Next Article in Journal
Effect of Water Molecule on Photo-Assisted Nitrous Oxide Decomposition over Oxotitanium Porphyrin: A Theoretical Study
Next Article in Special Issue
Bioinspired Design and Computational Prediction of SCS Nickel Pincer Complexes for Hydrogenation of Carbon Dioxide
Previous Article in Journal
Hydrogenation of Carbon Dioxide on Supported Rh Catalysts
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Oxygen Reduction Reaction Catalyzed by Pt3M (M = 3d Transition Metals) Supported on O-doped Graphene

1
Key Laboratory for Green Chemical Technology of the Ministry of Education, School of Chemical Engineering and Technology, Collaborative Innovation Center of Chemical Science and Engineering, Tianjin University, Tianjin 300072, China
2
Department of Chemistry and Biochemistry, Southern Illinois University, Carbondale, IL 62901, USA
3
State Key Laboratory for Structural Chemistry of Unstable and Stable Species, Institute of Chemistry, Chinese Academy of Sciences, Beijing 100190, China
*
Authors to whom correspondence should be addressed.
Catalysts 2020, 10(2), 156; https://doi.org/10.3390/catal10020156
Submission received: 27 December 2019 / Revised: 15 January 2020 / Accepted: 16 January 2020 / Published: 1 February 2020

Abstract

:
Pt3M (M = 3d transition metals) supported on oxygen-doped graphene as an electrocatalyst for oxygen reduction was investigated using the periodic density functional theory-based computational method. The results show that oxygen prefers to adsorb on supported Pt3M in a bridging di-oxygen configuration. Upon reduction, the O–O bond breaks spontaneously and the oxygen adatom next to the metal–graphene interface is hydrogenated, resulting in co-adsorbed O* and OH* species. Water formation was found to be the potential-limiting step on all catalysts. The activity for the oxygen reduction reaction was evaluated against the difference of the oxygen adsorption energy on the Pt site and the M site of Pt3M and the results indicate that the oxygen adsorption energy difference offers an improved prediction of the oxygen reduction activity on these catalysts. Based on the analysis, Pt3Ni supported on oxygen-doped graphene exhibits an enhanced catalytic performance for oxygen reduction over Pt4.

Graphical Abstract

1. Introduction

Proton-exchange membrane fuel cells (PEMFCs) can directly convert the chemical energy stored in hydrogen and oxygen into electricity [1,2,3], but their adoption for practical application as electrocatalysts is hindered by the high price and limited supply of platinum [4,5,6]. Efforts have been made to reduce, and even replace, platinum in electrocatalysts. Alloying platinum with inexpensive metals has been shown to maintain or improve the activity for the oxygen reduction reaction (ORR) while significantly reducing the cost [7,8,9,10,11,12]. Among inexpensive metals, 3d metals, including Ti [13,14], V [13,14], Cr [14], Mn [14], Fe [13,14,15,16], Co [13,14,15,17,18], Ni [13,14,15,19], and Cu [17,20,21] have been tested as a component of the electrocatalysts for ORR.
Graphitic carbon materials have long been used as electrode materials, with graphene-based materials being natural alternatives due largely to their generally high electrical conductivity, high specific surface areas, and strong chemical strength [22,23]. Therefore, graphene-supported Pt and metal alloy catalysts have been studied for the oxygen reduction process [24,25,26]. B-, N-, and S-doped graphene have also been investigated for ORR [27,28,29,30,31]. Various functional groups on graphene have been reported to enhance the ORR activity [32]. In particular, graphene oxides and oxygen-containing graphene materials are widely available. These oxygen-derived defect sites in graphene may form the anchor sites for the active catalyst and/or work synergistically with the active components to promote the reactions. Consequently, the use of graphene oxides either alone or as a support in the electrode has attracted extensive attention [6,33,34,35]. In the present study, we use O-doped graphene to model the oxygen-containing graphene materials.
Our previous study on the ORR process over Pt4, Pt3V, and Pt3Fe supported on O-doped graphene showed that V and Fe change the ORR mechanism and activity on the Pt-based catalysts differently [36]. Other studies also reported that early and late 3d transition metals could affect the adsorption of O2 and the ORR activity in a different way [13,37]. Herein, we examined the entire row of 3d transition metals in the form of Pt3M with M = Sc, Ti, V, Cr, Mn, Fe, Co, and Ni on O-doped graphene (OG) as the ORR catalysts. We compared the effect of transition metals on the ORR mechanism and activity. Through a detailed analysis of O2, OH, and H2O adsorption energies, as well as the ORR mechanism, we identified Ni as the most effective modifier.

2. Results and Discussion

2.1. Optimized Pt3M Structures on O-doped Graphene

Our previous study showed that the O-doping site is the anchoring site for the tetrahedral Pt4 cluster [36]. The OG-supported Pt3M clusters (OG-Pt3M) were constructed by replacing one Pt atom of the supported Pt4 cluster with M at either the interfacial or the top of tetrahedron. Structural optimization showed that the interfacial substituted Pt3M cluster is more stable, which was used for the ORR study. In this work, we expanded M to include all 3d metals, i.e., M = Sc, Ti, V, Cr, Mn, Fe, Co, and Ni. As we showed previously, the doped O atom is 2-C coordinated and 0.72 Å above the plane of the graphene sheet on pure OG [38]. As shown in Table 1, the doped O atom was pushed to the opposite side of the graphene sheet at a distance of 0.17~0.60 Å, depending on M in the supported cluster. The details of the optimized structures are shown in Table S1 of the Supporting Information.
Table 1 lists the Bader charges of C1, C2, O, Pt1, Pt2, Pt3, and M (or Pt0) in the optimized Pt3M on OG, with numbering of the atoms shown in the figures under Table 1. The charges of the doped O atom and the two linking C atoms were generally maintained at similar values for different Pt3M clusters. The exception was the C2 atom in Pt3Co and Pt3Ni due to the O–C2 bond breaking and the O–C3 bond formation. In supported Pt4, two Pt atoms were positively charged, whereas the other two were negatively charged, although the charge differences were not great. When a second metal was introduced to replace one of the Pt atoms to form Pt3M, charge localization occurred. In particular, the top Pt1 atom was significantly more negatively charged, whereas the M atom was positively charged and the charge on M decreased from 2.13 |e| on Ti to 0.57 |e| on Ni, due mainly to the change of the electronic structure and atomic radius. We note that the charge on the more active early transition metals (M for M = Sc, Ti, V, and Cr) was much larger than those on the late transition metals (M = Mn, Fe, Co, and Ni). These differences in charge redistribution will in turn affect the oxygen adsorption energies and ORR reactivity.
To characterize the stability of supported Pt3M on OG, we defined a net binding energy (Eb) as ∆Eb = Eb(Pt4) − Eb(Pt3M), i.e., the binding energy difference of OG–Pt3M and OG–Pt4, and plotted the results in Figure 1. According to this definition, a positive value would indicate that the Pt3M cluster binds OG stronger than Pt4. As shown in Figure 1, the binding energy of Pt3Mn and Pt3Fe was closer to Pt4, while the binding energy difference was highly negative for Pt3Ti and Pt3V, indicating that these two clusters bind OG much weaker than Pt4. On the other hand, Pt3M with M = Sc, Cr, Co, and Ni binds OG more strongly than Pt4, by ~0.1 eV, indicating their higher stabilities. The relative stability of OG-Pt3M is a reflection of the interaction between Pt3M and support and will affect the ORR performance.

2.2. Oxygen Adsorption and Reduction Mechanism

The mechanism of ORR on various electrocatalysts was extensively investigated using the density functional theory (DFT) method [4,27,39,40,41,42]. Previously, we studied the ORR mechanism over the OG supported Pt4, Pt3V, and Pt3Fe clusters and reported two possible pathways [36]. Herein, we expanded the investigation to include all 3d transition metals.
The first step for oxygen reduction is O2 adsorption. We calculated the possible adsorption structures of O2 on OG–Pt3M and found that O2 prefers to bind the supported cluster in a di-oxygen form in a bridging configuration, with one oxygen atom on Pt1 and another on M, as can be seen in Figure 2 and Figure S1. The O–O distance of these adsorbed di-oxygen species was stretched to 1.41~1.47 Å from 1.23 Å of the isolated O2 molecule. Interestingly, the Pt3M cluster maintained a tetrahedral structure after being combined with O2, in contrast to Pt4, which was transformed into a planar-ring structure by O2 adsorption, indicating that the transition metal atoms in Pt-based alloys contribute to stabilizing the cluster structure.
In addition to the di-oxygen form of adsorption, dissociative adsorption of O2 occurred on OG–Pt3Ti and OG–Pt3V, resulting in two O adatoms with O–O distances of 4.35 Å and 4.32 Å, respectively (shown in Figure 2). The dissociative adsorption was more stable than the di-oxygen state, by 1.61 eV for Pt3Ti and 2.65 eV for Pt3V. Based on Bader charge analysis, the Ti and V atoms in clusters had the highest positive charges among the 3d transition metals, with values of 2.13 |e| and 1.70 |e|, respectively. The oxygen atoms bound to the Ti and V atoms were more negatively charged than the oxygen adatom on Pt1, indicating that the oxygen atoms in the adsorbed di-oxygen state were strongly polarized by the large difference of electron distribution between Ti/V and Pt1. As a result, the dissociative adsorption of oxygen was more stable than the di-oxygen state on Pt3Ti and Pt3V, which also hindered the following H2O formation and release. In contrast, the charge difference between the two adsorbed oxygen atoms in the di-oxygen form on other Pt3M clusters fell in a small range of 0.07~0.17 |e|, and consequently, the adsorbed di-oxygen species was more stable. In order to compare the ORR activity on these supported Pt3M, we chose the di-oxygen species as the reference.
The reduction of adsorbed di-oxygen species occurs by adding a hydrogen atom either on the top O atom on Pt1 or the interfacial O atom on M. On Pt(111), ORR was believed to proceed via the OOH* intermediate [39,40,43]. However, a stable OOH* intermediate was not isolated on any of the OG–Pt3M catalysts. Upon hydrogenation, the di-oxygen species dissociates into co-adsorbed O* and OH* species on all Pt3M. For the early transition metals (M = Ti, V, and Cr), protonation of the top O of the di-oxygen species leads to more stable OH(t)* and O*, while for the late transition metals (M = Mn, Fe, Co, and Ni), the formation of OH(b)* with the interfacial O atom and leaving the top O as O* is more stable (suffix “t” and “b” correspond to the species on the top Pt1 atom and the interfacial M (or Pt0) atom, respectively).
Further reduction after either OH(t)* or OH(b)* formation may follow two pathways, as shown in Scheme 1. The calculated reaction free energy for every elementary step indicates that protonating two O*s to form two OH*s is more preferable than hydrogenating one O* completely to H2O* before hydrogenating the other O*. Further reducing the OH* species will produce H2O*, which will desorb and liberate the active sites for the continued O2 adsorption and reduction (see Tables S2 and S3 for details). The step with the most positive reaction free energy is the hydrogenation of the OH* species, which is likely to be the potential-limiting step. Between OH(t)* and OH(b)*, the reduction of OH(b)* has a lower reaction free energy and is more favorable than reducing the OH(t)*. In Scheme 1, the most favorable pathway is shown in bold black.
To compare the reaction on different OG–Pt3M catalysts, the free energy profiles for ORR through the interfacial site at 0 and 0.85 V (SHE) were constructed on the basis of the DFT results (see details in Figures S2 and S3). Compared with the results on OG–Pt4, the adsorption of O2 was strengthened on Pt3M with M = Sc, Ti, V, and Cr. On Pt3M with M = Mn, Fe, Co, and Ni, the O2 adsorption energy became 0.1~0.4 eV smaller than that on Pt4. These results are consistent with previous studies [37,44]. Figure 3 shows a comparison of the free energy profiles of ORR on OG–Pt4, OG–Pt3Ti, and OG–Pt3Ni. Obviously, oxygen bound much stronger on Pt3Ti than on Pt4: the binding energy on the supported Pt3Ti was increased by 0.71 eV, resulting in a highly stable OH* species, which is difficult to hydrogenate to H2O. In the case of Pt3Ni, the adsorption energy of oxygen was decreased by 0.4 eV from that of Pt4. By applying a potential of 0.85 V (SHE), the reduction of OH* to form H2O* on Pt3Ni became less endothermic than Pt4. In addition, the desorption of H2O* became easier than that on Pt4, indicating an enhanced ORR activity on Pt3Ni.

2.3. Correlation of the ORR Activity with the Oxygen Binding Energy Difference

To design an electrocatalyst with improved ORR activity, understanding the relationship between the activity and atomic/electronic structures of the catalyst is important [1,37,45]. On transition metal- and/or metal alloy-based catalysts, the linear relationship between the ORR activity and the d-band center of those electrocatalysts has been established [13,20,37,46,47]. The oxygen binding energy (ΔEO) has been used as a descriptor for the ORR activity on the transition metal and even on non-metallic surfaces [1,39,48,49]. In fact, the ΔEO descriptor can be considered an extension and application of the d-band center model in electrocatalysis, as the adsorption/binding energy of oxygen has been shown to correlate with the d states of the surface atoms to which the atom/molecule binds [13,37]. Generally, an upward shift of the d-band center to Fermi level results in a stronger binding [13,17,24]. However, both the d-band center model and the oxygen-binding energy descriptor have limitations [14,50,51]. For example, Lin et al. reported that the ORR activity of the Pt-cored alloy (Pt@M, M = Co, Cu, Pd, and Au) catalysts does not correlate well with the d-band center of surface Pt [17]. Yu et al. showed that the binding energies of O do not always correlate with those of OH [52].
To understand the performance difference of the Pt3M alloy catalysts in the present study, we plotted the average adsorption energy of di-oxygen species (1/2 ΔEO2) on the OG–Pt3M as a function of the d-band center (εd) of the cluster atoms (as shown in Figure S4) and found a very weak linear correlation. In fact, the present results indicate that a higher d-band center in the clusters involving late transition metals (M = Mn, Fe, Co, and Ni) corresponds to a lower oxygen adsorption energy. Previous reports showed that the adsorption energy of O* had a strong correlation with the adsorption energy of OH* [53,54,55]. Herein, we evaluated the possible correlation between the adsorption energies of O* and OH* on the interfacial sites of the supported Pt3M (seen in Figure S5) and also found a rather weak linear correlation, demonstrating that neither ΔEO nor εd is a good descriptor of the ORR activity of the OG–Pt3M.
Charge redistribution is expected to occur upon formation of the O* and OH* species. We analyzed the Bader charge of the interfacial M sites (q(M)) upon forming the O* and OH* adsorbates, as well as the those of the OH* and O* (q(O)) on different alloy clusters. As shown in Figure 4a, there was quite a good linear correlation between q(M) and the charges on OHb* and Ob*. The slope of the lines is ~1.06, indicating a similar electron transfer from M to OHb* and Ob*. In the 3d transition metal series, the late transition metals (including Ni, Co, Fe, and Mn) transferred fewer electronic charges to the adsorbates than the early transition metals and resulted in a relatively weaker binding of O*.
From Figure 4b, the positive charges on the interfacial M site generally increased by following the order of the 3d elements in the periodic table from right to left, with the exception of Sc upon O* formation. In OG–Pt3Sc, the positive charge on Sc with adsorbed Ob* was even smaller than that on Cr in Pt3Cr. We also plotted the adsorption energy of Ob* (ΔE(Ob)) against q(M) of Ob* (seen in Figure S6) and found no strong correlation between the two.
Since the O2 molecule prefers to form a di-oxygen species on the bridge sites of the top Pt1 and interfacial M (or Pt0) atoms, both sites were expected to contribute to the ORR activity. Therefore, we needed a descriptor that takes into the contribution of both sites to account for the observed ORR activity. Herein, we used the adsorption energy difference between the interfacial oxygen adatom Ob* and the top oxygen adatom Ot* species (ΔEOb-Ot) to characterize the clusters and plotted the adsorption energy difference on each OG–Pt3M as a function of q(M) of Ob* in Figure 4c. As shown in the figure, there was a reasonable correlation between the two (R2 = 0.64). The correlations shown in Figure 4 are useful to understand the observed ORR activity, although they cannot be used as a descriptor to predict superior ORR catalysts.
Both the O* and OH* species are important intermediates for ORR, and their stability will have a strong influence on the overall reaction activity. However, the adsorption energy of individual OH* and O* only shows a weak linear relationship on the single interfacial site. As the adsorption energy difference of ΔEOb-Ot presents a good correlation with the charge changes of cluster, we investigated the adsorption energy difference of OH* species adsorbed on the interfacial and top sites (ΔEOHb-OHt). The correlation between ΔEOHb-OHt and ΔEOb-Ot is shown in Figure 5a. Following the increased energy difference of O*, O* adsorbed on the interfacial site became less stable, resulting in a similar trend to that of the OH* species. Compared to pure Pt4, Pt3Ni and Pt3Co showed weaker binding toward the O* and OH* at the interfacial site. All the other Pt3Ms exhibited stronger binding toward O* at the interfacial M site. Therefore, the adsorption energy difference between the top Pt site and the interfacial M site provides a good indicator of the alloying effect on the ORR reactivity.
In Figure 5b, the reaction free energy of the potential limiting step was plotted as a function of the adsorption energy difference of O* on top and interfacial sites. Again, a good correlation (R2 = 0.81) was obtained. In general, a higher positive adsorption energy difference corresponds to a lower reaction-free energy, and therefore, a higher activity. If the OG–Pt4 is chosen as a reference, both Pt3Fe and Pt3Ni have smaller reaction free energies, indicating a better ORR performance than Pt4. Among all the OG–Pt3Ms, Ni exhibited the largest decrease in the reaction free energy, and thereby, the limiting potential. Consequently, OG–Pt3Ni is the best catalyst among all OG–Pt3Ms. This conclusion is supported by the free energy profiles (Figures S2 and S3). The limiting potential for ORR on OG–Pt3Ni was reduced to 0.83 V from 1.00 V on OG–Pt4.
The present analysis demonstrated that oxygen adsorption energy alone is not effective in predicting ORR reactivity. For the alloy catalysts, the adsorption of the ORR intermediates on different metal sites should be taken into account when both metals participate in the reaction. On the OG–Pt3M catalysts, ΔEOb-Ot provides a more effective indicator of the ORR activity. In reactions involving sites of different metals, the adsorption energy difference between the alloying metals would provide a more complete description of the ORR activity.

3. Model and Computational Details

Spin unrestricted DFT calculations were performed using the Vienna ab initio simulations program (VASP), with the projector augmented wave method [56,57,58]. In short, the Perdew–Burke–Ernzerhof functional [59] and a cutoff energy of 400 eV for the plane wave basis set were used in the calculations. The same oxygen-doped graphene supercell of 12.30 × 17.04 Å with a vacuum of 20 Å as that used in our previous work was employed in the present study [36], and the same 3 × 2 × 1 k-point grid and convergence criteria were used in the present study.
The binding energy (Eb) of metal clusters (Pt3M) on the OG support was calculated as Eb = EOG-Pt3M − (EOG + EPt3M), with EOG-Pt3M, EOG, and EPt3M being the total energies of the OG with supported metal cluster, isolated OG sheet, and metal cluster, respectively. The adsorption energies of O2 on the supported metal clusters were calculated with respect to the gas-phase O2 molecules. The Gibbs free energy change (ΔG) of the elementary reactions was used to identify the most favorable oxygen reduction reaction pathway. The free energies of every intermediate along the reaction pathways were calculated based on the computational hydrogen electrode (CHE) model proposed by Nørskov et al. [53,60,61]. The reference potential was set to the standard hydrogen electrode (SHE), i.e., the free energy of (H+ + e) equaled half of the chemical potential of a gas-phase H2 at 0 V under the standard conditions of pH = 0 and the pressure of H2 was 1 bar. The free energy change for each elementary step was determined as ΔG = ΔE + ΔZPE − TΔS + ΔGU, where ΔE corresponds to the reaction energy calculated using DFT. ΔZPE and −TΔS are the corrections due to zero-point energy (ZPE) and entropy (S) at 300 K. Both ZPE and S were calculated based on the harmonic vibrational frequencies. Adding ΔGU = −eU, where e is the unit positive charge and U is the electrode potential with respect to SHE, will allow the external potential to modulate the chemical potential change of an elementary step. The entropic contribution of gas-phase molecules was taken from the National Institute of Standards and Technology database [62].
Bader charge analysis was conducted for every OG-supported Pt3M (OG–Pt3M) catalyst [63,64]. The d-band centers (εd) of Pt atoms in the cluster were calculated based on the follow equation:
ε d = ϵ F ρ · E   d ϵ ϵ F ρ   d ϵ
in which ρ, E, and ϵ F represent the density of electronic states, the energy of electron, and Fermi energy, respectively [50,65]. The reaction free energy of an elementary step refers to the Gibbs free energy difference of reactant and product. The kinetic barrier originating from the non-electrochemical activation of the reactant was not considered in the present study. The barriers associated with the electron and proton transfer were believed to be small and were also not considered in the present study [66].

4. Conclusions

We performed a DFT computational study on the mechanism and activity of O-doped graphene-supported Pt4 and Pt3M (M = Sc, Ti, V, Cr, Mn, Fe, Co, and Ni) catalysts toward oxygen reduction reactions. Co-adsorbed H2O* and O* species are highly unlikely and water formation is the potential limiting step on the supported Pt3M catalysts.
The late transition metal alloyed clusters (M = Mn, Fe, Co, and Ni), especially Pt3Co and Pt3Ni, bound the oxygen intermediates weaker than Pt4 and would have a better ORR activity. On the other hand, alloying with an early transition metal atom (M = Sc, Ti, V, and Cr) increased the binding energy of the oxygen intermediates, thereby hindering the oxygen reduction process. Among the systems examined, OG–Pt3Ni lowered the adsorption energies of O and OH the most and was the best catalyst for ORR. The present study demonstrated that the oxygen adsorption energy alone may not be a good predictor of the ORR activity. Instead, the oxygen adsorption energy difference on the ensemble of the active sites (ΔEOb-Ot) is an effective indicator of the ORR activity.

Supplementary Materials

The following are available online at https://www.mdpi.com/2073-4344/10/2/156/s1, Table S1: Optimized structures for every O-doped graphene (OG) supported Pt3M (M = 3d transition metals) and Pt4 alloy clusters. Table S2. Free energy change (ΔG) of every elementary step of ORR for different OG supported Pt3M and Pt4 clusters as the interfacial site is first hydrogenated. Table S3. Free energy change (ΔG) of every elementary step of ORR for different OG supported Pt3M and Pt4 clusters as the top site is first hydrogenated. Figure S1. End-on adsorption of O2 on the bottom M atom of OG-Pt3M and the corresponding relative adsorption energy with respect to bridging di-oxygen form. Figure S2. Free energy diagrams for oxygen reduction on OG supported Pt3M (M = 3d transition metals) and Pt4 clusters at 0 V. Figure S3. Free energy diagrams for oxygen reduction on OG supported Pt3M (M = 3d transition metals) and Pt4 clusters at 0.85 V (SHE). Figure S4. Average adsorption energies of O2 on the OG-Pt3M clusters as a function of the d-band center (εd) of the Pt atoms in clusters. Figure S5. Relationship between the adsorption energy of O* and OH* on the interfacial sites of Pt3M alloy clusters. Figure S6. The adsorption energy of oxygen atom on the bottom site (ΔE(Ob)) as a function of the Bader charge q(M) of Ob* intermediate.

Author Contributions

C.C. and M.S. performed the calculations and drafted the manuscript; X.Z. and J.H. helped to collect and analyze the data; H.W. and Q.G. revised the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by National Natural Sciences Foundation of China (Grant No.21576204, 21206117, and 21373148).

Acknowledgments

The High Performance Computing Center of Tianjin University is acknowledged for providing services to computing resources.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Zagal, J.H.; Koper, M.T.M. Reactivity Descriptors for the Activity of Molecular MN4 Catalysts for the Oxygen Reduction Reaction. Angewandte Chemie Int. Ed. 2016, 55, 14510–14521. [Google Scholar] [CrossRef] [PubMed]
  2. Zhong, X.; Yu, H.; Wang, X.; Liu, L.; Jiang, Y.; Wang, L.; Zhuang, G.; Chu, Y.; Li, X.; Wang, J.-G. Pt@Au Nanorods Uniformly Decorated on Pyridyne Cycloaddition Graphene as a Highly Effective Electrocatalyst for Oxygen Reduction. ACS Appl. Mater. Interfaces 2014, 6, 13448–13454. [Google Scholar] [CrossRef] [PubMed]
  3. Kumar, A.; Zhang, Y.; Liu, W.; Sun, X. The chemistry, recent advancements and activity descriptors for macrocycles based electrocatalysts in oxygen reduction reaction. Coord. Chem. Rev. 2020, 402, 213047. [Google Scholar] [CrossRef]
  4. Sha, Y.; Yu, T.H.; Merinov, B.V.; Goddard, W.A., III. DFT Prediction of Oxygen Reduction Reaction on Palladium-Copper Alloy Surfaces. ACS Catal. 2014, 4, 1189–1197. [Google Scholar] [CrossRef]
  5. Antolini, E. Effect of Structural Characteristics of Binary Palladium-Cobalt Fuel Cell Catalysts on the Activity for Oxygen Reduction. ChemPlusChem 2014, 79, 765–775. [Google Scholar] [CrossRef]
  6. Wu, P.; Du, P.; Zhang, H.; Cai, C. Graphyne As a Promising Metal-Free Electrocatalyst for Oxygen Reduction Reactions in Acidic Fuel Cells: A DFT Study. J. Phys. Chem. C 2012, 116, 20472–20479. [Google Scholar] [CrossRef]
  7. Hyun, K.; Lee, J.H.; Yoon, C.W.; Kwon, Y. The Effect of Platinum Based Bimetallic Electrocatalysts on Oxygen Reduction Reaction of Proton Exchange Membrane Fuel Cells. Int. J. Electrochem. Sci. 2013, 8, 11752–11767. [Google Scholar]
  8. Zhang, B.-W.; Zhang, Z.-C.; Liao, H.-G.; Gong, Y.; Gu, L.; Qu, X.-M.; You, L.-X.; Liu, S.; Huang, L.; Tian, X.-C.; et al. Tuning Pt-skin to Ni-rich surface of Pt3Ni catalysts supported on porous carbon for enhanced oxygen reduction reaction and formic electro-oxidation. Nano Energy 2016, 19, 198–209. [Google Scholar] [CrossRef]
  9. Stephens, I.E.L.; Bondarenko, A.S.; Perez-Alonso, F.J.; Calle-Vallejo, F.; Bech, L.; Johansson, T.P.; Jepsen, A.K.; Frydendal, R.; Knudsen, B.P.; Rossmeisl, J.; et al. Tuning the Activity of Pt (111) for Oxygen Electroreduction by Subsurface Alloying. J. Am. Chem. Soc. 2011, 133, 5485–5491. [Google Scholar] [CrossRef]
  10. Deng, K.; Xu, Y.; Li, Y.; Dai, Z.; Wang, Z.; Li, X.; Wang, H.; Wang, L. Integration mesoporous surface and hollow cavity into PtPdRh nano-octahedra for enhanced oxygen reduction electrocatalysis. Nanotechnology 2020, 31, 025401. [Google Scholar] [CrossRef]
  11. Zhao, Z.; Chen, C.; Liu, Z.; Huang, J.; Wu, M.; Liu, H.; Li, Y.; Huang, Y. Pt-Based Nanocrystal for Electrocatalytic Oxygen Reduction. Adv. Mater. 2019, 31, 1808115. [Google Scholar] [CrossRef] [PubMed]
  12. Wu, Z.; Dang, D.; Tian, X. Designing Robust Support for Pt Alloy Nanoframes with Durable Oxygen Reduction Reaction Activity. ACS Appl. Mater. Interfaces 2019, 11, 9117–9124. [Google Scholar] [CrossRef] [PubMed]
  13. Stamenkovic, V.R.; Mun, B.S.; Arenz, M.; Mayrhofer, K.J.J.; Lucas, C.A.; Wang, G.; Ross, P.N.; Markovic, N.M. Trends in electrocatalysis on extended and nanoscale Pt-bimetallic alloy surfaces. Nat. Mater. 2007, 6, 241–247. [Google Scholar] [CrossRef]
  14. Cheng, D.; Qiu, X.; Yu, H. Enhancing oxygen reduction reaction activity of Pt-shelled catalysts via subsurface alloying. Phys. Chem. Chem. Phys. 2014, 16, 20377–20381. [Google Scholar] [CrossRef]
  15. Nguyen, M.T.; Wakabayashi, R.H.; Yang, M.; Abruna, H.D.; DiSalvo, F.J. Synthesis of carbon supported ordered tetragonal pseudo-ternary Pt2M′M′′ (M = Fe, Co, Ni) nanoparticles and their activity for oxygen reduction reaction. J. Power Sources 2015, 280, 459–466. [Google Scholar] [CrossRef] [Green Version]
  16. Wang, Q.; Tao, H.; Li, Z.; Wang, G. Effect of iron precursor on the activity and stability of PtFe/C catalyst for oxygen reduction reaction. J. Alloys Compd. 2020, 814, 152212. [Google Scholar] [CrossRef]
  17. Lin, S.-P.; Wang, K.-W.; Liu, C.-W.; Chen, H.-S.; Wang, J.-H. Trends of Oxygen Reduction Reaction on Platinum Alloys: A Computational and Experimental Study. J. Phys. Chem. C 2015, 119, 15224–15231. [Google Scholar] [CrossRef]
  18. Zhang, L.; Fischer, J.M.T.A.; Jia, Y.; Yan, X.; Xu, W.; Wang, X.; Chen, J.; Yang, D.; Liu, H.; Zhuang, L.; et al. Coordination of Atomic Co-Pt Coupling Species at Carbon Defects as Active Sites for Oxygen Reduction Reaction. J. Am. Chem. Soc. 2018, 140, 10757–10763. [Google Scholar] [CrossRef] [Green Version]
  19. Tian, X.; Zhao, X.; Su, Y.-Q.; Wang, L.; Wang, H.; Dang, D.; Chi, B.; Liu, H.; Hensen, E.J.M.; Lou, X.W.; et al. Engineering bunched Pt-Ni alloy nanocages for efficient oxygen reduction in practical fuel cells. Science 2019, 366, 850–856. [Google Scholar] [CrossRef]
  20. Ou, L. The origin of enhanced electrocatalytic activity of Pt-M (M = Fe, Co, Ni, Cu, and W) alloys in PEM fuel cell cathodes: A DFT computational study. Comput. Theor. Chem. 2014, 1048, 69–76. [Google Scholar] [CrossRef]
  21. Gatalo, M.; Moriau, L.; Petek, U.; Ruiz-Zepeda, F.; Sala, M.; Grom, M.; Galun, T.; Jovanovic, P.; Pavlisic, A.; Bele, M.; et al. CO-assisted ex-situ chemical activation of Pt-Cu/C oxygen reduction reaction electrocatalyst. Electrochim. Acta 2019, 306, 377–386. [Google Scholar] [CrossRef]
  22. Zhu, C.; Dong, S. Recent progress in graphene-based nanomaterials as advanced electrocatalysts towards oxygen reduction reaction. Nanoscale 2013, 5, 1753–1767. [Google Scholar] [CrossRef] [PubMed]
  23. Higgins, D.; Zamani, P.; Yu, A.; Chen, Z. The application of graphene and its composites in oxygen reduction electrocatalysis: A perspective and review of recent progress. Energy Environ. Sci. 2016, 9, 357–390. [Google Scholar] [CrossRef]
  24. Chen, H.S.; Liang, Y.T.; Chen, T.Y.; Tseng, Y.C.; Liu, C.W.; Chung, S.R.; Hsieh, C.T.; Lee, C.E.; Wang, K.W. Graphene-supported Pt and PtPd nanorods with enhanced electrocatalytic performance for the oxygen reduction reaction. Chem. Commun. 2014, 50, 11165–11168. [Google Scholar] [CrossRef]
  25. Yan, Z.H.; Wang, M.; Huang, B.X.; Liu, R.M.; Zhao, J.S. Graphene Supported Pt-Co Alloy Nanoparticles as Cathode Catalyst for Microbial Fuel Cells. Int. J. Electrochem. Sci. 2013, 8, 149–158. [Google Scholar]
  26. Boone, C.V.; Maia, G. Lowering metal loadings onto Pt-Pd-Cu/graphene nanoribbon nanocomposites affects electrode collection efficiency and oxygen reduction reaction performance. Electrochim. Acta 2019, 303, 192–203. [Google Scholar] [CrossRef]
  27. Fazio, G.; Ferrighi, L.; Di Valentin, C. Boron-doped graphene as active electrocatalyst for oxygen reduction reaction at a fuel-cell cathode. J. Catal. 2014, 318, 203–210. [Google Scholar] [CrossRef] [Green Version]
  28. Wang, S.Y.; Zhang, L.P.; Xia, Z.H.; Roy, A.; Chang, D.W.; Baek, J.B.; Dai, L.M. BCN Graphene as Efficient Metal-Free Electrocatalyst for the Oxygen Reduction Reaction. Angewandte Chemie Int. Ed. 2012, 51, 4209–4212. [Google Scholar] [CrossRef]
  29. Zhang, L.P.; Xia, Z.H. Mechanisms of Oxygen Reduction Reaction on Nitrogen-Doped Graphene for Fuel Cells. J. Phys. Chem. C 2011, 115, 11170–11176. [Google Scholar] [CrossRef]
  30. del Cueto, M.; Ocon, P.; Poyato, J.M.L. Comparative Study of Oxygen Reduction Reaction Mechanism on Nitrogen-, Phosphorus-, and Boron-Doped Graphene Surfaces for Fuel Cell Applications. J. Phys. Chem. C 2015, 119, 2004–2009. [Google Scholar] [CrossRef]
  31. Shi, X.; Zhang, J.; Huang, T. Significant advantages of sulfur-doped graphene in neutral media as electrocatalyst for oxygen reduction comparing with Pt/C. IOP Conf. Ser. Earth Environ. Sci. 2018, 121, 022011. [Google Scholar] [CrossRef]
  32. Cui, H.; Zhou, Z.; Jia, D. Heteroatom-Doped Graphene as Electrocatalysts for Air Cathodes. Mater. Horiz. 2017, 4, 7–19. [Google Scholar] [CrossRef]
  33. Xu, J.; Wang, Y.; Hu, S. Nanocomposites of graphene and graphene oxides: Synthesis, molecular functionalization and application in electrochemical sensors and biosensors. A review. Microchim. Acta 2017, 184, 1–44. [Google Scholar] [CrossRef]
  34. Su, C.; Loh, K.P. Carbocatalysts: Graphene Oxide and Its Derivatives. Acc. Chem. Res. 2013, 46, 2275–2285. [Google Scholar] [CrossRef]
  35. Tang, Y.; Chen, T.; Guo, W. Surfactant-Free Synthesis of Reduced Graphene Oxide Supported Well-Defined Polyhedral Pd-Pt Nanocrystals for Oxygen Reduction Reaction. Catalysts 2019, 9, 756. [Google Scholar] [CrossRef] [Green Version]
  36. Jin, N.; Han, J.; Wang, H.; Zhu, X.; Ge, Q. A DFT study of oxygen reduction reaction mechanism over O-doped graphene-supported Pt-4, Pt3Fe and Pt3V alloy catalysts. Int. J. Hydrogen Energy 2015, 40, 5126–5134. [Google Scholar] [CrossRef]
  37. Stamenkovic, V.; Mun, B.S.; Mayrhofer, K.J.J.; Ross, P.N.; Markovic, N.M.; Rossmeisl, J.; Greeley, J.; Norskov, J.K. Changing the activity of electrocatalysts for oxygen reduction by tuning the surface electronic structure. Angewandte Chemie Int. Ed. 2006, 45, 2897–2901. [Google Scholar] [CrossRef]
  38. Xu, L.; Ge, Q. Effect of defects and dopants in graphene on hydrogen interaction in graphene-supported NaAlH4. Int. J. Hydrogen Energy 2013, 38, 3670–3680. [Google Scholar] [CrossRef]
  39. Zhang, J.; Wang, Z.; Zhu, Z.P. A Density Functional Theory Study on Mechanism of Electrochemical Oxygen Reduction on FeN3-Graphene. J. Electrochem. Soc. 2015, 162, F1262–F1267. [Google Scholar] [CrossRef]
  40. Qaseem, A.; Chen, F.Y.; Wu, X.Q.; Johnston, R.L. Pt-free silver nanoalloy electrocatalysts for oxygen reduction reaction in alkaline media. Catal. Sci. Technol. 2016, 6, 3317–3340. [Google Scholar] [CrossRef] [Green Version]
  41. Roudgar, A.; Eikerling, M.; van Santen, R. Ab initio study of oxygen reduction mechanism at Pt-4 cluster. Phys. Chem. Chem. Phys. 2010, 12, 614–620. [Google Scholar] [CrossRef]
  42. Fu, C.; Liu, C.; Li, T.; Zhang, X.; Wang, F.; Yang, J.; Jiang, Y.; Cui, P.; Li, H. DFT calculations: A powerful tool for better understanding of electrocatalytic oxygen reduction reactions on Pt-based metallic catalysts. Comput. Mater. Sci. 2019, 170, 109202. [Google Scholar] [CrossRef]
  43. Lim, D.-H.; Wilcox, J. Mechanisms of the oxygen reduction reaction on defective graphene-supported Pt nanoparticles from first-principles. J. Phys. Chem. C 2012, 116, 3653–3660. [Google Scholar] [CrossRef] [Green Version]
  44. Greeley, J.; Stephens, I.E.L.; Bondarenko, A.S.; Johansson, T.P.; Hansen, H.A.; Jaramillo, T.F.; Rossmeisl, J.; Chorkendorff, I.; Norskov, J.K. Alloys of platinum and early transition metals as oxygen reduction electrocatalysts. Nat. Chem. 2009, 1, 552–556. [Google Scholar] [CrossRef]
  45. Liu, J.; Liu, H.; Chen, H.; Du, X.; Zhang, B.; Hong, Z.; Sun, S.; Wang, W. Progress and Challenges Toward the Rational Design of Oxygen Electrocatalysts Based on a Descriptor Approach. Adv. Sci. 2019. [Google Scholar] [CrossRef]
  46. Pasti, I.A.; Gavrilov, N.M.; Mentus, S.V. DFT study of chlorine adsorption on bimetallic surfaces—Case study of Pd3M and Pt3M alloy surfaces. Electrochim. Acta 2014, 130, 453–463. [Google Scholar] [CrossRef]
  47. Hwang, S.J.; Kim, S.K.; Lee, J.G.; Lee, S.C.; Jang, J.H.; Kim, P.; Lim, T.H.; Sung, Y.E.; Yoo, S.J. Role of Electronic Perturbation in Stability and Activity of Pt-Based Alloy Nanocatalysts for Oxygen Reduction. J. Am. Chem. Soc. 2012, 134, 19508–19511. [Google Scholar] [CrossRef]
  48. Karlberg, G.S.; Rossmeisl, J.; Norskov, J.K. Estimations of electric field effects on the oxygen reduction reaction based on the density functional theory. Phys. Chem. Chem. Phys. 2007, 9, 5158–5161. [Google Scholar] [CrossRef] [Green Version]
  49. Tao, H.; Liu, S.; Luo, J.-L.; Choi, P.; Liu, Q.; Xu, Z. Descriptor of catalytic activity of metal sulfides for oxygen reduction reaction: A potential indicator for mineral flotation. J. Mater. Chem. A 2018, 6, 9650–9656. [Google Scholar] [CrossRef]
  50. Xin, H.; Vojvodic, A.; Voss, J.; Norskov, J.K.; Abild-Pedersen, F. Effects of d-band shape on the surface reactivity of transition-metal alloys. Phys. Rev. B 2014, 89, 115114. [Google Scholar] [CrossRef]
  51. Xie, Y.; Wang, Z.-W.; Zhu, T.-Y.; Shu, D.-J.; Hou, Z.-F.; Terakura, K. Breaking the scaling relations for oxygen reduction reaction on nitrogen-doped graphene by tensile strain. Carbon 2018, 139, 129–136. [Google Scholar] [CrossRef]
  52. Yu, T.H.; Hofmann, T.; Sha, Y.; Merinov, B.V.; Myers, D.J.; Heske, C.; Goddard, W.A. Finding Correlations of the Oxygen Reduction Reaction Activity of Transition Metal Catalysts with Parameters Obtained from Quantum Mechanics. J. Phys. Chem. C 2013, 117, 26598–26607. [Google Scholar] [CrossRef] [Green Version]
  53. Norskov, J.K.; Rossmeisl, J.; Logadottir, A.; Lindqvist, L.; Kitchin, J.R.; Bligaard, T.; Jonsson, H. Origin of the Overpotential for Oxygen Reduction at a Fuel-cell Cathode. J. Phys. Chem. B 2004, 108, 17886–17892. [Google Scholar] [CrossRef]
  54. Hansen, H.A.; Viswanathan, V.; Norskov, J.K. Unifying Kinetic and Thermodynamic Analysis of 2 e(−) and 4 e(−) Reduction of Oxygen on Metal Surfaces. J. Phys. Chem. C 2014, 118, 6706–6718. [Google Scholar] [CrossRef]
  55. Seitz, L.C.; Dickens, C.F.; Nishio, K.; Hikita, Y.; Montoya, J.; Doyle, A.; Kirk, C.; Vojvodic, A.; Hwang, H.Y.; Norskov, J.K.; et al. A highly active and stable IrOx/SrIrO3 catalyst for the oxygen evolution reaction. Science 2016, 353, 1011–1014. [Google Scholar] [CrossRef] [PubMed]
  56. 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] [PubMed]
  57. Kresse, G.; Hafner, J. Ab initio molecular dynamics for liquid metals. Phys. Rev. B 1993, 47, 558–561. [Google Scholar] [CrossRef]
  58. Blöchl, P.E.; Jepsen, O.; Andersen, O.K. Improved tetrahedron method for Brillouin-zone integrations. Phys. Rev. B 1994, 49, 16223. [Google Scholar] [CrossRef]
  59. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 1996, 77, 3865–3868. [Google Scholar] [CrossRef] [Green Version]
  60. Rossmeisl, J.; Karlberg, G.S.; Jaramillo, T.; Norskov, J.K. Steady state oxygen reduction and cyclic voltammetry. Faraday Discuss. 2008, 140, 337–346. [Google Scholar] [CrossRef]
  61. Tripković, V.; Skúlason, E.; Siahrostami, S.; Nørskov, J.K.; Rossmeisl, J. The oxygen reduction reaction mechanism on Pt (111) from density functional theory calculations. Electrochim. Acta 2010, 55, 7975–7981. [Google Scholar] [CrossRef]
  62. Johnson, R.D., III. Computational Chemistry Comparison and Benchmark Database; NIST Standard Reference Database Number 101; U.S. Secretary of Commerce; NIST: Gaithersburg, MD, USA, 2010; Volume 219, p. 611. [Google Scholar]
  63. Henkelman, G.; Arnaldsson, A.; Jonsson, H. A fast and robust algorithm for Bader decomposition of charge density. Comput. Mater. Sci. 2006, 36, 354–360. [Google Scholar] [CrossRef]
  64. Bader, R.F. Atoms in Molecules: A Quantum Theory; Oxford University Press: New York, NY, USA, 1990. [Google Scholar]
  65. Seifitokaldani, A.; Savadogo, O.; Perrier, M. Density Functional Theory (DFT) Computation of the Oxygen Reduction Reaction (ORR) on Titanium Nitride (TiN) Surface. Electrochim. Acta 2014, 141, 25–32. [Google Scholar] [CrossRef]
  66. Skulason, E.; Tripkovic, V.; Bjorketun, M.E.; Gudmundsdottir, S.; Karlberg, G.; Rossmeisl, J.; Bligaard, T.; Jonsson, H.; Norskov, J.K. Modeling the Electrochemical Hydrogen Oxidation and Evolution Reactions on the Basis of Density Functional Theory Calculations. J. Phys. Chem. C 2010, 114, 18182–18197. [Google Scholar] [CrossRef]
Figure 1. Net binding energy of Pt3M on O-doped graphene (OG) with respect to Pt4 on OG.
Figure 1. Net binding energy of Pt3M on O-doped graphene (OG) with respect to Pt4 on OG.
Catalysts 10 00156 g001
Figure 2. Adsorption configuration and O–O bond distance (shown in Å) of O2 on OG–Pt3M. The left three columns show O2 adsorbed as di-oxygen species, while the two structures on the right are the O adatoms on supported Pt3Ti and Pt3V. Red, dark blue, and grey spheres represent O, Pt, and C, respectively. The color of M is different and depends on the element.
Figure 2. Adsorption configuration and O–O bond distance (shown in Å) of O2 on OG–Pt3M. The left three columns show O2 adsorbed as di-oxygen species, while the two structures on the right are the O adatoms on supported Pt3Ti and Pt3V. Red, dark blue, and grey spheres represent O, Pt, and C, respectively. The color of M is different and depends on the element.
Catalysts 10 00156 g002
Scheme 1. Reaction pathways of the oxygen reduction reaction (ORR). Labels “b” or “t” correspond to the species on the interfacial M (or Pt0) atom or the top Pt1 atom. The pathway following the bold black arrows is the thermodynamically more favorable one.
Scheme 1. Reaction pathways of the oxygen reduction reaction (ORR). Labels “b” or “t” correspond to the species on the interfacial M (or Pt0) atom or the top Pt1 atom. The pathway following the bold black arrows is the thermodynamically more favorable one.
Catalysts 10 00156 sch001
Figure 3. Free energy profiles of ORR on OG-supported Pt4 (black line), Pt3Ti (red line), and Pt3Ni (blue line) clusters at 0 V (a) and 0.85 V (b) with respect to the standard hydrogen electrode (SHE).
Figure 3. Free energy profiles of ORR on OG-supported Pt4 (black line), Pt3Ti (red line), and Pt3Ni (blue line) clusters at 0 V (a) and 0.85 V (b) with respect to the standard hydrogen electrode (SHE).
Catalysts 10 00156 g003
Figure 4. The Bader charges on the (a) OH* intermediate, (b) Ob* adatoms, and (c) the adsorption energy difference between Ob* and Ot* as a function of the charge changes of interfacial M sites. The vertical dash line indicates the location of the Bader charge values of the corresponding element.
Figure 4. The Bader charges on the (a) OH* intermediate, (b) Ob* adatoms, and (c) the adsorption energy difference between Ob* and Ot* as a function of the charge changes of interfacial M sites. The vertical dash line indicates the location of the Bader charge values of the corresponding element.
Catalysts 10 00156 g004
Figure 5. The correlations of (a) the adsorption energy difference between OH* species adsorbed on the interfacial and top sites (ΔEOHb-OHt), and (b) the free energy change for the rate limiting step (ΔGlimit) of ORR to the oxygen adsorption energy difference of both sites (ΔEOb-Ot).
Figure 5. The correlations of (a) the adsorption energy difference between OH* species adsorbed on the interfacial and top sites (ΔEOHb-OHt), and (b) the free energy change for the rate limiting step (ΔGlimit) of ORR to the oxygen adsorption energy difference of both sites (ΔEOb-Ot).
Catalysts 10 00156 g005
Table 1. Bader charges of the metal atoms, doped oxygen and the carbon atoms bonded with oxygen, and the distance between the doped oxygen atom and the graphene plane (d(O-OG) in Å). The corresponding atoms are labeled in the accompanying structures under the table.
Table 1. Bader charges of the metal atoms, doped oxygen and the carbon atoms bonded with oxygen, and the distance between the doped oxygen atom and the graphene plane (d(O-OG) in Å). The corresponding atoms are labeled in the accompanying structures under the table.
ClusterPt4Pt3ScPt3TiPt3VPt3CrPt3MnPt3FePt3CoPt3Ni
Bader ChargeC1+0.72+0.78+0.93+0.70+0.71+0.68+0.73+0.62+0.75
C2+0.65+0.73+0.32+0.67+0.57+0.58+0.60−0.20−0.22
O−1.61−1.63−1.42−1.65−1.64−1.64−1.65−1.64−1.61
Pt1−0.16−0.59−0.84−0.61−0.51−0.49−0.44−0.39−0.35
Pt2+0.25−0.1+0.04−0.13−0.02−0.03+0.01+0.06+0.09
Pt3+0.19−0.11−0.52−0.09−0.07−0.04+0.01+0.08+0.11
M(or Pt0)−0.12+1.56+2.13+1.70+1.28+1.12+0.97+0.71+0.57
d(O-OG)0.170.510.600.570.420.380.400.200.18
Catalysts 10 00156 i001

Share and Cite

MDPI and ACS Style

Cui, C.; Sun, M.; Zhu, X.; Han, J.; Wang, H.; Ge, Q. Oxygen Reduction Reaction Catalyzed by Pt3M (M = 3d Transition Metals) Supported on O-doped Graphene. Catalysts 2020, 10, 156. https://doi.org/10.3390/catal10020156

AMA Style

Cui C, Sun M, Zhu X, Han J, Wang H, Ge Q. Oxygen Reduction Reaction Catalyzed by Pt3M (M = 3d Transition Metals) Supported on O-doped Graphene. Catalysts. 2020; 10(2):156. https://doi.org/10.3390/catal10020156

Chicago/Turabian Style

Cui, Chaonan, Mengnan Sun, Xinli Zhu, Jinyu Han, Hua Wang, and Qingfeng Ge. 2020. "Oxygen Reduction Reaction Catalyzed by Pt3M (M = 3d Transition Metals) Supported on O-doped Graphene" Catalysts 10, no. 2: 156. https://doi.org/10.3390/catal10020156

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