Next Article in Journal
High Ampacity Carbon Nanotube Materials
Previous Article in Journal
Transformation of Combustion Nanocatalysts inside Solid Rocket Motor under Various Pressures
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Prevention of Hydrogen Damage Using MoS2 Coating on Iron Surface

1
Beijing Advanced Innovation Center for Materials Genome Engineering, Corrosion and Protection Center, University of Science and Technology Beijing, Beijing 100083, China
2
Institute of Condensed Matter Physics, Linyi University, Linyi 276000, China
*
Authors to whom correspondence should be addressed.
Nanomaterials 2019, 9(3), 382; https://doi.org/10.3390/nano9030382
Submission received: 20 February 2019 / Revised: 28 February 2019 / Accepted: 1 March 2019 / Published: 6 March 2019

Abstract

:
The prevention of hydrogen penetration into steels can effectively protect steels from hydrogen damage. In this study, we investigated the effect of a monolayer MoS2 coating on hydrogen prevention using first-principles calculations. We found that monolayer MoS2 can effectively inhibit the dissociative adsorption of hydrogen molecules on an Fe(111) surface by forming a S–H bond. MoS2 coating acts as an energy barrier, interrupting hydrogen penetration. Furthermore, compared with the H-adsorbed Fe(111) film, the work function of the MoS2-coated film significantly increases under both equilibrium and strained conditions, indicating that the strained Fe(111) film with the MoS2 coating also becomes more corrosion resistant. The results reveal that MoS2 film is an effective coating to prevent hydrogen damage in steels.

Graphical Abstract

1. Introduction

Hydrogen damage is a severe problem since hydrogen degrades the mechanical properties of steels [1]. For example, hydrogen can induce low yield and fracture stresses in steels. Hydrogen atoms in structural materials usually come from the reduction of hydrogen ions or dissociative adsorptions of some H-containing gases, such as H2, H2O, and H2S. The prevention of hydrogen penetration into steels can effectively protect steels from hydrogen damage, which is typically achieved by applying a thin protective coating of TiN/TiC [2], SiC [3], an aluminum or a chromium rich layer [4,5,6], some alloy coatings [7], and conductive polymers [8] all of which show resistance to hydrogen or hydrogen isotopic permeability. It has been reported that AlN coatings [9] and Er2O3 coatings on 316L stainless steel [10] and MoS2/Ni80Cr20 films on pure iron [11] can act as protective barriers of metals against hydrogen or hydrogen isotope permeation. The graphene coating was found to decrease the hydrogen embrittlement susceptibility of the metal substrate, as the hydrogen content in graphene-coated copper was greatly reduced after hydrogen charging [12,13]. It is thought that less hydrogen is introduced to the bulk material by using these surface coatings.
Among various protection coatings, MoS2 possesses excellent physical and chemical properties that are suitable for preventing hydrogen permeation into metals. MoS2 not only has a 1185 °C melting point but is also chemically stable at an ambient atmosphere up to 315 °C. A good mechanical strength has also been reported for monolayer MoS2, which is a flexible and strong material with a high Young’s modulus comparable to steels [14]. Furthermore, monolayer MoS2 is a semiconductor with a direct band gap of ∼1.8 eV [15], and the low electrical conductivity and nearly insulating channel of MoS2 are also advantageous for hydrogen permeation barriers. For monolayer MoS2, the Mo atoms and S atoms combine with each other by covalent bonds. When coated on metals, the MoS2 layer usually exhibits a good adhesive performance, because S atoms can firmly bind to a metal surface [16]. The stable metal–S interface leads to a high diffusion barrier for hydrogen atoms to overcome in MoS2 coatings than that in metals. It is thought to protect the underlying metal substrate from corrosion and oxidation. Recently, MoS2 coatings of a few micrometers thick on pure iron substrates have been fabricated by magnetron sputtering [11]. However, the current understanding of MoS2 on steels as a protective barrier against hydrogen damage is still limited, which has motivated the present study.
More specifically, steels usually experience a moderate strain in the service environment due to external mechanical loads and residual stresses, which can affect the mechanical and electronic properties of steels [17,18]. Experimental studies have suggested that stress corrosion cracking and hydrogen embrittlement are dominating damages for steels. Thus, strain and hydrogen are two important factors to alter the physical properties of steels. The protective properties of MoS2 coating on iron films have also been investigated with applied stress and hydrogen.
Here we explore monolayer MoS2 as a promising coating for the protection of steels against hydrogen damage using first-principles calculations. To simulate the coating effects of MoS2 on steels, we chose pure iron film as a substrate, instead of steels with various additional elements, to provide a basic understanding. We found that the MoS2 monolayer can stably bind to the Fe(111) surface and effectively inhibit the dissociative adsorption and permeation of hydrogen. MoS2 interrupts hydrogen penetration by the formation of S–H bonds. In addition, compared with the H-adsorbed Fe(111) film, the work function of the MoS2-coated film significantly increases under both equilibrium and strained conditions. The present results suggest the feasibility of MoS2 coating as a protective barrier against hydrogen damage.

2. Calculation Methods

All calculations reported in this work were performed in the framework of the spin-polarized density functional theory (DFT), as implemented in the Vienna ab initio simulation package (VASP, 5.4.1, Universität Wien, Wien, Austria) [19]. The electron–ion interaction was described using the projector augmented wave (PAW) method [20]. The exchange correlation between electrons was treated with generalized gradient approximation (GGA) in the Perdew–Burke–Ernzerhof (PBE) form [21]. Using a 15 × 15 × 15 k-point mesh for the primitive cell, we obtained for body-centered cubic (bcc) Fe a lattice constant of 2.835 Å and a local magnetic moment of 2.20 μB per Fe atom, which agree very well with the experimental values of 2.866 Å [22] and the theoretical results performed with different k-point meshes [23,24,25,26].
For the study of H adsorption on the iron surface, we used a 4 × 4 surface cell of Fe(111) surface with six layers of Fe atoms, as shown in Figure 1a,b. A plane-wave energy cutoff of 400 eV with a k-point sampling of 3 × 3 × 1 in the Brillouin zone (BZ) was employed for the 4 × 4 surface cell. A vacuum region of 15 Å was introduced to eliminate the electronic interactions between the periodic images. The convergence criterion for the energy was set to 10−5 eV. For structural optimization, the bottom three Fe layers were fixed at their bulk positions and other atoms were fully relaxed until the force on each atom was lower than 0.01 eV/Å. In the case of asymmetric models, owing to the different electronegativity of two neighboring surfaces, a dipolar interaction appears in the z direction and affects the work function. To eliminate such interactions between the periodic replicas, dipole corrections were employed in the z direction for asymmetric slabs by using the method proposed by Neugebauer and Scheffler [27].

3. Results

When the H atom approaches the Fe(111) surface, we found that H chemically adsorbs on Fe atoms. Figure 1a,b displays the geometric structure of the film with one H atom adsorption on the surface. The H atom prefers to adsorb at the bridge site of the two top layer Fe atoms, such as Fe1 and Fe2 presented in Figure 1b. Table 1 lists the adsorption energy, Eads, which is defined as
E ads = E ( H / film ) E ( film ) 1 2 E ( H 2 )
where E(H/film) and E(film) are the energy of the Fe(111) film (or MoS2/Fe film) with or without H adsorption. The energy of H is referenced to the H2 molecule and hence reflects a dissociative adsorption energy. The adsorption energy for H adsorbing at the bridge site was −0.55 eV, suggesting that the H2 molecule can easily dissociate on the Fe(111) surface. Previous theoretical calculations have predicted that H adsorption energy on Fe(110) and Fe(100) surfaces is −0.71 eV and −0.38 eV [23], respectively. The H atom can form a strong chemical bond with the Fe3 atom with a distance of 1.636 Å, which is in good agreement with the literature results that report the closest H–Fe distance as 1.66 Å for H on Fe(110) and 1.68 Å for H on Fe(100) [23].
The chemisorbed H atom binds to the Fe(111) surface with a strong orbital overlap, which significantly affects the density of states (DOS) of the nearby Fe atoms, as shown in Figure 2. The H atom initially locates at the bridge site of the Fe1 and Fe2 atoms, and then binds to the Fe3 atom after geometric optimization. From Figure 2a, we observe that the projected density of states (PDOS) on Fe1 near the Fermi level slightly decreases after H adsorption, while the closest Fe3 exhibits an evident PDOS reduction as shown in Figure 2b—in particular the down spin. The decrease of Fe PDOS reflects the reduction of the charge density of Fe atoms, as the H atom withdraws 0.355 e from the Fe(111) surface after adsorption, as listed in Table 1. To further illustrate the detailed nature of the charge transfer, we show in Figure 2c the charge difference between the H-adsorbed Fe(111) system and the sum of the isolated Fe(111) and the H atom, which is defined as
Δ ρ = ρ H / Fe ( ρ Fe + ρ H )
where ρ H / Fe , ρ Fe , and ρ H are the charge density of the H/Fe(111) film, the Fe(111) film, and the H atom at the same lattice constant with frozen atom positions. The yellow regions represent the charge accumulation and the blue regions represent the depletion of electrons. To have a quantitative picture, we plot in Figure 2d the plane-averaged electron density difference along the perpendicular direction (z) to the Fe(111) surface. As seen in Figure 2c,d, the Fe(111) film acts as a donor and the H acts as an acceptor with electrons transferring from the Fe to the H due to the difference in electronegativity.
The above results show that hydrogen molecules can easily dissociate on the clean Fe surface and permeate into the bulk, as reported in literature [23]. The accumulation of H in bulk can generate hydrogen bubbles which are harmful for steels. In order to inhibit hydrogen permeation, a monolayer MoS2 is coated on the Fe(111) surface. For the model, a 3 3 × 3 3 MoS2 supercell was constructed in a hexagonal geometry on six layers of 4 × 4 Fe(111) film, as shown in Figure 1c,d. We have considered the MoS2/Fe(111) slab as a model, because the hexagonal unit cell on the Fe(111) surface structurally matches the surface cell of MoS2 and a better lattice match is achieved between both materials. The size of the unit cell of the 3 × 3 MoS2 slab is 5.473 × 5.473 Å with an angle of 120°, which is a good fit with the Fe(111) surface with the dimensions of 4.009 × 4.009 Å and an angle of 120°. In our setup, the MoS2 was subjected to a small strain (≈2.4%) to make it commensurable with Fe(111), and the effect of the lattice mismatch on the electronic structure of the MoS2 was negligible.
After optimizing the structures from four initial configurations in an interface with the monolayer MoS2, i.e., the top, bridge, fcc hollow, and hcp hollow sites formed by the three Fe atoms in the top layer, we obtained the most stable configurations of the MoS2/Fe(111) interfaces, as shown in Figure 1c,d. The S atom in the dotted red circle was located at the bridge site of two Fe atoms. In terms of the binding energy per interfacial sulfur atom, calculated as
E b = ( E MoS 2 / Fe E Fe E MoS 2 ) / N s
where NS is the number of interfacial sulfur atoms and NS = 27 for the calculated model, the Fe(111) surface had a medium adhesion with MoS2 with an Eb = −0.41 eV, which is larger than the weak interaction of MoS2–Au but smaller than the strong interaction of MoS2–Sc or MoS2–Ti [28]. The planar plane of the Fe(111) surface was distorted since the top layer Fe atoms were stretched by the MoS2. The average distance between the interfacial sulfur atoms with the top layer Fe atoms was 1.901 Å. The short interfacial distance also suggests that MoS2 forms a stable coating on an iron surface, which is beneficial for preventing hydrogen from transiting to Fe.
It is interesting to study the potential energy of the H atom adsorbing on the MoS2/Fe surface, and then moving through the interface region from the MoS2 part to the Fe substrate. To find the most energetically stable H adsorption site on the MoS2/Fe surface, we examined five possible initial positions for H on the clean monolayer MoS2, i.e., the top and bridge sites of Mo atoms and the top, bridge, and hollow sites of S atoms. In terms of the adsorption energy, we obtained the most energetically stable configuration, that is, H initially locates at the Mo bridge site and then binds to the surface S atom after structural relaxation. In this configuration, the atomic H chemically binds to the top surface of S with an Eads of 1.64 eV and a S–H distance of 1.417 Å (Table 1), in agreement with previous reports [29]. Figure 1c,d depicts the side and top views of the most stable H adsorption geometries on the MoS2/Fe surface, a similar adsorption configuration is obtained with an Eads of 1.26 eV. The positive Eads suggests that the H2 molecule cannot spontaneously dissociate on the MoS2/Fe film, which effectively suppresses the H dissociative adsorption. Experimental measurements have also demonstrated that the planar surface monolayer MoS2 is chemically rather inert while the edge sites of clusters and the defected layers are chemically reactive [30,31], which is consistent with the present results. Furthermore, as an atomic H migrates into the interface, it has been reported that an energy barrier of about 0.57 eV is required to pass through the center of the hexagonal structure in MoS2 [29], because there is a repulsive force induced by a strong electron cloud of MoS2. This energy barrier is up to 6.56 eV for the H2 molecule passing through the hole of the monolayer MoS2 [29], indicating that H is difficult to diffuse into the other side of MoS2. Even if H successfully moves to the interface at a high temperature or under a large tensile strain, the energy for binding to the Fe-contacted S atom is 0.15 eV higher than that of the up-surface S atom. All the results demonstrate that a MoS2 coating acts as an energy barrier which interrupts hydrogen penetration by the formation of S–H bonds.
To gain insight into the H prevention effect of the MoS2 coating, Figure 3a,b depicts the PDOS of the clean MoS2 film and the H-adsorbed MoS2/Fe(111) film. As seen from Figure 3a, a perfect monolayer MoS2 is a semiconductor with a band gap of 1.75 eV, in accordance with literature results [32]. There is no net magnetic moment for S and Mo atoms. When it is fabricated on Fe(111) film, the gap of MoS2 significantly shifts down in energy, as displayed in Figure 3b, i.e., the upward shift of the Fermi level, indicating that the MoS2 is n-doped on the Fe(111) surface. Furthermore, the electronic states of S and Mo are broadened within the band gap (−1.6~0.1 eV), due to the strong orbital interfacial hybridization between the Fe and S atoms. The H/MoS2/Fe(111) film still presents a ferromagnetic property, with split up and down orbitals for the PDOS in Figure 3b. This magnetic behavior mainly stems from the single d electron provided by the Fe atom. There is also a weak magnetic moment on the MoS2 layer induced by the Fe contact, in which Mo reaches 0.05 μB and the interface S atoms are less than 0.02 μB, in agreement with previous calculations for MoS2 on a single layer of Fe [33]. Moreover, for the H-adsorbed MoS2/Fe(111) film, the total DOS at the Fermi level, EF, is mainly contributed from the down spin of the Fe 3d states.
Figure 3c,d plots the interfacial charge transfer of the H-adsorbed MoS2/Fe(111) film. The differential charge density at the interface is defined as
Δ ρ = ρ H / MoS 2 / Fe ( ρ MoS 2 + ρ Fe + ρ H )
where ρ H / MoS 2 / Fe , ρ MoS 2 , ρ Fe , and ρ H are the charge densities of the H/MoS2/Fe, MoS2, and Fe(111) films and the H atom. Firstly, at the MoS2–Fe interface, there is a large amount of charge transfer from the top two layers of Fe to the interfacial S atoms in MoS2, a total of 4.91 e (calculated by Bader charge analysis). A strong binding between MoS2 and Fe with a distance of 1.901 Å allows a strong wave-function overlap between the Fe and the S states. Secondly, the adsorbed H donates 0.018 e to the neighboring S atom by forming a S–H bond, as listed in Table 1.
The above results demonstrate that the MoS2 coating on the Fe(111) film effectively prevents H adsorption on the iron surface or permeation into the bulk. The influence of the MoS2 coating on corrosion resistance is also reflected by work function, which is a sensitive parameter for the corrosive behavior of materials. Previous studies have suggested that materials with a lower work function possess a lower corrosion potential and consequently become easily corroded [34,35]. The work function (WF) is calculated as the difference between the vacuum level, Evacuum, and the Fermi energy, EF:
W F = E vacuum E F
Here, WF reflects the electronic energy level, so it is related to its electrostatic potential. The work function of a clean Fe(111) surface is 3.823 eV, which is lower than the closely packed crystallographic plane (low-index) Fe(110) surface [36]. After H adsorption, the WF is slightly enhanced to 3.842 eV, as listed in Table 1, because the H atom withdraws electrons from the Fe and a weak dipole pointing inward (from H to Fe) is formed. Figure 4a plots the WF of the H-adsorbed Fe(111) with and without the MoS2 coating. The WF of the H/MoS2/Fe(111) system significantly increases to 4.688 eV, which is lifted by 0.846 eV compared to that without the MoS2 coating, since the H-adsorbed MoS2 exhibits a high WF of 4.705 eV. It is indispensable to understand why the coating and H adsorption change the work function of these films.
There could be different factors, such as epitaxial strain, structural deformation, hydrogen, and the external environment, affecting the mechanical properties of steels. Experimental studies have suggested that stress corrosion cracking and hydrogen embrittlement are the most dominating damages for steels [17,18]. Next, we discuss the work function changes associated with the H adsorption in the strained system. A biaxial strain by rescaling the in-plane lattice constant was applied to the H/Fe(111) and the H/MoS2/Fe films. Without the MoS2 coating, the WF of the H/Fe(111) system increases/decreases by ~0.1 eV under compressive/tensile strain up to 6%, as plotted in Figure 4b. For instance, the WF is 3.939 eV under 6% compressive strain and 3.756 eV under 6% tensile strain. This property is consistent with the experimental and theoretical results that the WF of metals decreases with tensile strain [37,38]. In the case of the H/MoS2/Fe(111) system, the WF shows an opposite response to strain, i.e., it decreases to 4.631 eV under 6% compressive strain while it increases to 4.791 eV under 6% tensile strain. The changes in the WF of the MoS2-coated film can be attributed to the strained MoS2, which exhibited a WF decrease under compressive strain and a WF increase under tensile strain [39,40]. Under both compressive and tensile strains, a much higher work function is observed for the H/MoS2/Fe film than that of the system without the MoS2 coating, indicating that the strained Fe(111) film with the MoS2 coating becomes more corrosion resistant.
The effect of the coating and the strain on the work function depends on how they affect the Fermi energy. To clarify the origin of the WF changes in the strained systems, Figure 5 plots the orbital-resolved band structures for the H-adsorbed Fe(111) film without or with the MoS2 coating. Figure 5a–c shows the band structures for spin-up orbitals, corresponding with Figure 5d–f for the spin-down orbitals. Only five d orbitals are presented in Figure 5, because the energy bands near the Fermi level are mainly contributed from Fe 3d orbitals, as indicated by the PDOS in Figure 3b. Comparing the H/Fe(111) (Figure 5a,d) with the H/MoS2/Fe system (Figure 5b,e), one prominent property is that the bands near the Fermi level become less dispersive, especially for the down spin in Figure 5e. The flatter bands lead to a low Fermi velocity, indicating a quite low Fermi energy, and then give rise to an enhancement in the work function of H/MoS2/Fe. As a 6% tensile strain is applied to the H/MoS2/Fe(111) film, the flatter feature at the Fermi level is more pronounced in Figure 5f, suggesting that the WF of the H/MoS2/Fe system increases with tensile strain. The increased WF indicates that the MoS2-coated surface becomes more corrosion resistant.

4. Conclusions

In conclusion, we studied the MoS2 coating on an iron surface as a protective barrier against H damage. The monolayer MoS2 can be stably coated on the Fe(111) surface with a binding energy of −0.41 eV per surface S atom and an interfacial distance of 1.901 Å. Through the characterization of hydrogen-adsorbed MoS2/Fe(111), it was identified that MoS2 can effectively prevent hydrogen adsorption and penetration by the formation of a S–H bond. The hydrogen adsorption energy on the Fe(111) surface is enhanced from −0.55 eV to 1.26 eV with the MoS2 coating, suggesting that monolayer MoS2 can effectively inhibit the dissociative adsorption of hydrogen molecules. In addition, the work function of MoS2-coated Fe(111) films substantially increases by 0.846 eV, further indicating a more corrosion resistant property of the MoS2-coated Fe(111) films owing to their improved surface properties. The results demonstrate that the MoS2 coating is a proper barrier for H adsorption or permeation and can effectively avoid hydrogen damage. Based on the protective performance of monolayer MoS2, multiple layers of MoS2 or a thin film are expected to possess a better hydrogen prevention effect due to more barriers for hydrogen diffusion. Since it is still a challenge to produce a high quality monolayer MoS2 on a large scale on steels, we suggest that coating with multiple layers of MoS2 film might be more applicable.

Author Contributions

X.L. performed the calculation, data analysis, and wrote the manuscript; H.L., C.S., D.W., and Z.M. discussed the results and analyzed the data; and L.C. and L.Q. conceived the models and revised the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (Grant Nos. 51431004, 11634007, U1706221, and 51571030).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Hirth, J.P.; Medalist, R.F.M. Effects of hydrogen on the properties of iron and steel. Metall. Trans. A 1980, 11, 861–890. [Google Scholar] [CrossRef]
  2. Yao, Z.; Hao, J.; Zhou, C.; Shan, C.; Yu, J. The permeation of tritium through 316L stainless steel with multiple coatings. J. Nuclear Mater. 2000, 283-287, 1287–1291. [Google Scholar] [CrossRef]
  3. Causey, R.A.; Wampler, W.R. The use of silicon carbide as a tritium permeation barrier. J. Nuclear Mater. 1995, 220–222, 823–826. [Google Scholar] [CrossRef]
  4. Kalin, B.A.; Yakushin, V.L.; Fomina, E.P. Tritium barrier development for austenitic stainless steel by its aluminizing in a lithium melt. Fusion Eng. Des. 1998, 41, 119–127. [Google Scholar] [CrossRef]
  5. Park, I.J.; Jeong, K.H.; Jung, J.G.; Lee, C.S.; Lee, Y.K. The mechanism of enhanced resistance to the hydrogen delayed fracture in Al-added Fe-18Mn-0.6C twinning-induced plasticity steels. Int. J. Hydrogen Energy 2012, 37, 9925–9932. [Google Scholar] [CrossRef]
  6. Zamanzade, M.; Vehoff, H.; Barnoush, A. Cr effect on hydrogen embrittlement of Fe3Al-based iron aluminide intermetallics: Surface or bulk effect. Acta Mater. 2014, 69, 210–233. [Google Scholar] [CrossRef]
  7. Johnson, D.F.; Carter, E.A. First-principles assessment of hydrogen absorption into FeAl and Fe3Si: Towards prevention of steel embrittlement. Acta Mater. 2010, 58, 638–648. [Google Scholar] [CrossRef]
  8. Zarras, P.; Anderson, N.; Webber, C.; Irvin, D.J.; Irvin, J.A.; Guenthner, A.; Stenger-Smith, J.D. Progress in using conductive polymers as corrosion-inhibiting coatings. Radiat. Phys. Chem. 2003, 68, 387–394. [Google Scholar] [CrossRef]
  9. Wang, J.; Li, Q.; Xiang, Q.Y.; Cao, J.L. Performances of AlN coatings as hydrogen isotopes permeation barriers. Fusion Eng. Des. 2016, 102, 94–98. [Google Scholar] [CrossRef]
  10. Li, Q.; Wang, J.; Xiang, Q.Y.; Tang, T.; Rao, Y.C.; Cao, J.L. Thickness impacts on permeation reduction factor of Er2O3 hydrogen isotopes permeation barriers prepared by magnetron sputtering. Int. J. Hydrogen Energy 2016, 41, 3299–3306. [Google Scholar] [CrossRef]
  11. Zhou, P.; Li, W.; Jin, X. Electrochemical hydrogen permeation properties of MoS2 and Ni80Cr20 films prepared by magnetron sputtering on pure iron. J. Electrochem. Soc. 2018, 165, E256–E261. [Google Scholar] [CrossRef]
  12. Nam, T.H.; Lee, J.H.; Choi, S.R.; Yoo, J.B.; Kim, J.G. Graphene coating as a protective barrier against hydrogen embrittlement. Int. J. Hydrogen Energy 2014, 39, 11810–11817. [Google Scholar] [CrossRef]
  13. Chen, L.; Shi, C.; Li, X.; Mi, Z.; Jiang, C.; Qiao, L.; Volinsky, A.A. Passivation of hydrogen damage using graphene coating on α-Fe2O3 films. Carbon 2018, 130, 19–24. [Google Scholar] [CrossRef]
  14. Bertolazzi, S.; Brivio, J.; Kis, A. Stretching and breaking of ultrathin MoS2. ACS Nano 2011, 5, 9703–9709. [Google Scholar] [CrossRef] [PubMed]
  15. Mak, K.F.; Lee, C.; Hone, J.; Shan, J.; Heinz, T.F. Atomically thin MoS2: A new direct-gap semiconductor. Phys. Rev. Lett. 2010, 105, 136805. [Google Scholar] [CrossRef] [PubMed]
  16. Das, S.; Chen, H.Y.; Penumatcha, A.V.; Appenzeller, J. High performance multilayer MoS2 transistors with scandium contacts. Nano Lett. 2013, 13, 100–105. [Google Scholar] [CrossRef] [PubMed]
  17. Qiao, L.; Chu, W.; Miao, H.; Xiao, J.; Guo, P. Hydrogen-facilitated corrosion and stress corrosion cracking of austenitic stainless steel of type 310. Metall. Trans. A 1993, 24, 959–962. [Google Scholar]
  18. Alyousif, O.M.; Nishimura, R. Stress corrosion cracking and hydrogen embrittlement of sensitized austenitic stainless steels in boiling saturated magnesium chloride solutions. Corros. Sci. 2008, 50, 2353–2359. [Google Scholar] [CrossRef]
  19. Kresse, G.; Furthmüller, J. Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set. Phys. Rev. B 1996, 54, 11169–11186. [Google Scholar] [CrossRef]
  20. Kresse, G.; Joubert, D. From ultrasoft pseudopotentials to the projector augmented-wave method. Phys. Rev. B 1999, 59, 1758–1775. [Google Scholar] [CrossRef]
  21. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 1996, 77, 3865–3868. [Google Scholar] [CrossRef] [PubMed]
  22. Kittel, C. Introduction to Solid State Physics, 7th ed.; Wiley: New York, NY, USA, 1996. [Google Scholar]
  23. Jiang, D.E.; Carter, E.A. Diffusion of interstitial hydrogen into and through bcc Fe from first principles. Phys. Rev. B 2004, 70, 064102. [Google Scholar] [CrossRef]
  24. Counts, W.A.; Wolverton, C.; Gibala, R. First-principles energetics of hydrogen traps in α-Fe: Point defects. Acta Mater. 2010, 58, 4730–4741. [Google Scholar] [CrossRef]
  25. Momida, H.; Asari, Y.; Nakamura, Y.; Tateyama, Y.; Ohno, T. Hydrogen-enhanced vacancy embrittlement of grain boundaries in iron. Phys. Rev. B 2013, 88, 144107. [Google Scholar] [CrossRef]
  26. Tateyama, Y.; Ohno, T. Stability and clusterization of hydrogen-vacancy complexes in α-Fe: An ab initio study. Phys. Rev. B 2003, 67, 174105. [Google Scholar] [CrossRef]
  27. Neugebauer, J.; Scheffler, M. Adsorbate-substrate and adsorbate-adsorbate interactions of Na and K adlayers on Al(111). Phys. Rev. B 1992, 46, 16067–16080. [Google Scholar] [CrossRef] [Green Version]
  28. Zhong, H.; Quhe, R.; Wang, Y.; Ni, Z.; Ye, M.; Song, Z. Interfacial properties of monolayer and bilayer MoS2 contacts with metals: Beyond the energy band calculations. Sci. Rep. 2016, 6, 21786. [Google Scholar] [CrossRef] [PubMed]
  29. Koh, E.W.K.; Chiu, C.H.; Lim, Y.K.; Zhang, Y.W.; Pan, H. Hydrogen adsorption on and diffusion through MoS2 monolayer: First-principles study. Int. J. Hydrogen Energy 2012, 37, 14323–14328. [Google Scholar]
  30. Jaramillo, T.F.; Jørgensen, K.P.; Bonde, J.; Nielsen, J.H.; Horch, S.; Chorkendorff, I. Identification of active edge sites for electrochemical H2 evolution from MoS2 nanocatalysts. Science 2007, 317, 100–102. [Google Scholar] [CrossRef] [PubMed]
  31. Ye, G.; Gong, Y.; Lin, J.; Li, B.; He, Y.; Pantelides, S.T.; Zhou, W.; Vajtai, R.; Ajayan, P.M. Defects engineered monolayer MoS2 for improved hydrogen evolution reaction. Nano Lett. 2016, 16, 1097–1103. [Google Scholar] [CrossRef] [PubMed]
  32. Lebègue, S.; Eriksson, O. Electronic structure of two-dimensional crystals from ab initio theory. Phys. Rev. B 2009, 79, 115409. [Google Scholar] [CrossRef]
  33. Song, C.; Gong, S.; Zhang, Z.; Mao, H.; Zhao, Q.; Wang, J.; Xing, H. Tuning magnetic anistropy by charge injection and strain in Fe/MoS2 bilayer heterostructures. J. Phys. D Appl. Phys. 2015, 48, 485001. [Google Scholar] [CrossRef]
  34. Li, W.; Li, D.Y. Variations of work function and corrosion behaviors of deformed copper surfaces. Appl. Surf. Sci. 2005, 240, 388–395. [Google Scholar] [CrossRef]
  35. Lu, H.; Hua, G.; Li, D. Dependence of the mechanical behavior of alloys on their electron work function—An alternative parameter for materials design. Appl. Phys. Lett. 2013, 103, 261902. [Google Scholar] [CrossRef]
  36. Guo, L.; Hua, G.; Yang, B.; Lu, H.; Qiao, L.; Yan, X.; Li, D. Electron work functions of ferrite and austenite phases in a duplex stainless steel and their adhesive forces with AFM silicon probe. Sci. Rep. 2016, 6, 20660. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Li, W.; Cai, M.; Wang, Y.; Yu, S. Influences of tensile strain and strain rate on the electron work function of metals and alloys. Scr. Mater. 2006, 54, 921–924. [Google Scholar] [CrossRef]
  38. Gong, H.R.; Nishi, Y.; Cho, K. Effects of strain and interface on work function of a Nb-W metal gate system. Appl. Phys. Lett. 2007, 91, 242105. [Google Scholar] [CrossRef]
  39. Lanzillo, N.A.; Simbeck, A.J.; Nayak, S.K. Strain engineering the work function in monolayer metal dichalcogenides. J. Phys. Condens. Matter 2015, 27, 175501. [Google Scholar] [CrossRef] [PubMed]
  40. Shakya, J.; Kumar, S.; Kanjilal, D.; Mohanty, T. Work function modulation of molybdenum disulfide nanosheets by introducing systematic lattice strain. Sci. Rep. 2017, 7, 9576. [Google Scholar] [CrossRef] [PubMed]
Figure 1. (a) Side and (b) top views of H-adsorbed Fe(111) film with six layers. The brown, green, and silver balls represent Fe atoms in the A, B, and C layer, respectively. The possible adsorption sites of H are marked in (b), on the top site of surface Fe (T), the bridge site of two surface Fe atoms (B), the hcp (hexagonal close packed) hollow site (H), and the fcc (face-centered cubic) hollow site (F). The numbers 1, 2, and 3 in (b) present the Fe1, Fe2, and Fe3 atoms that surround the H atom. (c) Side and (d) top views of H-adsorbed MoS2/Fe(111) film. For the MoS2 and Fe(111) interfaces, one S atom (noted in the red dotted circle in (d)) located at the bridge site of the surface Fe atoms is the most energetically stable configuration.
Figure 1. (a) Side and (b) top views of H-adsorbed Fe(111) film with six layers. The brown, green, and silver balls represent Fe atoms in the A, B, and C layer, respectively. The possible adsorption sites of H are marked in (b), on the top site of surface Fe (T), the bridge site of two surface Fe atoms (B), the hcp (hexagonal close packed) hollow site (H), and the fcc (face-centered cubic) hollow site (F). The numbers 1, 2, and 3 in (b) present the Fe1, Fe2, and Fe3 atoms that surround the H atom. (c) Side and (d) top views of H-adsorbed MoS2/Fe(111) film. For the MoS2 and Fe(111) interfaces, one S atom (noted in the red dotted circle in (d)) located at the bridge site of the surface Fe atoms is the most energetically stable configuration.
Nanomaterials 09 00382 g001
Figure 2. Projected density of states (PDOS) onto (a) Fe1 and (b) Fe3 atoms in the clean Fe(111) film and the H-adsorbed Fe(111) film, respectively. (c) Charge transfer density between a H atom and Fe(111). The yellow and blue colors represent the electron accumulation and depletion, respectively. The charge density isosurface was set to 0.003 e Å−3. (d) Interfacial charge transfer between a H atom and the Fe(111) film as a function of the z coordinate perpendicular to the surface.
Figure 2. Projected density of states (PDOS) onto (a) Fe1 and (b) Fe3 atoms in the clean Fe(111) film and the H-adsorbed Fe(111) film, respectively. (c) Charge transfer density between a H atom and Fe(111). The yellow and blue colors represent the electron accumulation and depletion, respectively. The charge density isosurface was set to 0.003 e Å−3. (d) Interfacial charge transfer between a H atom and the Fe(111) film as a function of the z coordinate perpendicular to the surface.
Nanomaterials 09 00382 g002
Figure 3. (a) Projected density of states on S and Mo atoms in a clean monolayer MoS2. (b) Total density of states (DOS) and PDOS on Fe, MoS2, and H in the H/MoS2/Fe system. (c) Charge transfer density between the H, the MoS2, and the Fe(111) film. The yellow and blue colors represent electron accumulation and depletion, respectively. The charge density isosurface was set to 0.003 e Å−3. (d) Interfacial charge transfer between the H, the MoS2, and the Fe(111) film as a function of the z coordinate perpendicular to the surface.
Figure 3. (a) Projected density of states on S and Mo atoms in a clean monolayer MoS2. (b) Total density of states (DOS) and PDOS on Fe, MoS2, and H in the H/MoS2/Fe system. (c) Charge transfer density between the H, the MoS2, and the Fe(111) film. The yellow and blue colors represent electron accumulation and depletion, respectively. The charge density isosurface was set to 0.003 e Å−3. (d) Interfacial charge transfer between the H, the MoS2, and the Fe(111) film as a function of the z coordinate perpendicular to the surface.
Nanomaterials 09 00382 g003
Figure 4. (a) The work function of the Fe(111) films and the MoS2/Fe(111) films without and with H adsorption. (b) The work function of the H-adsorbed Fe(111) films and the H-adsorbed MoS2/Fe(111) films with a strain of up to ±6%.
Figure 4. (a) The work function of the Fe(111) films and the MoS2/Fe(111) films without and with H adsorption. (b) The work function of the H-adsorbed Fe(111) films and the H-adsorbed MoS2/Fe(111) films with a strain of up to ±6%.
Nanomaterials 09 00382 g004
Figure 5. Band structures along the M-Γ-K directions of (a,d) the H-adsorbed Fe(111), (b,e) the H-adsorbed MoS2/Fe film, and (c,f) the H-adsorbed MoS2/Fe film with 6% tensile strain for up and down spins. Blue, green, red, purple, and orange lines on the bands illustrate the contribution from dxy, dxz, dyz, d z 2 , and d x 2 y 2 states.
Figure 5. Band structures along the M-Γ-K directions of (a,d) the H-adsorbed Fe(111), (b,e) the H-adsorbed MoS2/Fe film, and (c,f) the H-adsorbed MoS2/Fe film with 6% tensile strain for up and down spins. Blue, green, red, purple, and orange lines on the bands illustrate the contribution from dxy, dxz, dyz, d z 2 , and d x 2 y 2 states.
Nanomaterials 09 00382 g005
Table 1. The adsorption energy (Eads), the closest H–Fe distance (dH-Fe) or H–S distance (dH-S), the amount of charge transfer of the adsorbed H atom (ΔQH), and the work function (WF) of H adsorptions on the Fe(111), MoS2, and MoS2/Fe(111) films. Eads is calculated by Equation (1).
Table 1. The adsorption energy (Eads), the closest H–Fe distance (dH-Fe) or H–S distance (dH-S), the amount of charge transfer of the adsorbed H atom (ΔQH), and the work function (WF) of H adsorptions on the Fe(111), MoS2, and MoS2/Fe(111) films. Eads is calculated by Equation (1).
Eads (eV)dH-Fe/dH-S (Å)ΔQH (e)WF (eV)
H on Fe(111)−0.551.6360.3553.842
H on MoS21.641.417−0.0424.705
H on MoS2/Fe(111)1.261.425−0.0184.688

Share and Cite

MDPI and ACS Style

Li, X.; Chen, L.; Liu, H.; Shi, C.; Wang, D.; Mi, Z.; Qiao, L. Prevention of Hydrogen Damage Using MoS2 Coating on Iron Surface. Nanomaterials 2019, 9, 382. https://doi.org/10.3390/nano9030382

AMA Style

Li X, Chen L, Liu H, Shi C, Wang D, Mi Z, Qiao L. Prevention of Hydrogen Damage Using MoS2 Coating on Iron Surface. Nanomaterials. 2019; 9(3):382. https://doi.org/10.3390/nano9030382

Chicago/Turabian Style

Li, Xiaolong, Li Chen, Hongmei Liu, Changmin Shi, Dongchao Wang, Zhishan Mi, and Lijie Qiao. 2019. "Prevention of Hydrogen Damage Using MoS2 Coating on Iron Surface" Nanomaterials 9, no. 3: 382. https://doi.org/10.3390/nano9030382

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