Next Article in Journal
High Photocatalytic Activity under Visible Light for a New Morphology of Bi2WO6 Microcrystals
Previous Article in Journal
Biocatalytic Synthesis of Natural Dihydrocoumarin by Microbial Reduction of Coumarin
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Oxygen Atom Function: The Case of Methane Oxidation Mechanism to Synthesis Gas over a Pd Cluster

1
College of Chemistry and Chemical Engineering, Taiyuan University of Technology, Taiyuan 030024, China
2
School of Chemistry and Chemical Engineering, Liaocheng University, Liaocheng 252000, China
3
Institute of Coal Chemistry, Chinese Academy of Sciences, Taiyuan 030001, China
*
Authors to whom correspondence should be addressed.
Catalysts 2019, 9(8), 666; https://doi.org/10.3390/catal9080666
Submission received: 19 June 2019 / Revised: 18 July 2019 / Accepted: 1 August 2019 / Published: 4 August 2019

Abstract

:
A dimer model Pd2 was established to study the adsorption of CHx (x = 1–4) and CH4 dehydrogenation, as well as syngas formation using density functional theory (DFT) at the atomic level. Meanwhile, insight into understanding the role of the oxygen atom on the partial oxidation of methane (POM) was also calculated based on a trimer model of Pd2O. For the adsorption of CHx, results showed that the presence of an oxygen atom was a disadvantage to the adsorption of CHx (x = 1–3) species. For CH4 dissociation, the process of CH2→CH + H was found to be the rate-limiting step (RSD) on both Pd2 and Pd2O. H2 was formed by the reaction of CH2 + 2H→CH2 + H2. For CO formation, it was primarily formed in the process of CH + O→CHO→CO + H on both the Pd2 and the Pd2O catalyst. Thermodynamic and kinetic calculations revealed that formation and maintainance of the oxygen atom on the Pd surface could promote a POM reaction to achieve high H2 and CO yield and selectivity. Our study provides a helpful understanding of the effect of an adsorbed oxygen atom on a POM reaction with a Pd catalyst.

1. Introduction

Methane (CH4), the main component of coalbed gas, is directly released to the environment in an amount of around 19 billion cubic meters per year [1]. In the consideration of energy conservation, environmental protection, and rational energy utilization, efficient utilization of methane has become increasingly important [2,3,4,5,6,7]. Due to the mildly exothermic heat property and suitable production ratio of H2/CO (syngas) for methanol and Fischer–Tropsch synthesis [8,9], conversions of methane via the partial oxidation of methane (POM) have been proposed in previous studies [10,11,12,13].
Metallic palladium (Pd) is one of the most efficient catalysts for CH4 oxidation, which was outlined both upon experimental and theoretical calculation [14,15,16,17,18]. Liu et al. claimed that Pd/TiO2 nanoparticles shows good activity for methane oxidation at low temperatures [19]. Li et al. investigated the synthesis gas production from partial oxidation of methane over highly dispersed Pd/SiO2 catalyst. It was found that the reaction of methane conversion on highly dispersed Pd catalyst was more activity and stability than the other catalysts [20]. However, it is difficult to prepare and research performance small cluster catalyst alone in experiment, but it can be easily solved by density functional theory (DFT). With the further research, Johan et al. investigated methane oxidation on Pd/Al2O3 by in situ X-ray absorption fine structure (XAFS) characterization. It was found that methane conversion was increased with the oxidation of Pd, which indicated that PdO was the active phase for methane oxidation [15]. Moreover, Xiong et al. found that Pd and PdO could coexist within a single particle and be mutually transformed at high temperatures by the methods of in situ X-ray diffraction (XRD) and a JEOL 2010 microscope transmission electron microscopy (TEM). A polycrystalline catalyst was formed when Pd and PdO were exposed to methane and oxygen together [21,22,23].
In addition to the above experiments, methane activation on a Pd bulk surface was also largely investigated using the density functional theory (DFT) method [24,25,26,27,28]. Adsorption and dehydrogenation of CH4 on Pd (100) and PdO (100) surfaces were studied in previous works that found that the effect of coadsorbed oxygen atoms tends to weaken the adsorbate–substrate interaction. However, the ordered metal surface changes into an amorphous state, identified as being dependent on temperature. Therefore, only obtaining results of the reactions on the surfaces were inadequate. In recent years, researchers have focused on the effect of methane activation on clusters. Trevor et al. observed methane activation on Pt clusters of 2–24 atoms by using time-of-flight mass spectroscopy; the experiments showed that the most activity-neutral cluster was Pt2–Pt5 [29]. The nanocluster had a large surface area and abundant defects, which suggested higher reactivity for the oxidation of methane than that of the bulk surface [17,30,31,32,33,34]. Therefore, an intuitive idea was that the study of the POM mechanism on a Pd cluster would be better matched with the actual reaction process.
In this work, a Pd cluster model of dimer Pd2 was established to study the reaction mechanisms of POM at the atomic level, while a similar cluster model of trimer Pd2O was used to investigate the effect of oxygen on methane activation in a Pd system. The adsorption of CHx (x = 1–-4) and methane dehydrogenation on dimer Pd2 and trimer Pd2O were studied by DFT. Subsequently, the formation of hydrogen and carbon monoxide were also identified in detail. Plausible reaction mechanisms and the effect of an oxygen atom on POM were illuminated.

2. Results and Discussion

The structures of the model clusters are shown in Figure 1. In dimer Pd2, the diatomic distance was 2.545 Å, which was in good agreement with the experiment result of 2.55 Å that Nasresfahani et al. obtained [35]. In trimer Pd2O, the Pd–Pd bond length was elongated to 2.635 Å. This indicated that, with the adsorption of the oxygen atom, the interaction between two Pd atoms was decreased. CH4 dehydrogenation and oxidation were considered as necessary processes for POM in order to understand the effect of the oxygen atom that Pd-catalyzed the reaction sufficiently.

2.1. Intermediate System in the Reaction

With similar reaction paths, methane dehydrogenation processes are presented as four competitive reaction steps [36,37,38,39]: methyl (CH3), methylene (CH2), methylidene (CH), and carbon (C) were generated in an orderly fashion during the cracking. The stable configurations of CH4–Pd2(o), CH3–Pd2(o), CH2–Pd2(o), and CH–Pd2(o) were obtained by geometry optimization, respectively. The length of the chemical bonds is shown in Figure 2, and the coordination energies are listed in Table 1.
As shown in Figure 2 and Table 1, CH4 dissociation catalyzed by Pd2 and Pd2O results in an exothermic formation of CH4–Pd2 and CH4–Pd2O systems. In the CH4–Pd2 system, the length of the C–Pd bond was 2.657 Å, while the length decreased to 2.432 Å in CH4–Pd2O [37]. The coordination energy of the two systems was –0.34 and –0.87 eV, respectively. These coordination-energy results, further confirmed that our work was reliable as they were consistent with previous work [38]. Similarly to the process of CH4 dissociation catalyzed by Pd2 and Pd2O, CHx (x = 0–3) dissociation catalyzed by Pd2 and Pd2O were exothermic reactions (Figure 2). For CH3–Pd2, the length of the C–Pd bond was 2.007 Å, while the length increased to 2.235 Å as CH3 was catalyzed by Pd2O. The distance between C and Pd was 1.988 Å in the CH2–Pd2 system, and 2.017 Å in the CH2–Pd2O system, and these lengths were lower than those of the CH3 system. The coordination energies of CH3–Pd2 and CH3–Pd2O were −2.63 and −2.42 eV, respectively. The coordination energy of Pd2 and Pd2O increased to −4.90 and −4.08 eV, respectively, when CH3 was replaced by CH2. The length of the C–Pd bonds of CH–Pd2 and CH– Pd2O were 1.829 and 1.867 Å, respectively. The coordination energies of the CH–Pd2 and CH–Pd2O systems were increased to −5.70 and −4.66 eV, respectively. These results indicated that the coordination energy decreased with the cracking process: CH4 < CH3 < CH2 < CH < C [26,37,38]. They also revealed that the interaction between CHx (x = 0–4) and the catalyst was reduced by the introduction of the oxygen atom.
To further confirm the effect of the oxygen atom on Pd2, the partial density of states (PDOS) of CHx–Pd2(O) was investigated. In Figure 3a, it is shown that CH4 was hardly activated due to the superwide orbital energy gap of CH4, which was about 8 eV. When CH4 was adsorbed on Pd2O, PDOS peaks were slightly shifted to more negative values than those of the CH4–Pd2 system, indicating the charge transfer from Pd2 to CH4. As shown in Figure 3b,c, when CH3 and CH2 were catalyzed by Pd2O, their peaks shifted to more negative values than those of the CH3/CH2–Pd2 systems. In Figure 3d, we can see that a similar phenomenon was observed in CH.

2.2. Methane Activation

In order to illuminate the effect of an oxygen atom on CH4 dehydrogenation, the cleavage energy of CH4 on Pd2 and Pd2O was calculated at the same level. In the reaction, CHX with the most stable geometry was chosen as the initial state, and the CHX–1+H with the most stable coadsorbed geometry was chosen as the final state. The transitional reaction states on Pd2 and Pd2O were labeled for TS1–TS4 and TS1’–TS4’, respectively. The configuration of various structures involved in the reaction mechanism associated with the change from CHx–Pd2(O) to CHx–1 + H–Pd2(O) are shown in Figure 4 and Figure 5.

2.2.1. Potential Energy Changes of Methane Activation

An energy barrier and energy changes were generally used to study the reaction mechanism. The energy changes are listed in Tables S1 and S2. The potential energy curves for CH4 dissociation on Pd2 and Pd2O are shown in Figure 6.
In the process whereby CH4→CH3 + H, the energy barrier (Eb) on Pd2 and Pd2O was 0.61 and 1.09 eV, respectively, which was in good agreement with CH4 dissociation over the Pt2 dimer and Pd(111) [40,41]. This indicated that CH4 activation on Pd2O is relatively difficult to achieve compared with that on Pd2 [42,43]. This reaction was an exothermic process (reaction energy (Er) < 0) on Pd2 but an endothermic process (Er > 0) on Pd2O.
For CH3 + H→CH2 + 2H, Eb on both Pd2 and Pd2O was higher than the first C–H dissociation. Eb was 1.03 and 1.28 eV on Pd2 and Pd2O, respectively. It was shown that the second C–H bond dissociated on Pd2O requires higher energy than that on Pd2 [41,43]. This reaction was an endothermic process (Er > 0) both on Pd2 and Pd2O. Therefore, the POM experiment required a high reaction temperature, about 800 °C, at the beginning [10].
In terms of CH2 + 2H→CH + 3H, the Eb reached 2.18 eV on Pd2 and 1.71 eV on Pd2O. The largest reaction barriers implied that this step was the rate-determining steps (RDS) in the process of CH4 activation on Pd2 and Pd2O, which was consistent with reported works [26,42].
For the step of CH + 3H→C + 4H, the Eb on Pd2 and Pd2O was 1.58 and 1.33 eV, respectively. The reactions on Pd2 and Pd2O were both endothermic processes (Er > 0) [26,43]. Above all, compared with the reaction energy barrier and reaction heat details of the steps of CH4 dehydrogenation on Pd2 and Pd2O, it can be seen that, except for CH4 and CH3, the adsorption of the oxygen atom was conducive to CHx (X = 1,2) dissociation.

2.2.2. Thermodynamic and Kinetic Analysis of Methane Activation

Furthermore, the thermodynamics and kinetics of the reaction on the Pd2(O) cluster were studied, as shown in Table 2 and Table 3. The activation energy (Ea) of CH4→CH3 + H was 0.45 eV on Pd2, but 1.78 eV on Pd2O, indicating that the cleavage of the first C–H bond is easier in the CH4–Pd2 system than that in the CH4–Pd2O system. The corresponding rate constant (κ) values of the two reactions, verified by a conclusion from a previous source, were 1.75 × 1010 and 2.09 × 101, respectively. CH3 + H→CH2 + 2H on Pd2(O) had the same trend as the first H atom dissociation.
The Ea of CH2 + 2H→CH + 3H was 3.13 eV on Pd2 and 2.65 eV on Pd2O. The corresponding κ values were 2.17 × 10–6 and 6.25 × 10−3, respectively. The relatively low Ea and high κ indicated that the cracking process from CH2 to CH on Pd2O could easily occur. In addition, CH2 + 2H→CH + 3H had the highest Ea and the lowest κ compared to the other dehydrogenation reaction on Pd2(O), indicating that CH2→CH was the RDS for CH4 decomposition on Pd2(O).

2.3. Hydrogen Production

The structure of MCH2 + H2 is an important intermediate during hydrogen production [40,44]. Thus, the structures of the corresponding intermediates of Pd2 and Pd2O were calculated. The frequency calculations of transition-state (TS) geometries are shown in Figure 7 and Figure 8. The transition states of the reaction on Pd2 and Pd2O were marked as TS5–TS7 and TS5’–TS7’, respectively. All frequencies of the initial states (IS), intermediate states, and the final state (FS) were positive, which indicates that their structures were stabilized. Intrinsic reaction coordinate (IRC) calculations showed that all optimized transition states correctly connected the relevant reactants and products.

2.3.1. Potential Energy Changes of Hydrogen Generation

The potential energy changes of hydrogen generation were researched to reflect the hydrogen elimination mechanism. As shown in Figure 9, two distinguishable reaction paths were used to describe the formation of H2 on the Pd2 and Pd2O catalysts. The results of potential energy changes are listed in Tables S3 and S4. The energy barrier of the CH2 + 2H→CH2 + H2 reaction was low, as a result of no need for any chemical-bond breakage, similarly to what occurred on Pd2 and Pd2O. From Figure 9, we can see that the barrier and overall heat absorption on Pd2 were relatively bigger than those on Pd2O. This illustrated that the presence of oxygen atoms was beneficial to the formation of hydrogen.

2.3.2. Thermodynamic and Kinetic Analysis of Hydrogen Generation

Ea and κ were calculated to understand the reaction processes on Pd2(O) in terms of the thermodynamics and kinetics. The relevant data are shown in Table 4 and Table 5. The reactions of hydrogen generation included CH2 + 2H→CH2 + 2HS, CH2 + 2H→CH2 + 2HN, and CH2 + 2H→CH2 + H2. The corresponding Ea and κ of these reactions on Pd2 were 0.03, 0.68, and 1.11 eV, and 2.44 × 1012, 2.20 × 109, and 5.27 × 106 s−1, respectively, and the corresponding Ea and κ values of the above reaction were 0.23, 0.48, and 1.08 eV, and 1.13 × 1012, 1.80 × 1011, and 5.77 × 106 s−1, respectively. It was seen that hydrogen generation in the presence of an oxygen atom was easier in terms of the thermodynamics of the reaction.

2.4. Carbon Monoxide Formation

The oxidation of CH4 to CO involves the oxidation process of CH and C. In order to investigate the effect of oxygen atoms on the formation of CO on a Pd catalyst, three possible reaction pathways of CO formation on Pd2 and Pd2O were studied using DFT calculation [45,46]. The processes were: (1) CH→C + H, C + O→CO; (2) CH + O→CHO→CO + H; and (3) CH + O→COH→CO + H.

2.4.1. Potential Energy Changes of Carbon Monoxide Formation

The geometric optimization structure of reactants, products, intermediates, and transition states in the oxidation process of CH and C are shown in Figure 10. All reactants and products structures were stable due to the positive calculations of frequencies. The transition states of reaction on Pd2 and Pd2O were denoted as TS8–TS13 and TS8’–TS13’, respectively.
Three distinguishable reaction paths for the formation of CO on Pd2 are shown in Figure 7, and the corresponding energy is shown in Table S5. In the first reaction path, two processes were involved. The reaction energy barrier (Eb) of the CH→C + H and C + O→CO processes was about 1.26 and 0.46 eV, respectively. The next two reactions were followed by a series of reactions. The maximum activation energy barriers in the reactions were CH + O→CHO→CO + H and CH + O→COH→CO + H, whose energy barriers were 0.70 and 0.84 eV for the reaction, respectively. Both reactions were exothermic (Er < 0).
Subsequently, the formation process of carbon monoxide on Pd2O was calculated and the effect of oxygen atoms on the process was discussed. The reaction processes and energies are shown in Table S6. The reactive barrier (Eb) in the first reaction path was 0.93 eV for the process of CH→C + H, and 0.39 eV for the process of C + O→CO. The maximum reactive energy barriers of reactions CH + O→CHO→CO + H and CH + O→COH→CO + H were 0.46 and 0.54 eV, respectively.

2.4.2. Thermodynamic and Kinetic Analysis of Carbon Monoxide Formation

Thermodynamic and kinetic data for the reaction on the Pd2 catalyst are listed in Table 6. In the first process, the activation energy (Ea) of CH→C + H and C + O→CO was 2.70 and 1.49 eV, respectively. The maximum activation energy (Ea) of the second possible pathway of CH + O→COH→CO + H was 1.73 eV. Compared with the above processes, the maximum activation energy (Ea) of CH + O→CHO→CO + H was relatively low, at about 1.60 eV. Corresponding rate constants (κ) were also in good agreement with the trend of their activation energies. The results indicated that the CH + O→CHO→CO + H reaction was the best pathway for the reaction on Pd2, both thermodynamically and kinetically.
All thermodynamic and kinetic data for the reaction on Pd2O are listed in Table 7. The minimum activation energy (Ea) was found to be about 1.19 eV in CH + O→CHO→CO + H. The corresponding reaction rate (κ) was 2.68 × 1013 s–1. Thus, CH + O→CHO→CO + H was the best pathway of the reaction on Pd2O, both thermodynamically and kinetically.

3. Computational Details

All calculations were performed using DFT in generalized gradient approximation (GGA), Perdew Burke, and Ernzerho (PBE) [36]. Exchange correlation was implemented with module DMol3 in Material Studio of Accelrys Inc. The double-numeric quality plus polarization (DNP) functions were applied to all atoms. Considering the calculation cost, all core electrons were represented by DFT semicore pseudopodium functions. The convergence criteria for energy, force, and displacement convergence were constant at 2 × 10−5 hartree, 4 × 10−3 hartree/Å, and 5 × 10−3 Å.
The frequency of all intermediates and transition states was calculated at the same level in order to determine the minimum (no imaginary frequencies) and the first-order saddle point (an imaginary frequency representing the required reaction coordinates). Transition states (TSs) were confirmed by LST/QST (linear synchronous transit and quadratic synchronous transit) method to obtain zero-point-energy (ZPE) corrections. With the aim to define each saddle point linking two desired minima, IRC calculations were used.
Considering the dissociation of methane, we first examined the adsorption direction and energies. Binding energies are defined by the following equation:
Ebinding = E(adsorbates/clues) − Eclues − Eadsorbates,
where: E(adsorbates/clues), total energy of adsorbates on Pd2; Eclues, total energy of Pd2/Pd2O; Eadsorbates, total energy of free adsorbates.
Convergence tolerance in TSs turned out to be the same as in the geometry optimization. Then, the electron energies (Eelec) and Mulliken atomic charges of all the structures were calculated. Standard entropy (S), enthalpy (H), and Gibbs free energy (G) were also obtained by using frequency analysis. The forward or reverse rates of the elementary step within the reaction stayed constant because of the constant enthalpy and entropy differences.
The energy barrier (Eb) is presented by the difference in the Gibbs free energy between the transition state and initial state. The reaction energy (Er) is given by the difference of enthalpy between the final state and initial state.
Eb = GTS − GIS,
Er = HFS − HIS,
where GTS is the Gibbs free energy of transition state; GIS is the Gibbs free energy of initial state; HFS and HIS are the enthalpy of the final and initial state, respectively.
According to the thermodynamics process of transition state theory (TST), the Arrhenius activation energy (Ea) is given by the following equation:
Ea = RT + ∆H,
where R is the gas constant; T is the reaction temperature; ∆H is the change of enthalpy between the transition state and initial state.
By using DFT to study the oxidation of methane, temperature and pressure are taken into account. Therefore, the pre-exponential factor A and the rate constant of a reaction k can also be determined according to TST:
A = k b T h exp ( Δ rS m R ) ,
where ∆rSm is the difference of entropy between the transition state and initial state; kb is the Boltzmann constant; h is the Planck constant.
Reaction rate constant k can be given as the following:
k = k b T h exp ( Δ rG m R ) = k b T h exp ( Δ rS m R ) exp ( Δ rH m RT )

4. Conclusions

In summary, CH4 adsorption and configuration evolution of configurations on Pd2 and Pd2O were studied using DFT calculations. Binding energy CHx (x = 1–3), adsorbed on Pd2, decreased with the oxygen atom adsorption. However, the energy barriers for CH4 dissociation increased with the presence of the oxygen atom. The reaction of CH2→CH + H was the RDS for CH4 activation on both the catalysts. In syngas synthesis, H2 was formed by the reaction of CH2 + 2H→CH2 + H2, and the energy barriers of this reaction were reduced with the presence of the oxygen atom, which indicated that H2 formation was easier on Pd2O than on Pd2. For CO formation, it was mainly formed in the process of CH + O→CHO→CO + H on both the Pd2 and the Pd2O catalyst. The addition of oxygen atoms was also beneficial to CO formation, resulting from thermodynamic and kinetic calculations. Thus, forming and maintaining an oxygen atom on the Pd surface could promote a POM reaction to achieve high H2 and CO selectivity. Our study provides a helpful understanding of the effect of an adsorbed oxygen atom to a POM reaction with a Pd catalyst.

Supplementary Materials

The following are available online at https://www.mdpi.com/2073-4344/9/8/666/s1, Table S1: Gibbs free energies (G), energy barriers (Eb), enthalpies (H), and reaction energies (Er) in the process of methane activation by Pd2. Table S2: Gibbs free energies (G), energy barriers (Eb), enthalpies (H) and reaction energies (Er) in the process of methane activation by Pd2O. Table S3: Gibbs free energies (G), energy barriers (Eb), enthalpies (H), and reaction energies (Er) in the process of hydrogen generation by Pd2. Table S4: Gibbs free energies (G), energy barriers (Eb), enthalpies (H), and reaction energies (Er) in the process of hydrogen generation by Pd2O. Table S5: Gibbs free energies (G), energy barriers (Eb), enthalpies (H), and reaction energies (Er) in the process of carbon monoxide formation by Pd2. Table S6. Gibbs free energies (G), energy barriers (Eb), enthalpies (H), and reaction energies (Er) in the process of carbon monoxide formation by Pd2.

Author Contributions

Conceptualization, C.D. and J.W.; writing—original draft preparation, Y.M.; writing—review and editing, Y.M. and Y.X.; methodology, X.G.; formal analysis, Z.L., K.Z., and P.L.

Funding

This research was funded by the Natural Science Foundation of China, grant number 21706178; the Natural Science Foundation of Shanxi Province, grant number 201801D121058.

Acknowledgments

The authors gratefully acknowledge the great support and guidance of the teachers of Group 602 from the Institute of Coal Chemistry, Chinese Academy of Sciences.

Conflicts of Interest

The authors declare no conflicts of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to publish the results.

References

  1. Yang, M. Climate change and energy policies, coal and coalmine methane in China. Energy Policy 2009, 37, 2858–2869. [Google Scholar] [CrossRef]
  2. Cargnello, M.; Delgado, J.; Jaen, J.; Hernandez, J.; Garrido, C.; Bakhmutsky, K.; Montini, T.; Calvino, J.; Gamez, J.; Gorte, R.J.; et al. Exceptional activity for methane combustion over modular Pd@CeO2 subunits on functionalized Al2O3. Science 2012, 337, 713–717. [Google Scholar] [CrossRef] [PubMed]
  3. Ruckenstein, E.; Hu, Y.H. Carbon dioxide reforming of methane over nickel alkaline earth metal oxide catalysts. Appl. Catal. A Gen. 1995, 133, 149–161. [Google Scholar] [CrossRef]
  4. Lee, J.H.; Lee, E.G.; Joo, O.S.; Jung, K.D. Stabilization of Ni/Al2O3 catalyst by Cu addition for CO2 reforming of methane. Appl. Catal. A Gen. 2004, 269, 1–6. [Google Scholar] [CrossRef]
  5. Stagg, S.M.; Resasco, D.E. Effects of Promoters and Supports on Coke Formation on Pt Catalysts During CH4 Reforming with CO2. In Studies in Surface Science and Catalysis; Elsevier: Amsterdam, The Nertherlands, 1997; pp. 543–550. [Google Scholar]
  6. Li, X.; Li, S.; Yang, Y.; Wu, M.; He, F. Studies on coke formation and coke species of nickel-based catalysts in CO2 reforming of CH4. Catal. Lett. 2007, 118, 59–63. [Google Scholar]
  7. Zhao, X.; Liu, Y.; Deng, J.; Xu, P.; Yang, J.; Zhang, K.; Han, Z.; Dai, H. Mesoporous Pd Pt alloys: High-performance catalysts for methane combustion. Mol. Catal. 2017, 442, 191–201. [Google Scholar] [CrossRef]
  8. Long, H.; Xu, Y.; Zhang, X.; Hu, S.; Shang, S.; Yin, Y.; Dai, X. Ni-Co/Mg-Al catalyst derived from hydrotalcite-like compound prepared by plasma for dry reforming of methane. J. Energy Chem. 2013, 22, 733–739. [Google Scholar] [CrossRef]
  9. Qiu, Y.; Chen, J.; Zhang, J. Effect of CeO2 and CaO Promoters on Ignition Performance for Partial Oxidation of Methane over Ni/MgO-Al2O3 Catalyst. J. Nat. Gas Chem. 2007, 16, 148–154. [Google Scholar] [CrossRef]
  10. Wang, Y.; Zeng, H.; Shi, Y.; Cai, N. Methane partial oxidation in a two-layer porous media burner with Al2O3 pellets of different diameters. Fuel 2018, 217, 45–50. [Google Scholar] [CrossRef]
  11. Nilsson, J.; Carlsson, P.A.; Martin, N.M.; Velin, P.; Meira, D.M.; Grönbeck, H.; Skoglundh, M. Oxygen step-response experiments for methane oxidation over Pd/Al2O3: An in situ XAFS study. Catal. Commun. 2018, 109, 24–27. [Google Scholar] [CrossRef]
  12. Zheng, C.; Cheng, L.; Cen, K.; Bingue, J.P.; Saveliev, A. Partial oxidation of methane in a reciprocal flow porous burner with an external heat source. Int. J. Hydrog. Energy 2012, 37, 4119–4126. [Google Scholar] [CrossRef]
  13. Guo, C.; Zhang, X.; Zhang, J.; Wang, Y. Preparation of La2NiO4 catalyst and catalytic performance for partial oxidation of methane. J. Mol. Catal. A Chem. 2007, 269, 254–259. [Google Scholar] [CrossRef]
  14. Xunli, C.; Zhang, S.M.L.; Michael, D.; Mingos, P.; David Hayward, O. Oscillatory behaviour during the oxidation of methane over palladium metal catalysts. Appl. Catal. A Gen. 2003, 240, 183–197. [Google Scholar] [Green Version]
  15. Nilsson, J.; Carlsson, P.A.; Fouladvand, S.; Martin, N.M.; Gustafson, J.; Newton, M.A.; Lundgren, E.; Grönbeck, H.; Skoglundh, M. Chemistry of Supported Palladium Nanoparticles during Methane Oxidation. ACS Catal. 2015, 5, 2481–2489. [Google Scholar] [CrossRef] [Green Version]
  16. Maxim, L.; Lyubovsky, P. Complete methane oxidation over Pd catalyst supported on a-alumina. In¯uence of temperature and oxygen pressure on the catalyst activity. Catal. Today 1999, 47, 29–44. [Google Scholar]
  17. Hicks, R.F.; Qi, H.; Young, M.L.; Lee, R.G. Structure Sensitivity of Methane Oxidation over Platinum and Palladium. J. Catal. 1990, 122, 280–294. [Google Scholar] [CrossRef]
  18. Gabasch, H. Methane Oxidation on Pd(111): In Situ XPS Identification of Active Phase. J. Phys. Chem. C 2007, 111, 7957–7962. [Google Scholar] [CrossRef]
  19. Niu, F.; Li, S.; Zong, Y.; Yao, Q. Catalytic Behavior of Flame-Made Pd/TiO2 Nanoparticles in Methane Oxidation at Low Temperatures. J. Phys. Chem. C 2014, 118, 19165–19171. [Google Scholar] [CrossRef]
  20. Li, B.; Li, H.; Weng, W.Z.; Zhang, Q.; Huang, C.J.; Wan, H.L. Synthesis gas production from partial oxidation of methane over highly dispersed Pd/SiO2 catalyst. Fuel 2013, 103, 1032–1038. [Google Scholar] [CrossRef]
  21. Xiong, H.; Lester, K.; Ressler, T.; Schlögl, R.; Allard, L.F.; Datye, A.K. Metastable Pd ↔ PdO Structures During High Temperature Methane Oxidation. Catal. Lett. 2017, 147, 1095–1103. [Google Scholar] [CrossRef]
  22. Lan, L.; Chen, S.; Zhao, M.; Gong, M.; Chen, Y. The effect of synthesis method on the properties and catalytic performance of Pd/Ce0.5Zr0.5O2-Al2O3 three-way catalyst. J. Mol. Catal. A Chem. 2014, 394, 10–21. [Google Scholar] [CrossRef]
  23. Lovón-Quintana, J.J.; Santos, J.B.; Lovón, A.S.; La-Salvia, N.; Valença, G.P. Low-temperature oxidation of methane on Pd-Sn/ZrO2 catalysts. J. Mol. Catal. A Chem. 2016, 411, 117–127. [Google Scholar] [CrossRef]
  24. Yang, Y. Structural and dynamical properties of water adsorption on PtO2(001). RSC Adv. 2018, 8, 15078–15086. [Google Scholar] [CrossRef]
  25. Yoo, J.S.; Schumann, J.; Studt, F.; Abild-Pedersen, F.; Nørskov, J.K. Theoretical Investigation of Methane Oxidation on Pd(111) and Other Metallic Surfaces. J. Phys. Chem. C 2018, 122, 16023–16032. [Google Scholar] [CrossRef]
  26. Jiang, Z.; Wang, B.; Fang, T. Adsorption and dehydrogenation mechanism of methane on clean and oxygen-covered Pd(1 0 0) surfaces: A DFT study. Appl. Surf. Sci. 2014, 320, 256–262. [Google Scholar] [CrossRef]
  27. Dianat, A.; Seriani, N.; Ciacchi, L.C.; Bobeth, M.; Cuniberti, G. DFT study of reaction processes of methane combustion on PdO(100). Chem. Phys. 2014, 443, 53–60. [Google Scholar] [CrossRef]
  28. Wang, Z.; Du, W.; Wang, J. Research and Application of improved Adaptive MOMEDA Fault Diagnosis Method. Measurement. 2019, 140, 63–75. [Google Scholar] [CrossRef]
  29. Geng, C.; Weiske, T.; Li, J.; Shaik, S.; Schwarz, H. On the Intrinsic Reactivity of Diatomic 3d Transition-Metal Carbides in the Thermal Activation of Methane: Striking Electronic Structure Effects. J. Am. Chem. Soc. 2018, 141, 599–610. [Google Scholar] [CrossRef]
  30. Qi, W.; Ran, J.; Wang, R.; Du, X.; Shi, J.; Ran, M. Kinetic mechanism of effects of hydrogen addition on methane catalytic combustion over Pt(1 1 1) surface: A DFT study with cluster modeling. Comput. Mater. Sci. 2016, 111, 430–442. [Google Scholar] [CrossRef]
  31. Zhao, H.; Wu, P.; Du, L.; Du, H. Effect of the nanopore on ferroelectric domain structures and switching properties. Comput. Mater. Sci. 2018, 148, 216–223. [Google Scholar] [CrossRef]
  32. Gong, H.; Lu, W.; Wang, L.; Li, G.; Zhang, S. The effect of deposition velocity and cluster size on thin film growth by Cu cluster deposition. Comput. Mater. Sci. 2012, 65, 230–234. [Google Scholar] [CrossRef]
  33. Trevor, D.J.; Cox, D.M.; Kaldor, A. Methane Activation on Unsupported Platinum Clusters. J. Am. Chem. Soc. 1990, 90, 3742–3749. [Google Scholar] [CrossRef]
  34. Lang, S.M.; Bernhardt, T.M. Methane activation and partial oxidation on free gold and palladium clusters: Mechanistic insights into cooperative and highly selective cluster catalysis. Faraday Discuss. 2011, 152, 337–351. [Google Scholar] [CrossRef]
  35. Nasresfahani, S.; Safaiee, R.; Sheikhi, M.H. Influence of Pd/Pd2 decoration on the structural, electronic and sensing properties of monolayer graphene in the presence of methane molecule: A dispersion-corrected DFT study. Surf. Sci. 2017, 662, 93–101. [Google Scholar] [CrossRef]
  36. Wang, Z.; Zheng, L.; Cai, W.; Du., W. A Novel Method for Intelligent Fault Diagnosis of Bearing Based on Capsule Neural Network. Complexity 2019, 7, 1–17. [Google Scholar] [CrossRef]
  37. Guo, J.; Xie, C.; Lee, K.; Guo, N.; Miller, J.T.; Janik, M.J.; Song, C. Improving the Carbon Resistance of Ni-Based Steam Reforming Catalyst by Alloying with Rh: A Computational Study Coupled with Reforming Experiments and EXAFS Characterization. ACS Catal. 2011, 1, 574–582. [Google Scholar] [CrossRef]
  38. Liu, H.; Teng, B.; Fan, M.; Wang, B.; Zhang, Y.; Harris, H.G. CH4 dissociation on the perfect and defective MgO(001) supported Ni4. Fuel 2014, 123, 285–292. [Google Scholar] [CrossRef]
  39. Qi, Q.; Wang, X.; Chen, L.; Li, B. Methane dissociation on Pt(111), Ir(111) and PtIr(111) surface: A density functional theory study. Appl. Surf. Sci. 2013, 284, 784–791. [Google Scholar] [CrossRef]
  40. Wang, R.; Ran, J.; Qi, W.; Niu, J.; Du, X. A comparison of methane activation on catalysts Pt 2 and PtNi. Comput. Theor. Chem. 2015, 1073, 94–101. [Google Scholar] [CrossRef]
  41. Li, K.; Zhou, Z.; Wang, Y.; Wu, Z. A theoretical study of CH4 dissociation on NiPd(111) surface. Surf. Sci. 2013, 612, 63–68. [Google Scholar] [CrossRef]
  42. Kinnunen, N.M.; Hirvi, J.T.; Suvanto, M.; Pakkanen, T.A. Role of the Interface between Pd and PdO in Methane Dissociation. J. Phys. Chem. C 2011, 115, 19197–19202. [Google Scholar] [CrossRef]
  43. Mayernick, A.D.; Janik, M.J. Methane oxidation on Pd–Ceria: A DFT study of the mechanism over PdxCe1−xO2, Pd, and PdO. J. Catal. 2011, 278, 16–25. [Google Scholar] [CrossRef]
  44. Yang, J.; Miao, J.; Li, X.; Xu, W. Density functional theory studies on the mechanism of activation of methane by homonuclear bimetallic Ni–Ni. Comput. Theor. Chem. 2012, 996, 117–124. [Google Scholar] [CrossRef]
  45. Zhang, M.; Yang, K.; Zhang, X.; Yu, Y. Effect of Ni(111) surface alloying by Pt on partial oxidation of methane to syngas: A DFT study. Surf. Sci. 2014, 630, 236–243. [Google Scholar] [CrossRef]
  46. Niu, J.; Ran, J.; Wang, R.; Du, X. Mechanism of methylene oxidation on Pt catalysts: A DFT study. Comput. Theor. Chem. 2015, 1067, 40–47. [Google Scholar] [CrossRef]
Figure 1. Structures of dimer Pd2 and trimer Pd2O (bond unit, Å; turquoise: Pd, red: O).
Figure 1. Structures of dimer Pd2 and trimer Pd2O (bond unit, Å; turquoise: Pd, red: O).
Catalysts 09 00666 g001
Figure 2. Optimized configurations of the CHx–Pd2(o) (x = 0–4) system (bond unit, Å; turquoise: Pd, gray: C, white: H, red: O).
Figure 2. Optimized configurations of the CHx–Pd2(o) (x = 0–4) system (bond unit, Å; turquoise: Pd, gray: C, white: H, red: O).
Catalysts 09 00666 g002
Figure 3. Partial density of states (PDOS) of CHx (x =1–4) before and after being catalyzed by Pd2 and Pd2O. (ad) CH4, CH3, CH2, and CH, respectively, before and after being catalyzed by Pd2 and Pd2O. Black, red, and blue lines, free CHx and CHx–Pd2(O), respectively. The blue dashed vertical lines indicate the Fermi levels.
Figure 3. Partial density of states (PDOS) of CHx (x =1–4) before and after being catalyzed by Pd2 and Pd2O. (ad) CH4, CH3, CH2, and CH, respectively, before and after being catalyzed by Pd2 and Pd2O. Black, red, and blue lines, free CHx and CHx–Pd2(O), respectively. The blue dashed vertical lines indicate the Fermi levels.
Catalysts 09 00666 g003
Figure 4. Optimized structure of reactants, intermediates, products, and the transition state of methane activation by Pd2 (turquoise: Pd, gray: C, white: H).
Figure 4. Optimized structure of reactants, intermediates, products, and the transition state of methane activation by Pd2 (turquoise: Pd, gray: C, white: H).
Catalysts 09 00666 g004
Figure 5. Optimized structure of reactants, intermediates, products, and the transition state of methane activation by Pd2O (turquoise: Pd, gray: C, white: H).
Figure 5. Optimized structure of reactants, intermediates, products, and the transition state of methane activation by Pd2O (turquoise: Pd, gray: C, white: H).
Catalysts 09 00666 g005
Figure 6. Energy profiles of dehydrogenation from CH4 to C: CH4→CH3 +H→CH2 + 2H→CH + 3H→C + 4H on Pd2 and Pd2O. Black line and red line represent the dissociation of CHx on Pd2 and Pd2O.
Figure 6. Energy profiles of dehydrogenation from CH4 to C: CH4→CH3 +H→CH2 + 2H→CH + 3H→C + 4H on Pd2 and Pd2O. Black line and red line represent the dissociation of CHx on Pd2 and Pd2O.
Catalysts 09 00666 g006
Figure 7. Optimized structure of reactants, intermediates, products, and the transition state of hydrogen generation by Pd2 (turquoise: Pd, gray: C, white: H).
Figure 7. Optimized structure of reactants, intermediates, products, and the transition state of hydrogen generation by Pd2 (turquoise: Pd, gray: C, white: H).
Catalysts 09 00666 g007
Figure 8. Optimized structure of reactants, intermediates, products, and the transition state of hydrogen generation by Pd2O (turquoise: Pd, gray: C, white: H).
Figure 8. Optimized structure of reactants, intermediates, products, and the transition state of hydrogen generation by Pd2O (turquoise: Pd, gray: C, white: H).
Catalysts 09 00666 g008
Figure 9. Energy profiles of hydrogen generation: CH2 + 2H→CH2 + H2 on Pd2 and Pd2O.
Figure 9. Energy profiles of hydrogen generation: CH2 + 2H→CH2 + H2 on Pd2 and Pd2O.
Catalysts 09 00666 g009
Figure 10. Energy profiles of carbon monoxide formation on Pd2 and Pd2O. Left curve, three reaction paths on Pd2; right curve, three reaction paths on Pd2O.
Figure 10. Energy profiles of carbon monoxide formation on Pd2 and Pd2O. Left curve, three reaction paths on Pd2; right curve, three reaction paths on Pd2O.
Catalysts 09 00666 g010
Table 1. Binding energies of CHx (x = 0–4) adsorbates on dimer Pd2 and trimer Pd2O.
Table 1. Binding energies of CHx (x = 0–4) adsorbates on dimer Pd2 and trimer Pd2O.
SystemCoordination Energy Pd2 (eV) Pd2O (eV)
CH4–Pd2(o)−0.32−0.87
CH3–Pd2(o)−2.63−2.42
CH2–Pd2(o)−4.90−4.08
CH–Pd2(o)−5.70−4.66
C–Pd2(o)−6.01−5.30
Table 2. The activation energy (Ea), difference enthalpy between the transition state and the initial state (ΔH), difference entropies between the transition state and the initial state (ΔS), pre-exponential factor (A), rate constant (κ) for methane activation on Pd2.
Table 2. The activation energy (Ea), difference enthalpy between the transition state and the initial state (ΔH), difference entropies between the transition state and the initial state (ΔS), pre-exponential factor (A), rate constant (κ) for methane activation on Pd2.
Pd2Ea (eV)ΔH (eV)ΔS (eV)A (S−1)κ (S−1)
CH4→CH3 + H1.171.1−0.211.62 × 10124.36 × 105
CH3 + H→CH2 + 2H1.971.9−0.182.10 × 10129.84 × 100
CH2 + 2H→CH + 3H4.184.1−0.177.44 × 10127.27 × 10–13
CH + 3H→C + 4H3.012.93−0.157.04 × 10139.30 × 10–6
Table 3. The activation energy (Ea), difference enthalpy between the transition state and the initial state (ΔH), difference entropies between the transition state and the initial state (ΔS), pre-exponential factor (A), rate constant (κ) for methane activation on Pd2O.
Table 3. The activation energy (Ea), difference enthalpy between the transition state and the initial state (ΔH), difference entropies between the transition state and the initial state (ΔS), pre-exponential factor (A), rate constant (κ) for methane activation on Pd2O.
Pd2OEa (eV)ΔH (eV)ΔS (eV)A (S–1)κ (S−1)
CH4→CH3 + H1.921.85−0.324.07 × 10113.75 × 100
CH3 + H→CH2 + 2H 1.941.87−0.421.11 × 10117.75 × 10–1
CH2 + 2H→CH + 3H3.473.4−0.072.31 × 10124.03 × 10–8
CH + 3H→C + 4H3.133.060.113.04 × 10123.74 × 10–5
Table 4. The activation energy (Ea), difference enthalpy between the transition state and the initial state (ΔH), difference entropies between the transition state and the initial state (ΔS), pre-exponential factor (A), rate constant (κ) for methane activation on Pd2.
Table 4. The activation energy (Ea), difference enthalpy between the transition state and the initial state (ΔH), difference entropies between the transition state and the initial state (ΔS), pre-exponential factor (A), rate constant (κ) for methane activation on Pd2.
Pd2Ea (eV)ΔH (eV)ΔS (eV)A (S–1)κ (S−1)
CH2 + 2H→CH2 + 2HS0.03−0.04−0.211.43 × 10122.44 × 1012
CH2 + 2HS→CH2 + 2HN0.680.61−0.069.11 × 10122.20 × 109
CH2 + 2HN→CH2 +H21.111.04−0.078.16 × 10125.27 × 106
Table 5. The activation energy (Ea), difference enthalpy between the transition state and the initial state (ΔH), difference entropies between the transition state and the initial state (ΔS), pre-exponential factor (A), rate constant (κ) for methane activation on Pd2O.
Table 5. The activation energy (Ea), difference enthalpy between the transition state and the initial state (ΔH), difference entropies between the transition state and the initial state (ΔS), pre-exponential factor (A), rate constant (κ) for methane activation on Pd2O.
Pd2OEa (eV)ΔH (eV)ΔS (eV)A (S–1)κ (S–1)
CH2 + 2H→CH2 + 2HS0.230.15−0.069.38 × 10121.13 × 1012
CH2 + 2HS→CH2 + 2HN0.480.090.095.21 × 10131.80 × 1011
CH2 + 2HN→CH2 +H21.08−0.08−0.086.57 × 10125.77 × 106
Table 6. The activation energy (Ea), difference enthalpy between the transition state and the initial state (ΔH), difference entropies between the transition state and the initial state (ΔS), pre-exponential factor (A), rate constant (κ) for methane activation on Pd2.
Table 6. The activation energy (Ea), difference enthalpy between the transition state and the initial state (ΔH), difference entropies between the transition state and the initial state (ΔS), pre-exponential factor (A), rate constant (κ) for methane activation on Pd2.
Pd2Ea(V)ΔH (eV)ΔS (eV)A (S−1)κ (S−1)
CH→C + H2.752.680.011.97 × 10–31.95 × 1013
C + O→CO1.51.42−0.195.82×1031.81 × 1012
CH + O→CHO1.61.530.031.87 × 1042.61 × 1013
CHO→CO + H0.530.45−0.184.25 × 1042.15 × 1012
CH + O→COH1.641.56−0.112.19 × 1034.81 × 1012
COH→CO + H1.731.650.054.42 × 1033.31 × 1013
Table 7. The activation energy (Ea), difference enthalpy between the transition state and the initial state (ΔH), difference entropies between the transition state and the initial state (ΔS), pre-exponential factor (A), rate constant (κ) for methane activation on Pd2O.
Table 7. The activation energy (Ea), difference enthalpy between the transition state and the initial state (ΔH), difference entropies between the transition state and the initial state (ΔS), pre-exponential factor (A), rate constant (κ) for methane activation on Pd2O.
Pd2OEa (eV)ΔH (eV)ΔS (eV)A (S–1)κ (S–1)
CH→C + H1.931.85−0.076.09 × 1017.59 × 1012
C + O→CO1.231.16−0.314.69 × 1043.98 × 1011
CH + O→CHO0.580.51−0.114.28 × 1094.83 × 1012
CHO→CO + H1.191.120.035.84 × 1062.68 × 1013
CH + O→COH0.620.54−0.151.75 × 1093.07 × 1012
COH→CO + H1.161.08−0.101.80 × 1065.32 × 1013

Share and Cite

MDPI and ACS Style

Meng, Y.; Xue, Y.; Ding, C.; Gao, X.; Zhang, K.; Liu, P.; Wang, J.; Li, Z. Oxygen Atom Function: The Case of Methane Oxidation Mechanism to Synthesis Gas over a Pd Cluster. Catalysts 2019, 9, 666. https://doi.org/10.3390/catal9080666

AMA Style

Meng Y, Xue Y, Ding C, Gao X, Zhang K, Liu P, Wang J, Li Z. Oxygen Atom Function: The Case of Methane Oxidation Mechanism to Synthesis Gas over a Pd Cluster. Catalysts. 2019; 9(8):666. https://doi.org/10.3390/catal9080666

Chicago/Turabian Style

Meng, Yuanyuan, Yuyuan Xue, Chuanmin Ding, Xiaofeng Gao, Kan Zhang, Ping Liu, Junwen Wang, and Zhe Li. 2019. "Oxygen Atom Function: The Case of Methane Oxidation Mechanism to Synthesis Gas over a Pd Cluster" Catalysts 9, no. 8: 666. https://doi.org/10.3390/catal9080666

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