at about 300–420 cm−1 which is red shifted due to the interaction of bridged sulfur atom with the Ge atom via covalent bond. As shown in Table 5, the calculated S–Ge vibration frequencies of FSI are red shifted (about 50 to 130 cm−1) compared to other final products, which indicates that the bond of dimeric Ge become longer because of the ring stress from the Ge–S–Ge bonds.
Table 5 The calculated vibrational frequencies of the initial state (H2Sad), local minima's (LM), and final products (FS) for the decomposition of H2S on Ge(100) surfaces
System νH–S (cm−1) νH–Ge (cm−1) νS–Ge (cm−1)
H2Sad 2577/2159   265
LM1I 2600 1926 342
LM1II 2544 1967 331
LM1III 2389 2024 348
LM2I   1962/1915 456
LM2II   1979/1968 461
LM3I   1970/1921 463
LM3II   2000/1968 448
LM3III   1948/1927 450
LM3IV   1913/1908 351/310
LM3V   1934/1916 364/295
FSI   1914/1903 291/282
FSII   1998/1994 421/267
FSIII   1991/1983 344/276
FSIV   1951/1923 396/333


4. Conclusions

We have performed the detailed investigation of the adsorption and dissociation of H2S on Ge(100) surface using density functional theory calculations. Our results show that among the three different conformation, H2Sad is found to be adsorbed stably on the surface. Possible reaction paths are proposed for dehydrogenation between adsorbed H2S and S adatoms. In the first step, adsorbed H2S partially dissociates and produced HS and H and in the second step, HS further dissociates into S adatom. Further, the EDD contour indicate that the increased electron density between the H atom and Ge(5) atom of a neighboring Ge dimer along [010] which leads to the lower energy barrier (0.10 eV) in the first dehydrogenation for H2Sad to LM1III. The results show that the four final products have more stable conformations via LM1, LM2, and LM3 intermediates. We found that FSI and FSIII are kinetically favored products whereas FSII and FSIII are thermodynamically favored products for the H2S adsorption on the Ge surface. In addition, the interactions between S and Ge atoms are identified by DOS spectra. The calculated results of IR spectra are in good agreement with the available experimental EELS data.

Acknowledgements

We thank Ministry of Science and Technology (MOST 101-2113-M-011-004-MY3), Taiwan for supporting this research financially. We are also grateful to National Center of High-Performance Computing, Institute of Nuclear Energy Research, and Atomic Energy Council in Taiwan for their support.

References

  1. K. Kita, K. Kyuno and A. Toriumi, Appl. Phys. Lett., 2004, 85, 52–54 CrossRef CAS PubMed.
  2. H. Kim, P. C. McIntyre, C. O. Chui, K. C. Saraswat and M. H. Cho, Appl. Phys. Lett., 2004, 85, 2902–2904 CrossRef CAS PubMed.
  3. A. Dimoulas, G. Mavrou, G. Vellianitis, E. Evangelou, N. Boukos, M. Houssa and M. Caymax, Appl. Phys. Lett., 2005, 86, 032908 CrossRef PubMed.
  4. P. W. Loscutoff and S. F. Bent, Annu. Rev. Phys. Chem., 2006, 57, 467–495 CrossRef CAS PubMed.
  5. D. Misra, R. Garg, P. Srinivasan, N. Rahim and N. A. Chowdhury, Mater. Sci. Semicond. Process., 2006, 9, 741–748 CrossRef CAS PubMed.
  6. K. Martens, B. De Jaeger, R. Bonzom, J. Van Steenbergen, M. Meuris, G. Groeseneken and H. Maes, IEEE Electron Device Lett., 2006, 27, 405–408 CrossRef CAS.
  7. N. Wu, Q. C. Zhang, C. X. Zhu, D. S. H. Chan, M. F. Li, N. Balasubramanian, A. Chin and D. L. Kwong, Appl. Phys. Lett., 2004, 85, 4127–4129 CrossRef CAS PubMed.
  8. G. W. Anderson, M. C. Hanf, P. R. Norton, Z. H. Lu and M. J. Graham, Appl. Phys. Lett., 1995, 66, 1123–1125 CrossRef CAS PubMed.
  9. L. M. Nelen, K. Fuller and C. M. Greenlief, Appl. Surf. Sci., 1999, 150, 65–72 CrossRef CAS.
  10. M. Houssa, G. Pourtois, B. Kaczer, B. De Jaeger, F. E. Leys, D. Nelis, K. Paredis, A. Vantomme, M. Caymax, M. Meuris and M. M. Heyns, Microelectron. Eng., 2007, 84, 2267–2273 CrossRef CAS PubMed.
  11. D. R. Fitzgerald and D. J. Doren, J. Am. Chem. Soc., 2000, 122, 12334–12339 CrossRef CAS.
  12. P. Kruger and J. Pollmann, Phys. Rev. Lett., 1995, 74, 1155–1158 CrossRef.
  13. T. Shirasawa, S. Mizuno and H. Tochihara, Surf. Sci., 2006, 600, 815–819 CrossRef CAS PubMed.
  14. C. Yang and H. C. Kang, J. Chem. Phys., 1999, 110, 11029–11037 CrossRef CAS PubMed.
  15. J. A. Kubby and J. J. Boland, Surf. Sci. Rep., 1996, 26, 61–204 CrossRef CAS.
  16. T. Weser, A. Bogen, B. Konrad, R. D. Schnell, C. A. Schug and W. Steinmann, Phys. Rev. B: Condens. Matter Mater. Phys., 1987, 35, 8184–8188 CrossRef CAS.
  17. M. Gothelid, G. LeLay, C. Wigren, M. Bjorkqvist, M. Rad and U. O. Karlsson, Appl. Surf. Sci., 1997, 115, 87–95 CrossRef CAS.
  18. M. Cakmak and G. P. Srivastava, Phys. Rev. B: Condens. Matter Mater. Phys., 1999, 60, 5497–5505 CrossRef CAS.
  19. C. Fleischmann, S. Sioncke, S. Couet, K. Schouteden, B. Beckhoff, M. Muller, P. Honicke, M. Kolbe, C. Van Haesendonck, M. Meuris, K. Temst and A. Vantomme, J. Electrochem. Soc., 2011, 158, H589–H594 CrossRef CAS PubMed.
  20. C. Fleischmann, M. Houssa, M. Muller, B. Beckhoff, H. G. Boyen, M. Meuris, K. Temst and A. Vantomme, J. Phys. Chem. C, 2013, 117, 7451–7458 CAS.
  21. H. J. Kuhr and W. Ranke, Surf. Sci., 1987, 189–190, 420–425 CrossRef CAS.
  22. D. Lee, K. Kubo, T. Kanashima and M. Okuyama, Jpn. J. Appl. Phys., 2012, 51 Search PubMed.
  23. T.-F. Teng, W.-L. Lee, Y.-F. Chang, J.-C. Jiang, J.-H. Wang and W.-H. Hung, J. Phys. Chem. C, 2009, 114, 1019–1027 Search PubMed.
  24. S. Sioncke, J. Ceuppens, D. Lin, L. Nyns, A. Delabie, H. Struyf, S. De Gendt, M. Muller, B. Beckhoff and M. Caymax, Microelectron. Eng., 2011, 88, 1553–1556 CrossRef CAS PubMed.
  25. J. S. Kachian and S. F. Bent, J. Am. Chem. Soc., 2009, 131, 7005–7015 CrossRef CAS PubMed.
  26. G. Kresse and J. Furthmuller, Comput. Mater. Sci., 1996, 6, 15–50 CrossRef CAS.
  27. G. Kresse and J. Hafner, Phys. Rev. B: Condens. Matter Mater. Phys., 1993, 48, 13115–13118 CrossRef CAS.
  28. G. Kresse and J. Furthmüller, Phys. Rev. B: Condens. Matter Mater. Phys., 1996, 54, 11169–11186 CrossRef CAS.
  29. D. Vanderbilt, Phys. Rev. B: Condens. Matter Mater. Phys., 1990, 41, 7892–7895 CrossRef.
  30. G. Henkelman, B. P. Uberuaga and H. Jonsson, J. Chem. Phys., 2000, 113, 9901–9904 CrossRef CAS PubMed.
  31. G. Mills, H. Jonsson and G. K. Schenter, Surf. Sci., 1995, 324, 305–337 CrossRef CAS.
  32. T. F. Teng, C. Y. Chou, W. H. Hung and J. C. Jiang, J. Phys. Chem. C, 2011, 115, 19203–19209 CAS.
  33. Z. G. Wang, X. T. Zu, J. L. Nie and H. Y. Xiao, Chem. Phys., 2006, 325, 525–530 CrossRef CAS PubMed.
  34. P. Krüger and J. Pollmann, Appl. Phys. A, 1994, 59, 487–502 CrossRef.
  35. R. Rossmann, H. L. Meyerheim, V. Jahns, J. Wever, W. Moritz, D. Wolf, D. Dornisch and H. Schulz, Surf. Sci., 1992, 199–209 CrossRef CAS.
  36. S. Hong, Y. E. Cho, J. Y. Maeng and S. Kim, J. Phys. Chem. B, 2004, 108, 15229–15232 CrossRef CAS.
  37. K. T. Leung, L. J. Terminello, Z. Hussain, X. S. Zhang, T. Hayashi and D. A. Shirley, Phys. Rev. B: Condens. Matter Mater. Phys., 1988, 38, 8241–8248 CrossRef CAS.

Footnote

Electronic supplementary information (ESI) available: ESI Table S1 lists the lattice parameters of Ge obtained with various pseudopotentials. Table S2 displays dimer bond length (dD) and dimer tilting angle (θ) for a clean Ge(100) surface, compared with results from other calculations and experimental data. Fig. S1 illustrates the partial DOS of H2S adsorbed on the Ge(100) surface. Fig. S2 shows a top view of all transition structures on the reaction paths of adsorbed H2S. See DOI: 10.1039/c4ra08887e

This journal is © The Royal Society of Chemistry 2015
Click here to see how this site uses Cookies. View our privacy policy here.