Next Article in Journal
Density-Sensitive Implicit Functions Using Sub-Voxel Sampling in Additive Manufacturing
Previous Article in Journal
Dynamic Recrystallization Simulation for X12 Alloy Steel by CA Method Based on Modified L-J Dislocation Density Model
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Combined Density Functional Theory and Reduction Kinetics Investigation of Enhanced Adsorption of Hydrogen onto Fe2O3 by Surface Modification with Nickel

1
State Key Laboratory of Complex Nonferrous Metal Resources Clean Utilization, Kunming University of Science and Technology, Kunming 650093, China
2
Faculty of Metallurgy and Energy Engineering, Kunming University of Science and Technology, Kunming 650093, China
*
Author to whom correspondence should be addressed.
Metals 2019, 9(12), 1292; https://doi.org/10.3390/met9121292
Submission received: 10 November 2019 / Revised: 25 November 2019 / Accepted: 26 November 2019 / Published: 30 November 2019

Abstract

:
Based on the density functional theory, the geometric structure, adsorption energy and density of states of H2/α-Fe2O3 (001) system and (H2 + Ni)/α-Fe2O3 (001) system were determined. The results showed that the absolute value of adsorption energy between H2 molecule and α-Fe2O3 (001) surface and the bond length of H2 molecule were increased by the presence of Ni atom. The presence of Ni atom promotes the adsorption of H2 molecule on α-Fe2O3 (001) surface. Reduction behavior of iron and nickel oxides in H2 atmosphere was determined by thermogravimetric analysis. The samples included Fe2O3, Fe2O3-NiO and Fe2O3-Ni systems. The effect of Ni and NiO on the reduction behavior of Fe2O3 was investigated. The ease of reduction within the following three systems decreases sequentially: Fe2O3-Ni > Fe2O3-NiO > Fe2O3. The activation energy of Fe2O3-Ni, Fe2O3-NiO and Fe2O3 systems at two temperature stages (viz. < 400 °C and > 400 °C) were 172 kJ·mol−1, 197 kJ·mol−1 and 263 kJ·mol−1 respectively.

Graphical Abstract

1. Introduction

Nickel is a silver-white metal with good mechanical strength, plasticity and high chemical stability. Nickel is mainly used in stainless steel, electroplating and catalyst industries [1,2]. About 2/3 of the world’s nickel resources are used to produce stainless steel. About 70% of the production cost of austenitic stainless steel is the consumption of nickel [3,4,5]. Nickel catalysts are widely used in the hydrogenation of various unsaturated hydrocarbons due to their good catalytic activity, high mechanical strength and good thermal conductivity. They are also good catalysts for dehydrogenation, oxidative dehalogenation and desulfurization [6].
To provide a theoretical insight into the reaction mechanism of substances by investigating the electronic structure of multi-electron systems, density functional theory (DFT), which expresses the electronic energy as a functional of the electron density, was also included in this work. It has become an important means to study the properties and reaction mechanism of substances by simulation based on density functional theory and combining with macroscopic experimental phenomena [7,8,9]. In the present study, based on the first principle of density functional theory, the geometric structure, adsorption energy and density of states of H2/α-Fe2O3 (001) system and (H2 + Ni)/α-Fe2O3 (001) system were investigated.
Cores et al. [10] studied the reduction process of Fe3O4-NiO by pulverized coal and H2. It was found that there are two steps in the reduction reaction: The first step was the reduction of NiO and NiFe2O4, and the second step was the reduction of iron oxides. Zhou et al. [11] studied the reaction process of Fe2O3-NiO by methane. It was found that the first stage of the reduction process of Fe2O3 was Fe2O3 reduced to FeO, and the second stage was FeO reduced to Fe. The addition of NiO promotes the initial stage of reduction process of Fe2O3. Abdel-Halim et al. [12] studied the preparation of ferronickel alloy by reduction of Fe2O3/NiO compound with carbon. The results showed that the reaction was carried out step by step. At low temperatures, Boudouard reaction was the controlling step. Feng et al. [13] studied the effect of chloride on sulfide adsorption on cerussite surface through DFT calculations and XPS measurements. In the presence of chloride species, it had a significant effect on the surface structures and electronic properties of cerussite.
In this paper, based on the density functional theory, the effect of Ni atom on the adsorption behavior of H2 molecule on α-Fe2O3 (001) surface was investigated. The reduction of the Fe-Ni-O-H system was determined by a thermogravimetric analyzer. According to the weight loss curve of the sample, the kinetic mechanism was studied.

2. Computational and Experimental Methods

2.1. Computational Details

The structure of the hexagonal hematite (α-Fe2O3) is an R-3c space group with lattice constants of a = b = 5.04 Å, c = 13.75 Å, α = β = 90°, and γ = 120°. α-Fe2O3 (001) surface is one of the dominant crystal facets of nature α-Fe2O3. In this paper, the nine atomic layer slab models (P = 2 × 2) of the α-Fe2O3 (001) surface were studied [14]. A vacuum region of 15 Å was added. This region was optimized by relaxing the outmost six layers and fixing the bottom three layers [15,16,17]. The clean surface of α-Fe2O3 (001) is shown in Figure 1.
Calculations were processed using the Cambridge Sequential Total Energy Package (CASTEP) module. The Perdew–Burke–Ernzerhof (PBE) exchange correlation function form in the generalized gradient approximation (GGA) was selected, it is suitable for the study of iron-oxygen system [15,18]. Pseudopotential chooses the ultra-soft pseudopotentials in reciprocal space representation. The plane-wave has the cutoff energy of 350 eV in all calculations, and the value of k-points is 4 × 4 × 1 [15,16,19,20]. The values of convergence criterial for energy is 2.0 × 10−5 eV/atom, maximum force is 0.005 Ha/Å, maximum stress is 0.1 GPa, displacement in Self Consistent Field (SCF) is 0.002 Å. The spin-polarized runs through the all calculation. The optimized cell parameters (a = b = 5.01 Å, c = 13.91 Å) are in good agreement with the experimental values.
The optimized H2 molecule was placed vertically at the Fe-top adsorption site on the α-Fe2O3 (001) surface, and Ni atoms were placed at the O-top adsorption site on the α-Fe2O3 (001) surface as shown in Figure 2. The adsorption energy, Eads, was calculated using following Equation [21]:
Eads(H2/α-Fe2O3 (001)) = EH2 − Fe2O3 (001) − Eα − Fe2O3 (001) − EH2
Eads((H2 + Ni)/α − Fe2O3 (001)) = E(H2 + Ni)/α − Fe2O3 (001) − Eα − Fe2O3 (001) − EH2 − ENi
where EH2/α-Fe2O3 (001), E(H2+Ni)/α-Fe2O3 (001), Eα-Fe2O3 (001), EH2 and ENi represent the total energies of H2/α-Fe2O3 (001) system, (H2+Ni)/α-Fe2O3 (001) system, α-Fe2O3 (001) system, the adsorbate H2 molecule and the adsorbate Ni atom, respectively.

2.2. Experimental Details

The Fe2O3-Ni and Fe2O3-NiO samples were mixed homogeneously by a ball mill. The three samples were studied in the experiment: Fe2O3 (200 mg), Fe2O3 + Ni (200 mg + 20 mg), Fe2O3 + NiO (200 mg + 25 mg). The mass of Ni component in Fe2O3 + NiO and Fe2O3 + Ni samples was the same. A thermogravimetric analyzer (HENGJIU Inc., Beijing, China) was used to study the reduction experiment of Fe-Ni-O system in H2 atmosphere. The three samples were reduced by H2 (99.95%) with gas flow rate of 30 mL/min and a heating rate of 10 °C/min from 20 to 1000 °C. The phase composition of the samples was determined by X-ray diffraction (JEOL Inc., Tokyo, Japan). The diffraction spectra in the range 10° to 90° at a rate of 10°/min were measured.

2.3. Kinetic Approach

Study on kinetics of reduction process by thermal analysis [22,23,24], the kinetic equation under non-isothermal conditions is obtained.
d α d T = A β e x p ( E R T ) f ( α )
where α is the degree of conversion, β is the heating rate, E is the activation energy of the reaction, R is 8.314 J·mol−1·K−1, f(α) is the differential mechanism function.
The kinetic equation of reduction process is obtained by Coats–Redfern approximation equation.
ln [ G ( α ) T 2 ] = ln [ A R β E ( 1 2 R T E ) ] E R T
For most reactions, the activation energy E >> RT, and Equation (4) can be simplified:
ln [ G ( α ) T 2 ] = ln ( A R β E ) E R T
For a constant heating rate, the linear relationship between ln(G(α)/T2) and 1/T can be obtained, and E can be determined from the slope.

3. Results and Discussion

3.1. Computational Results of H2 Molecule on the α-Fe2O3 (001) Surface in the Presence of Ni Atom

The structure parameters and adsorption energy of H2/α-Fe2O3 (001) system and (H2 + Ni)/α-Fe2O3 (001) system are shown in Table 1. Negative value of Eads represents a stable adsorption configuration. The greater the absolute value of adsorption energy represents stronger interaction and the stronger interaction between adsorbate and substrate. As shown in Table 1, the absolute value of adsorption energy of (H2 + Ni)/α-Fe2O3 (001) system is 0.93 eV higher than that of H2/α-Fe2O3 (001) system. It is shown that the presence of Ni atom strengthens the interaction between the H2 molecule and the α-Fe2O3 (001) surface. When Ni atom was added to H2/α-Fe2O3 (001) system, the bond length of H2 molecule increased from 0.758 to 0.889 Å and a 63.36° tilt occurred between the H2 molecule and the α-Fe2O3 (001) surface. The variation of structure parameters of H2 molecule indicates that the existence of Ni atom can affect the structure of H2 molecule itself. When Ni atom was added to H2/α-Fe2O3 (001) system, which causes the inclination angle between H2 molecule and the α-Fe2O3 (001) surface, thus leading to the change of surface charge distribution of α-Fe2O3 (001) [21,25]. Compared with H2/α-Fe2O3 (001) system, the average bond length of Fe–O in (H2 + Ni)/α-Fe2O3 (001) system increased from 2.0253 to 2.0258 Å, indicating that α-Fe2O3 (001) in (H2 + Ni)/α-Fe2O3 (001) system has been activated, which will be more conducive to the adsorption of H2 molecules on α-Fe2O3 (001) surface [26].
Figure 3 shows the density of states (DOS) of α-Fe2O3 (001) surface and H2 molecules of H2/α-Fe2O3 (001) system in the presence or absence of Ni atom. The green dotted line in the two figures represent Fermi level. When the energy is greater than Fermi level, it is unoccupied state, and when energy is less than Fermi level, it is occupied state. As shown in Figure 3a, when the Ni atom is added to the H2/α-Fe2O3 (001) system, within the range of −2.0 to 1.50 eV, the electrons tend to move to higher energy levels, resulting in the increase in density of the electrons with higher energies. Meanwhile, within the range of −7.50 to 1.50 eV, the peaks also show a slight displacement to the more positive side. It is found that the presence of Ni atom in the H2/α-Fe2O3 (001) system leads to the transition to higher energy levels of electron of H2 molecule, and thus an obvious displacement of the peaks to more positive side is shown Figure 3b. Moreover, a new peak was found in the unoccupied states. These results indicate that the stability of the H2 molecule decreases and the tendency of cracking occurs with the addition of Ni atom [27].
Based on the above theoretical analysis, the existence of Ni atom has a great influence on the structure of H2 molecule and α-Fe2O3 (001) surface and promotes the adsorption of H2 molecule on α-Fe2O3 (001) surface. From the microscopic point of view, Ni components can promote the reduction of Fe2O3 in H2 atmosphere, and we will validate this conclusion through experiments.

3.2. Reduction Behavior in the Fe-Ni-O-H System

The reduction curves of Fe2O3, Fe2O3-Ni and Fe2O3-NiO exposed to H2 are shown in Figure 4. When the ordinate value of Figure 4 is 1.0, it means the reduction degree is 100%. Comparing the reduction degree curves of three samples, reduction reaction of Fe2O3-Ni was firstly completed, followed by Fe2O3-NiO, and Fe2O3 was finally completed. It can be considered that Ni components were determined as accelerator in Fe2O3 system, and final reaction temperature of Fe2O3 was reduced. There are two reasons for the promotion behavior of Ni components. Firstly, Ni has large specific surface area, many surface active sites and strong adsorption capacity for H2 [28]. When Ni is present within the Fe2O3 system, Ni adsorbs more hydrogen atoms and transfers them to Fe2O3 system, which promotes the reduction reaction of Fe2O3 system. Secondly, NiO reacts with Fe2O3 to produce NiFe2O4, which is easier to reduce. When NiO was added to Fe2O3 system, the morphological structure of samples changed significantly and the porosity of the compacts increased and promoted the reduction of Fe2O3 [12,29]. Compared with Ni, the promotion effect of NiO is not as obvious. The reaction rate curves of Fe2O3, Fe2O3-Ni and Fe2O3-NiO are shown in Figure 5. The temperature corresponding to the maximum reaction rate of the three samples ranged from low to high: Fe2O3-Ni < Fe2O3-NiO < Fe2O3. As shown in Figure 5, the reaction rate curves of three samples are mainly composed of two peaks, and the boundary point of the two peaks is about 400 °C. According to the result, we analyzed the reduction products of three samples at different temperatures by XRD and explored the reduction process of three samples in H2 atmosphere.
As shown in Figure 6, the reduction products phases of three samples at different temperatures were characterized by XRD. Figure 6a shows that the phases of reduction products of Fe2O3 at different temperatures. The phases present are Fe2O3 and Fe3O4 at 380 °C. At 550 °C, the phases present are Fe3O4 and Fe. At 590 °C, FeO diffraction peaks appear, and the phases present are Fe3O4, FeO and Fe. At 1000 °C, Fe2O3 and H2 react completely, and the phase present is Fe. From the results of XRD, the reduction of Fe2O3 by H2 is a multi-stage process [30,31]. When the temperature is less than 400 °C, Fe2O3 is reduced to Fe3O4. In the temperature range of 400–570 °C, the following reaction occurred: Fe3O4→Fe. When the temperature is higher than 570 °C, the following reactions occurred: Fe3O4→FeO→Fe. Figure 6b shows that the constituent phases of Fe2O3-NiO after reduction at different temperatures. Fe2O3, Fe3O4 and Ni are the phases present at 380 °C. At 570 °C, awaruite [FeNi3] diffraction peaks appear. At this temperature, the phases present are Fe3O4, Fe and FeNi3. When the temperature is increased to 590 °C, the phases present are Fe3O4, FeO, Fe and FeNi3. At 1000 °C, Fe2O3 and NiO are completely reduced and the phases present are Fe and FeNi3. From the results of XRD, the reduction process of Fe2O3-NiO is divided into two steps. The first step is mainly the reduction of NiO and the transformation of Fe2O3 to low-valent iron oxides. At the same time, a small amount of Fe-Ni alloy is formed. The second step is the reduction of low-valent iron oxides and the formation of Fe-Ni alloys [32]. Comparing Figure 6b,c, we can see that the reduction products of Fe2O3-Ni and Fe2O3-NiO at different temperatures have the same constituent phases, and the reduction process is basically the same. The first step is the reduction of Fe2O3 to Fe3O4, and the second step is the reduction of Fe3O4 and the formation of Fe-Ni alloys.

3.3. Reduction Kinetics of Fe-Ni-O-H System

Because the reaction rate curves of three samples are mainly composed of two peaks, and the boundary point of the two peaks is about 400 °C. The kinetic calculation was divided into two sections (above 400 °C and below 400 °C). Table 2 lists the fifteen commonly used mechanism functions in gas-solid reactions, by analyzing the curve of ln(G(α))/T2 versus 1/T of fifteen mechanism functions and finding out the best mechanism function with the maximum correlation coefficient. For reduction of Fe2O3 by hydrogen at temperatures less than 400 °C, the best mechanism function was the Anti-Jander 3D function. This indicated that the reduction was dominated by a 3D diffusion model. For temperatures >400°C, Jander 2D (n = 2) function was determined to be the mechanism function of the Fe2O3 systems, and the reduction reaction follows a 2D diffusion model. The reduction process for the Fe2O3-Ni and Fe2O3-NiO systems was divided into two stages, <400 °C and >400 °C, and the kinetic mechanisms investigated. For temperatures <400 °C, Anti-Jander 3D function was the best mechanism function for the Fe2O3-Ni-H2 systems, and the reduction reaction follows a 3D diffusion model. For temperatures > 400 °C, Mampel Power (n = 2) function was determined to be the mechanism function of the Fe2O3-Ni-H2 systems, and the reduction reaction follows the Mampel Power law. For the Fe2O3-NiO-H2 system, at temperatures <400 °C, the Anti-Jander 3D function was the best mechanism function, and the reduction reaction follows a 3D diffusion model. For temperatures > 400 °C, Jander 2D (n = 2) function was the best mechanism function, and the reduction reaction follows a 2D diffusion model. The apparent activation energies for the reduction reactions are shown in Table 3.
According to the data in Table 3, the activation energy of Fe2O3 system doped with Ni components is lower than that of Fe2O3 system at two temperature ranges, whether the reaction temperature is lower than 400 °C or higher than 400 °C. The high activation energy is not conducive to the reaction, indicating that doping Ni components in Fe2O3 system is conducive to the reduction reaction. The activation energies of reduction process of Fe2O3-Ni, Fe2O3-NiO and Fe2O3 are 172 kJ·mol−1, 197 kJ·mol−1 and 263 kJ·mol−1, respectively. The difficulty of reduction reaction in three systems is represented from low to high: Fe2O3-Ni < Fe2O3-NiO < Fe2O3; it is consistent with the conclusion in Figure 4.

4. Conclusions

Based on DFT calculation, it can be seen that H2 molecule is more easily adsorbed on the α-Fe2O3 (001) surface when Ni atoms exist. It can be concluded that Ni components can promote the reduction of Fe2O3 in H2 atmosphere. This conclusion is verified by reduction degree curve and reduction kinetics. The absolute value of adsorption energy of (H2 + Ni)/α-Fe2O3 (001) system is 0.93 eV higher than that of H2/α-Fe2O3 (001) system. It is shown that the presence of Ni atom promotes the adsorption of H2 on α-Fe2O3 (001) surface. The existence of Ni atom has a great influence on the structure of H2 molecule and α-Fe2O3 (001) surface. The reduction of Fe2O3 by H2 is a multi-stage process, when the temperature is below 400 °C and the reduction reaction follows a 3D diffusion model. At temperatures above 400 °C the reduction reaction follows a 2D diffusion model. The activation energies of two temperature stages are 180 kJ·mol−1 and 83 kJ·mol−1, respectively. Ni components are determined as accelerators in Fe2O3 system, and the reaction termination temperatures of Fe2O3 were reduced. When Ni components were added, the activation energies of reduction process of Fe2O3 system decreased.

Author Contributions

Funding acquisition, H.W.; investigation, H.Z. and B.L.; methodology, H.Z., B.L., Y.W. and H.W.; project administration, B.L. and Y.W.; resources, B.L., Y.W. and H.W.; software, H.Z.; supervision, H.W.; writing—original draft, H.Z. and B.L.; writing—review and editing, B.L., Y.W. and H.W.

Acknowledgments

This research was funded by the Key R&D Program of Yunnan Province (2018IA055), National Natural Science Foundation of China (No. U1602272 and 51664039) and the Analysis and Testing Foundation of Kunming University of Science and Technology (2018M20172102040).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Quaicoe, I.; Nosrati, A.; Skinner, W.; Addai-Mensah, J. Agglomeration and column leaching behaviour of goethitic and saprolitic nickel laterite ores. Miner. Eng. 2014, 65, 1–8. [Google Scholar] [CrossRef]
  2. Dabrowa, J.; Kucza, W.; Cieslak, G.; Kulik, T.; Danielewski, M.; Yeh, J.W. Interdiffusion in the FCC-structured Al-Co-Cr-Fe-Ni high entropy alloys: Experimental studies and numerical simulations. J. Alloy. Compd. 2016, 674, 455–462. [Google Scholar] [CrossRef]
  3. Ma, B.Z.; Wang, C.Y.; Yang, W.J.; Yin, F.; Chen, Y.Q. Screening and reduction roasting of limonitic laterite and ammonia-carbonate leaching of nickel-cobalt to produce a high-grade iron concentrate. Miner. Eng. 2013, 46, 106–113. [Google Scholar] [CrossRef]
  4. Shirazi, H.; Miyamoto, G.; Hossein Nedjad, S.; Chiba, T.; Nili Ahmadabadi, M.; Furuhara, T. Microstructure evolution during austenite reversion in Fe-Ni martensitic alloys. Acta Mater. 2018, 144, 269–280. [Google Scholar] [CrossRef]
  5. Ma, B.Z.; Wang, C.Y.; Yang, W.J.; Chen, Y.Q.; Yang, B. Comprehensive utilization of Philippine laterite ore, part 1: Design of; technical route and classification of the initial ore based on; mineralogical analysis. Int. J. Miner. Process. 2013, 124, 42–49. [Google Scholar] [CrossRef]
  6. Hu, M.; Laghari, M.; Cui, B.H.; Xiao, B.; Zhang, B.P.; Guo, D.B. Catalytic cracking of biomass tar over char supported nickel catalyst. Energy 2018, 145, 228–237. [Google Scholar] [CrossRef]
  7. Parr, R.G. Density Functional Theory. Chem. Eng. News 1983, 68, 2470–2484. [Google Scholar] [CrossRef]
  8. Sun, X.; Hwang, J.Y.; Xie, S.Q. Density functional study of elemental mercury adsorption on surfactants. Fuel 2011, 90, 1061–1068. [Google Scholar] [CrossRef]
  9. de Oliveira, C.; Duarte, H.A. Disulphide and metal sulphide formation on the reconstructed (001) surface of chalcopyrite: A DFT study. Appl. Surf. Sci. 2010, 257, 1319–1324. [Google Scholar] [CrossRef]
  10. Cores, A.; FormosoO, A.; Larrea, M.T.; Qrtiz, J. Kinetic regularities in joint reduction of nickel and iron oxides under non-isothermal conditions. Ironmak. Steelmak. 1989, 16, 446–449. [Google Scholar]
  11. Zhou, S.W.; Wei, Y.G.; Peng, B.; Li, B.; Wang, H. Reduction kinetics for Fe2O3/NiO-doped compound in a methane atmosphere. J. Alloy. Compd. 2017, 735, 365–371. [Google Scholar] [CrossRef]
  12. Abdel-Halim, K.S.; Khedr, M.H.; Nasr, M.I.; Abdel-wahab, M.S. Carbothermic reduction kinetics of nanocrystallite Fe2O3/NiO composites for the production of Fe/Ni alloy. J. Alloy. Compd. 2008, 463, 585–590. [Google Scholar] [CrossRef]
  13. Feng, Q.C.; Wen, S.M.; Deng, J.S.; Zhao, W.J. Combined DFT and XPS investigation of enhanced adsorption of sulfide species onto cerussite by surface modification with chloride. Appl. Surf. Sci. 2017, 425, 8–15. [Google Scholar] [CrossRef]
  14. Weiss, W.; Ranke, W. Surface chemistry and catalysis on well-defined epitaxial iron-oxide layers. Prog. Surf. Sci. 2002, 70, 1–151. [Google Scholar] [CrossRef]
  15. Tao, L.; Guo, X.; Zheng, C.G. Density functional study of Hg adsorption mechanisms on α-Fe2O3 with H2S. Proc. Combust. Inst. 2013, 34, 2803–2810. [Google Scholar] [CrossRef]
  16. Liu, T.; Xue, L.C.; Guo, X.; Zheng, C.G. DFT study of mercury adsorption on α-Fe2O3 surface: Role of oxygen. Fuel 2014, 115, 179–185. [Google Scholar] [CrossRef]
  17. Alvarez-Ramirez, F.; Martinez-Magadan, J.M.; Gomes, J.R.B.; Illas, F. On the geometric structure of the (0001) hematite surface. Surf. Sci. 2004, 558, 4–14. [Google Scholar] [CrossRef]
  18. Bergermayer, W.; Schweiger, H. Ab initio thermodynamics of oxide surfaces: O2 on Fe2O3(0001). Phys. Rev. B 2004, 69, 195409–195421. [Google Scholar] [CrossRef]
  19. Lin, C.F.; Qin, W.; Dong, C.Q. H2S adsorption and decomposition on the gradually reduced-Fe2O3 (001) surface: A DFT study. Appl. Surf. Sci. 2016, 387, 720–731. [Google Scholar] [CrossRef]
  20. Sun, Y.Q.; Lu, Q.; Wan, X.Y.; Zhang, S.H.; Zhang, J.M. Computational Insights into Interactions between Ca Species and α-Fe2O3 (001). J. Iron Steel Res. Int. 2014, 21, 413–418. [Google Scholar] [CrossRef]
  21. Xu, Y.D.; Zheng, Y.Y.; Mao, J.H.; Xia, P. Analysis on corrosion properties of reinforcement in concrete in chloride environment based on first-principles. J. Southeast Univ. (Nat. Sci. Ed.) 2016, 46, 619–623. [Google Scholar]
  22. Hu, R.Z.; Shi, Q.Z. Thermal Analysis Kinetics; Science Press: Beijing, China, 2008; pp. 2–8. [Google Scholar]
  23. Li, B. Basic Study on Drying Characteristics and Pre-Reduction of Garnierite. Ph.D. Thesis, Kunming University of Science and Technology, Kunming, China, 2012. [Google Scholar]
  24. Zhang, H.P.; Li, B.; Wei, Y.G.; Wang, H.; Yang, Y.D.; Alexander, M. Nonisothermal reduction kinetics in the Fe-Cu-O system using H2. JOM 2019, 5, 1813–1821. [Google Scholar] [CrossRef]
  25. Zhao, W. Research on the Equilibrium Geometric Structure and Electronic Structure in the Interface of Metal Iron and the Solution. Ph.D. Thesis, Tsinghua University, Beijing, China, 2008. [Google Scholar]
  26. Dong, C.Q.; Liu, X.L.; Qin, W.; Lu, Q.; Wang, X.Q.; Shi, S.M.; Yang, Y.Q. Deep reduction behavior of iron oxide and its effect on direct CO oxidation. Appl. Surf. Sci. 2012, 258, 2562–2569. [Google Scholar] [CrossRef]
  27. Hu, L.; Liu, D.C.; Chen, X.M.; Yang, B.; Bai, P.P.; Duan, S.F. Thermal decomposition of gallium arsenide under vacuum: Theoretical calculation and experiment. Chin. J. Nonferr. Met. 2014, 24, 2410–2417. [Google Scholar]
  28. Kang, T.O.; Lee, K.I.; Yoon, J.K. The reduction mechanism of nickel oxide with graphite. J. Korean Inst. Met. Mater. 1977, 15, 147–154. [Google Scholar]
  29. Nasr, M.I.; Omar, A.A.; Khedr, M.H.; El-Geassy, A.A. Effect of nickel oxide doping on the kinetics and mechanism of iron oxide reduction. ISIJ Int. 1995, 35, 1043–1049. [Google Scholar] [CrossRef]
  30. Guo, D.B.; Hu, M.; Pu, C.X.; Xiao, B.; Hu, Z.Q.; Liu, S.M.; Wang, X.; Zhu, X.L. Kinetics and mechanisms of direct reduction of iron ore-biomass composite pellets with hydrogen gas. Int. J. Hydrogen Energy 2015, 40, 4733–4740. [Google Scholar] [CrossRef]
  31. Ma, B.Z.; Xing, P.; Yang, W.J.; Wang, C.G.; Chen, Y.Q.; Wang, H. Solid-state metalized reduction of magnesium-rich low-nickel oxide ores using coal as the reductant based on thermodynamic analysis. Metall. Mater. Trans. B 2017, 48, 2037–2046. [Google Scholar] [CrossRef]
  32. Zhang, Y.; Wei, W.; Yang, X.; Wei, F. Reduction of Fe and Ni in Fe-Ni-O systems. J. Min. Metall. B 2013, 49, 13–20. [Google Scholar] [CrossRef]
Figure 1. Diagram of α-Fe2O3 (001) clean surface.
Figure 1. Diagram of α-Fe2O3 (001) clean surface.
Metals 09 01292 g001
Figure 2. Initial adsorption location of H2 molecule and Ni atom.
Figure 2. Initial adsorption location of H2 molecule and Ni atom.
Metals 09 01292 g002
Figure 3. DOS of (a) α-Fe2O3 (001) surface (black line: H2/α-Fe2O3 (001) system; red line: (H2 + Ni)/α-Fe2O3 (001) system), (b) H2 molecule (blue line: H2/α-Fe2O3 (001) system; pink line: (H2 + Ni)/α-Fe2O3 (001) system).
Figure 3. DOS of (a) α-Fe2O3 (001) surface (black line: H2/α-Fe2O3 (001) system; red line: (H2 + Ni)/α-Fe2O3 (001) system), (b) H2 molecule (blue line: H2/α-Fe2O3 (001) system; pink line: (H2 + Ni)/α-Fe2O3 (001) system).
Metals 09 01292 g003
Figure 4. Degree of reduction within the three systems when exposed to H2.
Figure 4. Degree of reduction within the three systems when exposed to H2.
Metals 09 01292 g004
Figure 5. The reaction rate curve of three systems.
Figure 5. The reaction rate curve of three systems.
Metals 09 01292 g005
Figure 6. XRD patterns of three samples after reduction for different temperature (a): Fe2O3; (b): Fe2O3-NiO; (c): Fe2O3-Ni.
Figure 6. XRD patterns of three samples after reduction for different temperature (a): Fe2O3; (b): Fe2O3-NiO; (c): Fe2O3-Ni.
Metals 09 01292 g006aMetals 09 01292 g006b
Table 1. Adsorption energy and structure parameters of two adsorption systems.
Table 1. Adsorption energy and structure parameters of two adsorption systems.
SystemEads/eVdH-HdO-Fe△αH2-surface/(°)
H2/α-Fe2O3 (001)−1.880.7582.02530
(H2 + Ni)/α-Fe2O3 (001)−2.810.8892.025863.36
Table 2. Some mechanism functions for gas-solid reactions.
Table 2. Some mechanism functions for gas-solid reactions.
No.Equationf (α)G(α)
1Mampel Power (n = 1)1α
2Mampel Power (n = 3/2)2/3α−1/2α3/2
3Mampel Power (n = 2)1/2α−1α2
4Jander 2D (n = 2)(1 − α)1/2[1 − (1 − α)1/2]−1[1 − (1 − α)1/2]2
5Jander 3D (n = 2)3/2(1 − α)2/3[1 − (1 − α)1/3]−1[1 − (1 − α)1/3]2
6G-B3/2[(1 − α)−1/3 − 1]−11 − 2/3α − (1 − α)2/3
7Anti Jander 3D3/2(1 + α)2/3[(1 + α)1/3 − 1]−1[(1 + α)1/3 − 1]2
8Avrami-Erofeev (n = 3/2)2/3(1 − α)[−ln(1 − α)]−1/2[−ln(1 − α)]3/2
9Avrami-Erofeev (n = 2)1/2(1 − α)[−ln(1 − α)]−1[−ln(1 − α)]2
10Avrami-Erofeev (n = 3)1/3(1 − α)[−ln(1 − α)]−2[−ln(1 − α)]3
11Avrami-Erofeev (n = 4)1/4(1 − α)[−ln(1 − α)]−3[−ln(1 − α)]4
12Interface Reaction R3 (n = 1/3)3(1 − α)2/31 − (1 − α)1/3
13Reaction Order (n = 1/4)4(1 − α)3/41 − (1 − α)1/4
14Interface Reaction R3 (n = 3)(1 − α)2/33[1 − (1 − α)1/3]
15Reaction Order (n = 3)(1 − α)3/22[(1 − α)−1/2 − 1]
Table 3. Apparent activation energies and mechanism functions for the reduction of Fe2O3, Fe2O3-Ni and Fe2O3-NiO compacts.
Table 3. Apparent activation energies and mechanism functions for the reduction of Fe2O3, Fe2O3-Ni and Fe2O3-NiO compacts.
SampleTemperature Range (°C)Activation Energy (kJ·mol−1)Mechanism
Fe2O3<400180Anti-Jander 3D
>40083Jander 2D (n = 2)
Fe2O3-Ni<400106Anti-Jander 3D
>40066Mampel Power (n = 2)
Fe2O3-NiO<400127Anti-Jander 3D
>40070Jander 2D (n = 2)

Share and Cite

MDPI and ACS Style

Zhang, H.; Li, B.; Wei, Y.; Wang, H. Combined Density Functional Theory and Reduction Kinetics Investigation of Enhanced Adsorption of Hydrogen onto Fe2O3 by Surface Modification with Nickel. Metals 2019, 9, 1292. https://doi.org/10.3390/met9121292

AMA Style

Zhang H, Li B, Wei Y, Wang H. Combined Density Functional Theory and Reduction Kinetics Investigation of Enhanced Adsorption of Hydrogen onto Fe2O3 by Surface Modification with Nickel. Metals. 2019; 9(12):1292. https://doi.org/10.3390/met9121292

Chicago/Turabian Style

Zhang, Haipei, Bo Li, Yonggang Wei, and Hua Wang. 2019. "Combined Density Functional Theory and Reduction Kinetics Investigation of Enhanced Adsorption of Hydrogen onto Fe2O3 by Surface Modification with Nickel" Metals 9, no. 12: 1292. https://doi.org/10.3390/met9121292

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