Next Article in Journal
Imaging of Endometriotic Lesions Using cRGD-MN Probe in a Mouse Model of Endometriosis
Next Article in Special Issue
First Principles Study of the Structure–Performance Relation of Pristine Wn+1Cn and Oxygen-Functionalized Wn+1CnO2 MXenes as Cathode Catalysts for Li-O2 Batteries
Previous Article in Journal
Gap-Free Tuning of Second and Third Harmonic Generation in Mechanochemically Synthesized Nanocrystalline LiNb1−xTaxO3 (0 ≤ x ≤ 1) Studied with Nonlinear Diffuse Femtosecond-Pulse Reflectometry
Previous Article in Special Issue
Numerical Study on Overcoming the Light-Harvesting Limitation of Lead-Free Cs2AgBiBr6 Double Perovskite Solar Cell Using Moth-Eye Broadband Antireflection Layer
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Density Functional Theory Study of Methanol Steam Reforming on Pt3Sn(111) and the Promotion Effect of a Surface Hydroxy Group

1
College of Science, China University of Petroleum (East China), Qingdao 266580, China
2
School of Materials Science and Engineering, China University of Petroleum (East China), Qingdao 266580, China
3
SINOPEC Dalian Research Institute of Petroleum and Petrochemicals Co., Ltd., Dalian 116045, China
*
Authors to whom correspondence should be addressed.
Nanomaterials 2024, 14(3), 318; https://doi.org/10.3390/nano14030318
Submission received: 13 January 2024 / Revised: 2 February 2024 / Accepted: 2 February 2024 / Published: 5 February 2024

Abstract

:
Methanol steam reforming (MSR) is studied on a Pt3Sn surface using the density functional theory (DFT). An MSR network is mapped out, including several reaction pathways. The main pathway proposed is CH3OH + OH → CH3O → CH2O → CH2O + OH → CH2OOH → CHOOH → COOH → COOH + OH → CO2 + H2O. The adsorption strengths of CH3OH, CH2O, CHOOH, H2O and CO2 are relatively weak, while other intermediates are strongly adsorbed on Pt3Sn(111). H2O decomposition to OH is the rate-determining step on Pt3Sn(111). The promotion effect of the OH group is remarkable on the conversions of CH3OH, CH2O and trans-COOH. In particular, the activation barriers of the O–H bond cleavage (e.g., CH3OH → CH3O and trans-COOH → CO2) decrease substantially by ~1 eV because of the involvement of OH. Compared with the case of MSR on Pt(111), the generation of OH from H2O decomposition is more competitive on Pt3Sn(111), and the presence of abundant OH facilitates the combination of CO with OH to generate COOH, which accounts for the improved CO tolerance of the PtSn alloy over pure Pt.

Graphical Abstract

1. Introduction

Methanol steam reforming (MSR) has been widely accepted as a candidate method of generating hydrogen for the on-board application of direct methanol fuel cells (DMFCs) [1,2]. Platinum (Pt) is generally applied as a DMFC catalyst because of its thermal stability and high catalytic activity [2,3,4,5,6]. However, CO molecules are primarily produced from methanol (CH3OH) decomposition and gradually accumulate on Pt, which ultimately leads to CO poisoning and the loss of activity of Pt catalysts [7,8]. Alloying is an effective way to enhance the resistance of metal catalysts. Recently, a PtSn alloy showed promise as an efficient DMFC catalyst with considerable CH3OH electrocatalytic rates compared to Pt [8,9,10,11,12,13,14,15], and it is reported to be active for CO oxidation [16,17,18]. Therefore, an in-depth study of MSR reactions on PtSn is an essential prerequisite to rationally design more efficient and stable PtSn-based catalysts for DMFC applications.
Generally, the MSR process can be summarized as the following two main reaction mechanisms based on previous experimental research studies [19,20,21]. The first mechanism (M1) proceeds with the direct dehydrogenation of CH3OH and the formation of CO; then, CO is oxidized to CO2 via the water–gas shift (WGS) reaction (H2O + CO → H2 + CO2) [22,23]. The second mechanism (M2) includes the reactions of intermediates with adsorbed OH, which is generated from water decomposition (H2O → OH + H), to yield CH2OO, CHOOH, CHOO and, finally, H2 and CO2 [24,25,26]. From the perspective of theoretical research, different catalyst models account for different intermediates and MSR mechanisms. Using density functional theory (DFT) calculations, Luo et al. [27] investigated MSR reactions on Co(0001) and Co(111), and their results showed that the direct decomposition of CH2O to CO is favored rather than CH2OOH formation, indicating the preference of the M1 mechanism. Fajín and Cordeiro [28] performed a DFT investigation on bimetallic Ni−Cu alloy surfaces and also confirmed the M1 mechanism. They found that the MSR evolves mostly through CH3OH decomposition followed by the WGS reaction. In these studies, the surface OH group did not take part in the main reaction pathway, but it can become involved in or influence the MSR process on other metal and alloy surfaces. Lin et al. proposed that MSR reactions on Cu(111) [25,29] and PdZn(111) [26,30] follow the M2 mechanism, that is, the stepwise dehydrogenation of CH3OH occurs first, followed by CH2O formation; then, CH2O combines with OH, which produces a CH2OOH intermediate. Finally, CH2OOH is further dehydrogenated to yield CO2. CH3O dehydrogenation is identified as the rate-determining step on both Cu(111) and PdZn(111) surfaces. Li et al. [31] also confirmed the M2 mechanism of the MSR on an α-MoC(100) surface using DFT calculations. The results suggest that the stepwise O−H and C−H bond scissions of CH3OH yield CH2O. Then, CH2OOH is formed through the combination of CH2O and OH, which is preferred over the decomposition path of CH2O to CHO and H. In addition to its direct involvement in the MSR reaction pathway, the surface OH group can also exert an important influence on the MSR process. Huang et al. [32] studied CH3OH decomposition on PdZn(111) using the DFT and found that the presence of co-adsorbed OH species would hinder C–H bond scission while significantly reducing the energy barrier of the O–H bond scission. Thus, CH3OH preferentially undergoes O–H bond scission to form CH3O because of the influence of OH. Although a great number of efforts have been made to determine the MSR mechanisms of various catalyst models, the detailed MSR process, as well as intermediate information, has not been unambiguously elucidated for specific new catalyst models. At present, there are no theoretical reports available to elucidate the complete MSR mechanism on a PtSn alloy surface. Furthermore, the effect of OH species on the MSR process should also be clarified.
In this work, a periodic DFT investigation is carried out to elucidate the MSR mechanism on a PtSn alloy surface. Among PtxSn catalysts with different Sn contents, Pt3Sn has been proven to have the best performance for the oxidation of methanol and CO in DMFCs [17,18]. Thus, Pt3Sn(111) is chosen as a representative PtSn alloy for DFT calculations. The adsorption structures, elementary reactions and potential energy surfaces (PESs) are illustrated for methanol decomposition and steam reformation processes, and the effect of the OH group on the catalytic mechanism is discussed in detailed.

2. Computational Methods

DFT calculations were conducted using the DMol3 program package [33,34,35]. Exchange and correlation effects were treated using the GGA-PW91 functional [36,37,38]. The DSPP method [39] was applied for Pt and Sn atoms, while C, H and O atoms were treated with an all-electron basis set. The valence electron functions were expanded into a set of numerical atomic orbitals on a double-numerical basis with polarization functions. A Fermi smearing of 0.005 Hartree and a real-space cutoff of 4.5 Å were used. Spin-polarization was applied in all calculations.
The lattice constant of the Pt3Sn was calculated to be 4.01 Å, in good agreement with the experimental value of 4.00 Å [40]. The Pt3Sn(111) surface was built using a p(2 × 2) unit cell with a four-layer slab, and each layer consisted of three Pt atoms and one Sn atom. The height of the vacuum region was set at 12 Å. The reciprocal space was sampled with a (5 × 5 × 1) k-points grid generated automatically using the Monkhorst–Pack method [41]. The uppermost two layers of the slab were relaxed with adsorbates, while two substrate layers were fixed at bulk positions.
High-symmetry sites on Pt3Sn(111) are presented in Figure 1. The adsorption energies (Eads) were calculated as follows: [42,43]
Eads = Eadsorbate + EslabEadsorbate/slab
where Eadsorbate/slab is the energy of the adsorbate/slab adsorption system, and Eadsorbate and Eslab are the energies of the free adsorbate and the clean slab, respectively. By this definition, stable adsorption will have a positive adsorption energy.
Transition state (TS) searches were performed at the same theoretical level using the complete linear synchronous transit/quadratic synchronous transit (LST/QST) method [22,23,24,44]. In this method, an LST maximization was performed, followed by an energy minimization in directions conjugating to the reaction pathway to obtain an approximated TS. The approximated TS was used to perform a QST maximization, and then another conjugated gradient minimization was performed. This cycle was repeated until a stationary point was located. The convergence criterion for the TS searches was set to 0.01 hartree/Å for the root mean square of the atomic forces. The energy barrier (Ea) was determined as the energy difference between the corresponding TS and the initial state (IS), and the reaction energy (Er) was defined as the energy difference between the final state (FS) and the IS.

3. Results and Discussion

3.1. Adsorption Structures and Energies

Figure 2 shows the most stable adsorption geometries of intermediates in MSR, and Table 1 shows the corresponding adsorption energies (Eads) and geometric parameters. For clarity, the geometries and energies of the sub-stable adsorptions of the involved intermediates are presented in Figure S1 and Table S1 of the Supporting Information. In our previous study of methanol decomposition on Pt3Sn(111) [43], several intermediates were calculated in detail, and the most stable adsorption sites along with Eads can be summarized as follows CH3OH at TSn (0.47 eV), CH3O at TSn (1.71 eV), CH2OH at TPt (1.94 eV), CH2O at F2PtSn (0.38 eV), CHOH at B2Pt (3.14 eV), CHO at TPt (2.28 eV), COH at H3Pt (4.05 eV), CO2 at BPtSn (0.11 eV), OH at B2Pt (2.51 eV) and O at F2PtSn (4.12 eV). In this work, we focus on the reformation process, especially the OH-involved paths. Accordingly, the reaction intermediates of MSR are described in detail below.
Carboxymethyl (CH2OOH) is formed through the combination of CH2O and an OH group and preferentially adsorbs at a bridge site via the η1(O)-η1(O) mode, which is different from the unidentate η3(O) modes at the hollow sites of Cu(111) [25], PdZn(111) [26] and Co(111) [27]. The Eads values of CH2OOH are 1.89 eV (BPtSn) and 1.65 (B2Pt), respectively. At the BPtSn site (Figure 2), two C–O bond lengths are 1.35 and 1.52 Å, and the O–Sn and O–Pt distances are 2.15 and 2.31 Å, respectively. Dioxomethylene (CH2OO) was reported to adsorb at a bridge site in a bidentate η1(O)-η1(O) mode on Cu(111), PdZn(111) and Co(111) surfaces [25,26,27]. However, we found that CH2OO has two adsorption modes on Pt3Sn(111) which are the η2(O)-η1(O) mode at the F2PtSn site and the η1(O)-η1(O) mode at the BPtSn site. As listed in Table 1, the η2(O)-η1(O) mode (Figure 2) is more stable with an Eads of 3.24 eV, and the two O–Pt and O–Sn distances are 2.09, 2.26 and 2.27 Å, respectively. The Eads of the η1(O)-η1(O) mode (Figure S1) was calculated to be 3.08 eV, consistent with the previous DFT result for CH2OO adsorption via the same η1(O)-η1(O) mode on Cu(111) [25]. For Formic acid (CHOOH), the most stable adsorption site is TSn, and the corresponding Eads is 0.49 eV. The molecule plane of CHOOH is almost vertical with the OH group pointing down toward the surface (Figure 2). The two C–O bond lengths are 1.23 and 1.32 Å, respectively. The CHOOH at the TPt has a similar adsorption configuration (Figure S1) with a lower Eads of 0.38 eV. The other four adsorption geometries of CHOOH at F2PtSn, F3Pt, H2PtSn and H3Pt involve molecule planes almost parallel to Pt3Sn(111) with ~3.70 Å above the surface (Figure S1). Formate (CHOO) can adsorb at B2Pt and BPtSn with the η1(O)-η1(O) mode, and the BPtSn site is preferred. At the BPtSn site, the molecular plane is perpendicular to the Pt3Sn(111), with an Eads of 2.52 eV; the O–Pt and O–Sn distances are 2.17 and 2.29 Å, respectively (Figure 2). At the B2Pt site, the Eads decreases to 2.08 eV. Carboxyl (COOH) has two isomers which are cis- and trans-COOH, respectively [25]. The cis-COOH can adsorb at the TPt, TSn and T2Pt sites with corresponding Eads values of 2.48, 1.26 and 2.37 eV, respectively. The TPt site can thus be identified as the most stable binding site for cis-COOH; the molecular plane is nearly perpendicular to Pt3Sn(111), with C–Pt and two C–O bond lengths of 2.03, 1.22 and 1.37 Å, respectively (Figure 2). For trans-COOH, the Eads values are 2.35 (TPt), 1.09 (TSn), 2.39 (T2Pt) and 2.41 (TPtSn) eV. The cis isomer binds slightly more strongly to Pt3Sn(111) than its trans counterpart (2.48 vs. 2.41 eV), similar to COOH adsorption on Cu(111) [25]. CO2 adsorbs weakly above the BPtSn and B2Pt sites with the same Eads value of 0.11 eV. At bridge sites, this linear molecule lies almost parallel to the Pt3Sn(111) at a distance of ~4.00 Å above the surface (Figure 2). These results are consistent with those of previous DFT studies of weak CO2 adsorptions over Cu(111) [25], Co(0001) [45] and Co(111) [27]. H2O adsorbs above the TSn site via the O–Sn bond, and the two O–H axes are parallel to the Pt3Sn(111) surface (Figure 2) with bond lengths of 0.98 Å. The binding strength of H2O is very weak, mirrored by a low Eads of 0.01 eV, which is also consistent with weak H2O adsorption on Cu(111) [25], Co(111) [27] and Co(0001) [45]. The most stable sites and Eads values for intermediates via η(O) can be summarized as followed: CH2OOH at BPtSn (1.89 eV), CH2OO at F2PtSn (3.24 eV), HCOOH at TSn (0.49 eV), CHOO at BPtSn (2.52 eV), cis-COOH at TPt (2.48 eV), trans-COOH at BPtSn (2.41 eV) and H2O at TSn (0.01 eV). Taking into account the adsorption properties of other intermediates (CH3OH, CH2OH, CH3O, CHOH, CH2O, COH, CHO, etc.) [43], Sn strengthens the binding of these intermediates to the Pt3Sn(111) surface via η(O).

3.2. Elementary Reaction Steps

The decomposition reactions of CH3OH, CH2OH, CH3O, CH2O and CHO via O−H, C−H and C−O bond scissions were calculated in our previous study [43]. We found that CH3OH decomposition began with O−H bond scission, followed by C−H bond cleavages, that is, CH3OH → CH3O → CH2O → CHO → CO. To identify the optimal MSR pathway, multiple reactions were further investigated in this work, including H2O dissociation into OH and H and subsequent OH-involving reactions with CH3OH and its dehydrogenated intermediates. The configurations of the involved IS, TS and FS are presented in Figure 3 and Figure 4. Sixteen reactions (R1–R16) were considered in total with their thermodynamic and kinetic parameters.
H2O Activation. In the IS, H2O adsorbs weakly above the TSn site. For the reaction R1, the O–H bond is ruptured, with the H atom migrating toward the adjacent Pt atom. The O–H distance of H2O is elongated from 0.98 Å in the IS to 1.62 Å in TS1, as shown in Figure 3. Finally, the OH binds at the B2Pt site, and the H sits at the H3Pt site. This reaction is exothermic by 0.47 eV, with an energy barrier of 0.97 eV. For comparison, the Ea of H2O decomposition on Pt3Sn is much lower than that on Cu(111) (1.11 eV) [25].
CH3OH + OH. In reaction R2, CH3OH and OH adsorb at the TSn and TPt sites in the IS, respectively, and in the FS, CH3O and H2O locate at the same sites as in the IS. In TS2 (Figure 3), the distance of the breaking O−H bond in CH3OH is 1.24 Å, smaller than that in the direct dehydrogenation of CH3OH (0.97 Å) [43] This step is slightly exothermic by 0.04 eV, and the Ea is only 0.02 eV, which is 0.97 eV lower than direct methanol dehydrogenation by O−H bond cleavage at the same site of the TSn [43].
CHxO + OH (x = 0–3). In reaction R3 of CH3O with OH, the energy barrier is 0.84 eV with a reaction energy of −0.87 eV. In TS3, the distance of the breaking O−H bond is 1.27 Å. In reaction R4, the CH2O fragment is weakly bound at the F2PtSn, while the OH fragment stays at TPt site, yielding CH2OOH at the B2Pt site. In TS4, two fragments move to the TPt site, and the distance of the cleaved O−H bond is 2.09 Å. This step is exothermic by 0.45 eV and has an activation barrier of 0.43 eV, lower than that of 0.75 eV for CH2O → CHO [25]. A similar process also occurs on Cu(111) [25] and PdZn(111) [26] For reaction R5, co-adsorbed CHO at the H3Pt site and OH at the TSn site are taken as the IS, and the HCOOH at the H3Pt site is the FS. The distance between C and O atoms is shortened from 3.67 Å in the IS to 1.92 Å in TS5 and to 1.36 Å in the FS. This reaction has an activation barrier of 0.63 eV with an exothermicity of 0.70 eV. For reaction R6, the IS is the co-adsorption of OH at the TSn site and CO at the TPt site, and the FS is COOH at the TPt site. In TS6, the distance of the forming C−O bond is 1.91 Å. This step is exothermic by 0.25 eV, with an activation barrier of 0.39 eV.
CH2OOH dehydrogenation. Two reaction pathways exist for CH2OOH dehydrogenation. The first is C–H bond scission (R7, CH2OOH → CHOOH + H), producing a CHOOH fragment above the BPtSn site with H at the H3Pt site. For TS7 (Figure 3), the C−H distance of the breaking C−H bond is 1.53 Å, which stretches from 1.11 Å in the IS to 3.95 Å in the FS. This reaction has an activation barrier of 0.40 eV with a reaction energy of −0.74 eV. The second is O–H bond scission (R8, CH2OOH → CH2OO + H), which starts with the CH2OOH at the B2Pt site and ends with a co-adsorbed CH2OO fragment at the B2Pt site and H at the TPt site. In TS8 (Figure 3), the O–H distance of the breaking O–H bond is 1.49 Å. This reaction has a higher activation barrier of 1.64 eV and is endothermic by 0.91 eV. Based on thermodynamic and kinetic viewpoints, CH2OOH dehydrogenation on Pt3Sn(111) tends to yield CHOOH rather than CH2OO, that is, the C–H bond scission of reaction R7 is more competitive than the O–H bond scission of reaction R8.
CH2OO and CHOOH dehydrogenation. CH2OO dehydrogenation, denoted as reaction R9, yields bidentate CHOO binding at the BPtSn site and H at the TPt site (Figure 4). This step has a low activation barrier of 0.37 eV and a high exothermicity of 1.55 eV. For TS9, H moves down and locates above the B2Pt site, while CHOO remains at the BPtSn site; the C–H distance of the breaking C–H bond is 1.11 Å. CHOOH dehydrogenation includes C–H bond scission (reaction R10) and O–H bond cleavage (reaction R11). The C–H bond cleavage of CHOOH yields a BPtSn-site-adsorbed COOH fragment and a TPt-site-adsorbed H atom (Figure 4). This reaction involves an energy barrier of 0.44 eV and an endothermicity of 0.01 eV. For TS10, the breaking C–H bond is elongated to 2.07 Å. The O–H bond cleavage of CHOOH is slightly exothermic by 0.02 eV, and the activation barrier is 0.78 eV. For TS11, the O–H bond is elongated by 1.35 Å, and the leaving H adsorbs at the B2Pt site. In the FS, CHOO binds to the Pt3Sn(111) surface in a bidentate configuration, and the detached H locates at the TPt site.
CHOO and COOH dehydrogenation. The CHOO is produced from CH2OO dehydrogenation or CHOOH dehydrogenation via the O–H bond cleavage. The further dehydrogenation of CHOO generates CO2 and an H atom, which is denoted as reaction R12 (Figure 4). In TS12, the C–H distance of the breaking C–H bond is 2.40 Å. After the C–H bond scission, the detached H atom adsorbs at the TPt site, while the CO2 adsorbs above the BPtSn site. This reaction is exothermic by 0.31 eV, and the activation barrier is 1.06 eV. The dehydrogenation process of CHOO could also be accomplished with assistance from an adsorbed OH group (reaction R13). This step starts with co-adsorbed CHOO at the BPtSn site and OH at the TPt site and ends with weakly bonded CO2 and H2O on the surface. The activation barrier of this step is 1.53 eV, and the reaction energy is −1.10 eV. Compared with the direct dehydrogenation of CHOO (R12), the OH-assisted reaction of CHOO with OH to H2O and CO2 (R13) has a relatively higher energy barrier, suggesting that R12 is more favorable than R13. The isomerization of cis-COOH to form trans-COOH (R14) is necessary for COOH dehydrogenation because the O–H bond of the adsorbed COOH points away from the surface in the cis-mode but swings toward the surface in the trans-mode, which is helpful for O–H bond activation. This isomerization step involves an energy barrier of 0.53 eV. Subsequently, CO2 is produced by removing the H atom from trans-COOH (R15), which accounts for an activation barrier of 1.04 eV and a reaction energy of −0.23 eV. For TS15, the O–H distance of the breaking O–H bond is 1.38 Å, and the CO2 is above the BPtSn site and an H atom locates at the TPt site. Similar to CHOO, trans-COOH can also react with OH to generate H2O and CO2 (R16). This step starts with co-adsorbed trans-COOH at the TPt site and OH at the BPtSn site and ends with CO2 above the BPtSn site and H2O above the TPt site. This OH-assisted step is exothermic by 0.74 eV, with a lower activation barrier of 0.11 eV.

3.3. MSR Mechanisms

Based on the calculated results, the potential energy surfaces of MSR on Pt3Sn(111) are presented in Figure 5. CH3OH decomposition with the assistance of OH to form CH3O + H2O and CH3OH dehydrogenation via O–H bond scission to form CH3O + H involve activation barriers of 0.02 and 1.01 eV, respectively. Compared with the direct dehydrogenation of CH3OH to CH3O on Pt3Sn(111), the involvement of the OH group greatly promotes this dehydrogenation step. For the intermediate CH3O, however, the OH group is not helpful for C–H bond cleavage because the direct dehydrogenation of CH3O to CH2O only needs to overcome an activation barrier of 0.42 eV compared with the case of CH3O + OH → CH2O + H2O (Ea = 0.84 eV). For the intermediate CH2O, the transition state of the C–H bond activation with the participation of the OH group was not found in spite of an elaborate search. CH2O has two competitive paths: the direct dehydrogenation, CH2O → CHO (Ea = 0.75 eV), and a combination with the OH group, CH2O + OH → CH2OOH (Ea = 0.43 eV). Therefore, the combination of CH2O with OH is more favorable. The further dehydrogenation of the newly formed CH2OOH has two possibilities, which are O–H and C–H bond activations. We found that the C–H bond scission of CH2OOH → CHOOH + H (Ea = 0.40 eV) is more competitive than the O–H bond cleavage of CH2OOH → CH2OO + H (Ea = 1.64 eV). Similar to CH2OOH, the intermediate CHOOH also tends to break the C–H bond (Ea = 0.44 eV) rather than the O–H bond (Ea = 0.78 eV). CHOOH dehydrogenation yields cis-COOH, followed by an isomerization step toward trans-COOH. Compared with cis-COOH, the adsorption geometry of trans-COOH is favored for O–H bond activation: trans-COOH → CO2 + H (Ea = 1.04 eV). The participation of the OH group substantially reduces the dehydrogenation barrier of trans-COOH via trans-COOH + OH → CO2 + H2O (Ea = 0.11 eV), indicating the promotion effect of the OH group.
Figure 6 summarizes the MSR reaction network based on the direct decomposition of methanol in our previous work [43] and the results calculated in this study. The most favorable pathway follows the M2 mechanism, in which important intermediates were identified as follows:
CH3OH + OH → CH3O + H2O
CH3O → CH2O + H
CH2O + OH → CH2OOH
CH2OOH → CHOOH + H
CHOOH → COOH + H
COOH + OH → CO2 + H2O
H2 production originates from H2O decomposition and the dehydrogenation of important intermediates (CH3O, CH2OOH and CHOOH). The promotion effect of the surface OH group on the conversions of CH3OH, CH2O and trans-COOH is remarkable. In particular, the energy barriers of the O–H bond activation (e.g., CH3OH → CH3O and trans-COOH → CO2) decrease substantially by ~1 eV due to the involvement of the surface OH group, while OH fails to facilitate C–H bond activation. The above results are consistent with previous DFT calculations of CH3OH decomposition by Huang et al. [32] in which the presence of a surface OH group on PdZn(111) impeded the C–H bond scission of CH3OH but substantially decreased the O–H bond-activation barrier. For comparison, Jin et al. [46] found that the OH group on Pt(111) could also be beneficial to MSR reactions, such as CH3OH → CH3O and CH2O + OH → CH2OOH. However, it is relatively difficult to dissociate water and generate the OH group on Pt(111) compared with the direct dehydrogenation of CH3OH. The OH group is only available when the difference in the energy barrier between H2O decomposition and CH3OH dehydrogenation is comparable. Thus, the MSR process on Pt(111) still follows the M1 mechanism, which is stepwise CH3OH decomposition to CO followed by WGS reactions: CH3OH → CH2OH → CHOH → CHO → CO → CO + OH → COOH. In this study, the Pt3Sn(111) surface reduced the difference in the Ea between H2O → H + OH (Ea = 0.97 eV) and CH3OH → CH3O + H (Ea = 1.01 eV)/CH3OH → CH2OH + H (Ea = 1.09 eV). The relatively lower Ea of H2O decomposition indicates the availability of the OH group, which facilitates the MSR process. The initial H2O decomposition to the OH group involves the highest activation barrier of 0.97 eV through the main reaction pathway. Thus, H2O decomposition could be identified as the rate-determining step for MSR on Pt3Sn(111) rather than the commonly accepted C–H bond-cleavage steps such as CH3OH → CH2OH on Pt(111) [46] and CH3O → CH2O on both Cu(111) [25] and PdZn(111) [26]. Compared with the dehydrogenation reactions of CH3OH, the initial H2O → OH step involves relatively higher selectivity on Pt3Sn, which accounts for the improved CO tolerance of PtSn alloys over pure Pt.

4. Conclusions

DFT calculations were performed to investigate possible intermediates and MSR reaction pathways on Pt3Sn(111). The MSR network was mapped out. The most favorable pathway was identified as follows: CH3OH + OH → CH3O → CH2O → CH2O + OH → CH2OOH → CHOOH → COOH → COOH + OH → CO2 + H2O. Along this main reaction pathway, the adsorption strengths of CH3OH, CH2O, CHOOH, H2O and CO2 are relatively weak (Eads < 0.5 eV), while other intermediates are strongly adsorbed at the TSn site for CH3O (Eads = 1.71 eV), at the TPt site for cis-COOH (Eads = 2.48 eV) and at the BPtSn site for CH2OOH (Eads = 1.89 eV) and trans-COOH (Eads = 2.41 eV). H2 production originates from H2O decomposition and the dehydrogenation of important intermediates (CH3O, CH2OOH and CHOOH). H2O decomposition into OH involves an activation barrier of 0.97 eV and was identified as the rate-determining step for the MSR process on Pt3Sn(111). The promotion effect of the surface OH group on the conversions of CH3OH, CH2O and trans-COOH is remarkable. In particular, the energy barriers of the O–H bond activation (e.g., CH3OH → CH3O and trans-COOH → CO2) decrease substantially by ~1 eV due to the involvement of the surface OH group. Compared with the case on Pt(111), the formation of a surface OH group from H2O decomposition is more competitive on Pt3Sn(111), and the presence of abundant OH facilitates the combination of CO with OH to generate COOH, which accounts for the improved CO tolerance of PtSn alloys over pure Pt.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/nano14030318/s1, Figure S1. The other adsorption configurations of reaction intermediates along reaction pathway of Methanol Steam Reformation (MSR) to CO2 on Pt3Sn(111); Table S1. Sub-stable Adsorption Sites, Energies (in eV) and Structural Parameters (in Angstroms) for Intermediates Involved in MSR on Pt3Sn(111).

Author Contributions

Conceptualization, H.Z. and X.L.; Methodology, D.L., Y.C. and H.R.; Software, H.Z., W.Z. and W.G.; Validation, R.L.; Formal analysis, D.L. and R.L.; Investigation, P.H.; Resources, H.Z., Q.S., M.L., X.L. and H.R.; Data curation, W.Z. and H.R.; Writing—original draft, P.H.; Writing—review & editing, H.Z.; Visualization, Y.C.; Supervision, H.Z. and W.G.; Project administration, W.G.; Funding acquisition, Q.S., M.L. and W.G. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by the National Key Research and Development Program of China (2019YFA0708703), the National Natural Science Foundation of China (21776315 and 12104513), the Taishan Scholars Program of Shandong Province (tsqn201909071) and the Shandong Provincial Natural Science Foundation of China (ZR2020QA050 and ZR2023MB034).

Data Availability Statement

Data are contained within the article.

Conflicts of Interest

Authors Qianyao Sun and Ming Li were employed by the company SINOPEC Dalian Research Institute of Petroleum and Petrochemicals Co., Ltd. The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

References

  1. Palo, D.R.; Dagle, R.A.; Holladay, J.D. Methanol Steam Reforming for Hydrogen Production. Chem. Rev. 2007, 107, 3992–4021. [Google Scholar] [CrossRef] [PubMed]
  2. Sá, S.; Silva, H.; Brandão, L.; Sousa, J.M.; Mendes, A. Catalysts for Methanol Steam Reforming—A Review. Appl. Catal. B Environ. 2010, 99, 43–57. [Google Scholar] [CrossRef]
  3. Song, S.Y.; Wang, X.; Li, S.L.; Wang, Z.; Zhu, Q.; Zhang, H.J. Pt Nanohelices with Highly Ordered Horizontal Pore Channels as Enhanced Photothermal Materials. Chem. Sci. 2015, 6, 6420–6424. [Google Scholar] [CrossRef] [PubMed]
  4. Antolini, E. Formation of Carbon-Supported PtM Alloys for Low Temperature Fuel Cells: A Review. Mater. Chem. Phys. 2003, 78, 563–573. [Google Scholar] [CrossRef]
  5. Zhao, L.M.; Wang, S.P.; Ding, Q.Y.; Xu, W.B.; Sang, P.P.; Chi, Y.H.; Lu, X.Q.; Guo, W.Y. The Oxidation of Methanol on PtRu(111): A Periodic Density Functional Theory Investigation. J. Phys. Chem. C 2015, 119, 20389–20400. [Google Scholar] [CrossRef]
  6. Wang, X.; Zhang, Y.B.; Song, S.Y.; Yang, X.G.; Wang, Z.; Jin, R.C.; Zhang, H.J. L-Arginine-Triggered Self-Assembly of CeO2 Nanosheaths on Palladium Nanoparticles in Water. Angew. Chem. Int. Ed. 2016, 55, 4542–4546. [Google Scholar] [CrossRef]
  7. Greeley, J.; Mavrikakis, M. A First-Principles Study of Methanol Decomposition on Pt(111). J. Am. Chem. Soc. 2002, 124, 7193–7201. [Google Scholar] [CrossRef]
  8. Hoseini, S.J.; Barzegar, Z.; Bahrami, M.; Roushani, M.; Rashidi, M. Organometallic Precursor Route for the Fabrication of PtSn Bimetallic Nanotubes and Pt3Sn/Reduced-graphene Oxide Nanohybrid Thin Films at Oilewater Interface and Study of Their Electrocatalytic Activity in Methanol Oxidation. J. Organomet. Chem. 2014, 769, 1–6. [Google Scholar] [CrossRef]
  9. Antolini, E.; Gonzalez, E.R. Effect of Synthesis Method and Structural Characteristics of Pt–Sn Fuel Cell Catalysts on the Electro-oxidation of CH3OH and CH3CH2OH in Acid Medium. Catal. Today 2011, 160, 28–38. [Google Scholar] [CrossRef]
  10. Ishikawa, Y.; Liao, M.S.; Cabrera, C.R. Oxidation of Methanol on Platinum, Ruthenium and Mixed Pt–M Metals (M = Ru, Sn): A Theoretical Study. Surf. Sci. 2000, 463, 66–80. [Google Scholar] [CrossRef]
  11. Honma, I.; Toda, T. Temperature Dependence of Kinetics of Methanol Electro-oxidation on PtSn Alloys. J. Electrochem. Soc. 2003, 150, A1689–A1692. [Google Scholar] [CrossRef]
  12. Zhou, W.J.; Zhou, B.; Li, W.Z.; Zhou, Z.H.; Song, S.Q.; Sun, G.Q.; Xin, Q.; Douvartzides, S.; Goula, M.; Tsiakaras, P. Performance Comparison of Low-Temperature Direct Alcohol Fuel Cells with Different Anode Catalysts. J. Power Sources 2004, 126, 16–22. [Google Scholar] [CrossRef]
  13. Kim, M.; Pratt, S.J.; King, D.A. In Situ Characterization of the Surface Reaction between Chemisorbed Ammonia and Oxygen on Pt(100). J. Am. Chem. Soc. 2000, 122, 2409–2410. [Google Scholar] [CrossRef]
  14. Alcala, R.; Shabaker, J.W.; Huber, G.W.; Marco, A.; Sanchez-Castillo, M.A.; Dumesic, J.A. Experimental and DFT Studies of the Conversion of Ethanol and Acetic Acid on PtSn-Based Catalysts. J. Phys. Chem. B 2005, 109, 2074–2085. [Google Scholar] [CrossRef] [PubMed]
  15. Radael, G.N.; Oliveira, G.G.; Pontes, R.M. A DFT study of ethanol interaction with the bimetallic clusters of PtSn and its implications on reactivity. J. Mol. Graph. Model. 2023, 125, 108621. [Google Scholar] [CrossRef] [PubMed]
  16. García-Rodríguez, S.; Peña, M.A.; Fierro, J.L.G.; Rojas, S. Controlled Synthesis of Carbon-Supported Pt3Sn by Impregnation-reduction and Performance on the Electrooxidation of CO and Ethanol. J. Power Sources 2010, 195, 5564–5572. [Google Scholar] [CrossRef]
  17. Colmati, F.; Antolini, E.; Gonzalez, E.R. Pt–Sn/C Electrocatalysts for Methanol Oxidation Synthesized by Reduction with Formic Acid. Electrochim. Acta 2005, 50, 5496–5503. [Google Scholar] [CrossRef]
  18. Lim, D.H.; Choi, D.H.; Lee, W.D.; Lee, H.I. A New Synthesis of a Highly Dispersed and CO Tolerant PtSn/C Electrocatalyst for Low-Temperature Fuel Cell; its Electrocatalytic Activity and Long-term Durability. Appl. Catal. B 2009, 89, 484–493. [Google Scholar] [CrossRef]
  19. Cai, F.F.; Jessica, J.I.; Fu, Y.; Kong, W.; Zhang, J.; Sun, Y. Low-temperature Hydrogen Production from Methanol Steam Reforming on Zn-modified Pt/MoC Catalysts. Appl. Catal. B Environ. 2020, 264, 118500. [Google Scholar] [CrossRef]
  20. Conant, T.; Karim, A.; Lebarbier, V.; Wang, Y.; Girgsdies, F.; Schlögl, R.; Datye, A. Stability of Bimetallic Pd–Zn Catalysts for the Steam Reforming of Methanol. J. Catal. 2008, 257, 64–70. [Google Scholar] [CrossRef]
  21. Liu, Z.; Guo, B. Microwave Heated Polyol Synthesis of Carbon-supported PtSn Nanoparticles for Methanol Electrooxidation. Electrochem. Commun. 2006, 8, 83–90. [Google Scholar] [CrossRef]
  22. Wang, K.; Gasteiger, H.A.; Markovic, N.M.; Ross, P.N., Jr. On the Reaction Pathway for Methanol and Carbon Monoxide Electrooxidation on Pt-Sn Alloy versus Pt-Ru Alloy Surfaces. Electrochim. Acta 1996, 41, 2587–2593. [Google Scholar] [CrossRef]
  23. Trimm, D.L.; Önsan, Z.I. On Board Fuel Conversion for Hydrogen-fuel-cell-driven Vehicles. Catal. Rev. 2001, 43, 31–84. [Google Scholar] [CrossRef]
  24. Takezawa, N.; Iwasa, N. Steam Reforming and Dehydrogenation of Methanol: Difference in the Catalytic Functions of Copper and Group VIII Metals. Catal. Today 1997, 36, 45–56. [Google Scholar] [CrossRef]
  25. Lin, S.; Johnson, R.S.; Smith, G.K.; Xie, D.; Guo, H. Pathways for Methanol Steam Reforming involving Adsorbed Formaldehyde and Hydroxyl Intermediates on Cu(111): Density Functional Theory Studies. Phys. Chem. Chem. Phys. 2011, 13, 9622–9631. [Google Scholar] [CrossRef] [PubMed]
  26. Lin, S.; Xie, D.Q.; Guo, H. Pathways of Methanol Steam Reforming on PdZn and Comparison with Cu. J. Phys. Chem. C 2011, 115, 20583–20589. [Google Scholar] [CrossRef]
  27. Luo, W.J.; Asthagiri, A. Density Functional Theory Study of Methanol Steam Reforming on Co(0001) and Co(111) Surfaces. J. Phys. Chem. C 2014, 118, 15274–15285. [Google Scholar] [CrossRef]
  28. Fajín, J.L.C.; Cordeiro, M.N.D.S. Insights into the Mechanism of Methanol Steam Reforming for Hydrogen Production over Ni–Cu-Based Catalysts. ACS Catal. 2022, 12, 512–526. [Google Scholar] [CrossRef]
  29. Lin, S.; Xie, D.Q.; Guo, H. Methyl Formate Pathway in Methanol Steam Reforming on Copper: Density Functional Calculations. ACS Catal. 2011, 1, 1263–1271. [Google Scholar] [CrossRef]
  30. Lin, S.; Xie, D.Q.; Guo, H. First-principles Study of the Methyl Formate Pathway of Methanol Steam Reforming on PdZn(111) with Comparison to Cu(111). J. Mol. Catal. A Chem. 2012, 356, 165–170. [Google Scholar] [CrossRef]
  31. Li, J.; Wan, Q.; Lin, G.Z.; Lin, S. DFT Study on the Catalytic Role of α-MoC(100) in Methanol Steam Reforming. J. Chem. Phys. 2022, 35, 639–646. [Google Scholar] [CrossRef]
  32. Huang, Y.; He, X.; Chen, Z.X. Density Functional Study of Methanol Decomposition on Clean and O or OH Adsorbed PdZn(111). J. Chem. Phys. 2013, 138, 184701. [Google Scholar] [CrossRef]
  33. Delley, B. An All-electron Numerical Method for Solving the Local Density Functional for Polyatomic Molecules. J. Chem. Phys. 1990, 92, 508–517. [Google Scholar] [CrossRef]
  34. Delley, B. Fast Calculation of Electrostatics in Crystals and Large Molecules. J. Chem. Phys. 1996, 100, 6107–6110. [Google Scholar] [CrossRef]
  35. Delley, B. From Molecules to Solids with the DMol3 Approach. J. Chem. Phys. 2000, 113, 7756–7764. [Google Scholar] [CrossRef]
  36. Perdew, J.P. Density Functional Approximation for the Correlation Energy of the Inhomogeneous Electron Gas. Phys. Rev. B Condens. Matter Mater. Phys. 1986, 33, 8822–8824. [Google Scholar] [CrossRef] [PubMed]
  37. Perdew, J.P.; Wang, Y. Accurate and Simple Analytic Representation of the Electron-Gas Correlation Energy. Phys. Rev. B Condens. Matter Mater. Phys. 1992, 45, 13244–13249. [Google Scholar] [CrossRef]
  38. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865–3868. [Google Scholar] [CrossRef]
  39. Delley, B. Hardness Conserving Semilocal Pseudopotentials. Phys. Rev. B Condens. Matter Mater. Phys. 2002, 66, 155125–155133. [Google Scholar] [CrossRef]
  40. Hoheisel, M.; Speller, S.; Kuntze, J.; Atrei, A.; Bardi, U.; Heiland, W. Structure of Pt3Sn(110) Studied by Scanning Tunneling Microscopy. Phys. Rev. B Condens. Matter Mater. Phys. 2001, 63, 245403. [Google Scholar] [CrossRef]
  41. Monkhorst, H.J.; Pack, J.D. Special Points for Brillouin-Zone Integrations. Phys. Rev. B 1976, 13, 5188–5192. [Google Scholar] [CrossRef]
  42. Zhu, H.Y.; Guo, W.Y.; Jiang, R.B.; Zhao, L.M.; Lu, X.Q.; Li, M.; Fu, D.L.; Shan, H.H. Decomposition of Methanthiol on Pt(111): A Density Functional Investigation. Langmuir 2010, 26, 12017–12025. [Google Scholar] [CrossRef]
  43. Lu, X.Q.; Deng, Z.G.; Guo, C.; Wang, W.L.; Wei, S.X.; Ng, S.P.; Chen, X.F.; Ding, N.; Guo, W.Y.; Wu, C.M.L. Methanol Oxidation on Pt3Sn (111) for Direct Methanol Fuel Cells: Methanol Decomposition. ACS Appl. Mater. Interfaces 2016, 8, 12194–12204. [Google Scholar] [CrossRef] [PubMed]
  44. Halgren, T.A.; Lipscomb, W.N. The Synchronous-transit Method for Determining Reaction Pathways and Locating Molecular Transition States. Chem. Phys. Lett. 1977, 49, 225–232. [Google Scholar] [CrossRef]
  45. Ma, Y.; Hernández, L.; Guadarrama-Pérez, C.; Balbuena, P.B. Ethanol Reforming on Co(0001) Surfaces: A Density Functional Theory Study. J. Phys. Chem. A 2012, 116, 1409–1416. [Google Scholar] [CrossRef] [PubMed]
  46. Jin, Y.X.; Wang, Y.F.; Zhang, R.X. Theoretical Insight into Hydrogen Production from Methanol Steam Reforming on Pt(111). Mol. Catal. 2022, 532, 112745. [Google Scholar] [CrossRef]
Figure 1. High-symmetry adsorption sites on Pt3Sn(111). The labels ①~⑧ represent the top of the Pt site (TPt), the top of the Sn site (TSn), the Pt-Pt bridge site (B2Pt), the Pt-Sn bridge site (BPtSn), the fcc/hcp site consisting of three surface nearest-neighboring Pt atoms (F3Pt/H3Pt) and the fcc/hcp site consisting of one Sn and two Pt atoms (F2PtSn/H2PtSn), respectively.
Figure 1. High-symmetry adsorption sites on Pt3Sn(111). The labels ①~⑧ represent the top of the Pt site (TPt), the top of the Sn site (TSn), the Pt-Pt bridge site (B2Pt), the Pt-Sn bridge site (BPtSn), the fcc/hcp site consisting of three surface nearest-neighboring Pt atoms (F3Pt/H3Pt) and the fcc/hcp site consisting of one Sn and two Pt atoms (F2PtSn/H2PtSn), respectively.
Nanomaterials 14 00318 g001
Figure 2. Stable adsorption structures of MSR intermediates on Pt3Sn(111). The C, H, O, Pt and Sn atoms are denoted as gray, white, red, blue and gray balls, respectively.
Figure 2. Stable adsorption structures of MSR intermediates on Pt3Sn(111). The C, H, O, Pt and Sn atoms are denoted as gray, white, red, blue and gray balls, respectively.
Nanomaterials 14 00318 g002
Figure 3. MSR reactions involving OH (R1–R8) on Pt3Sn(111). Parameters follow the same notation as in Figure 2.
Figure 3. MSR reactions involving OH (R1–R8) on Pt3Sn(111). Parameters follow the same notation as in Figure 2.
Nanomaterials 14 00318 g003
Figure 4. The MSR reactions involving OH (R9–R16) on Pt3Sn(111). Parameters follow the same notation as in Figure 2.
Figure 4. The MSR reactions involving OH (R9–R16) on Pt3Sn(111). Parameters follow the same notation as in Figure 2.
Nanomaterials 14 00318 g004
Figure 5. Potential energy surface (PES) of MSR on Pt3Sn(111). The detailed reaction pathways of the M1 and M2 mechanisms are shown in red and blue colors, respectively. Data on the direct decomposition of methanol (CH3OH → CH3O → CH2O → CHO → CO) were taken from our precious work [43].
Figure 5. Potential energy surface (PES) of MSR on Pt3Sn(111). The detailed reaction pathways of the M1 and M2 mechanisms are shown in red and blue colors, respectively. Data on the direct decomposition of methanol (CH3OH → CH3O → CH2O → CHO → CO) were taken from our precious work [43].
Nanomaterials 14 00318 g005
Figure 6. Proposed detailed MSR pathways on Pt3Sn(111). The M1 and M2 mechanisms are denoted by red and blue boxes, respectively. The OH-assisted steps are marked with green lines. The generated H and H2O are omitted for clarity.
Figure 6. Proposed detailed MSR pathways on Pt3Sn(111). The M1 and M2 mechanisms are denoted by red and blue boxes, respectively. The OH-assisted steps are marked with green lines. The generated H and H2O are omitted for clarity.
Nanomaterials 14 00318 g006
Table 1. The most stable adsorption sites, geometric parameters (in Å) and energies (in eV) for MSR intermediates on Pt3Sn(111).
Table 1. The most stable adsorption sites, geometric parameters (in Å) and energies (in eV) for MSR intermediates on Pt3Sn(111).
SpeciesSiteModedC/O-Pt/SnEads
CH3OHTSnη1(O)2.640.47
CH3OTSnη1(O)2.051.71
CH2OOHBPtSnη1(O)-η1(O)2.15, 2.311.89
CH2OOF2PtSnη2(O)-η1(O)2.09, 2.26, 2.273.24
CH2OHTPtη1(C)2.141.94
CH2OF2PtSnη1(C)-η2(O)2.13, 2.28, 2.430.38
HCOOHTSnη1(O)2.580.49
CHOOBPtSnη1(O)-η1(O)2.17, 2.292.52
CHOHB2Ptη2(C)2.09, 2.123.14
cis-COOHTPtη1(C)2.032.48
trans-COOHBPtSnη1(C)-η1(O)2.04, 2.552.41
CHOTPtη1(C)2.012.28
COHH3Ptη3(C)2.04, 2.06, 2.114.05
CO2BPtSn--0.11
H2OTSn--0.01
OHB2Ptη2(O)2.23, 2.242.51
OF2PtSnη3(O)2.13, 2.14, 2.144.12
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

He, P.; Zhu, H.; Sun, Q.; Li, M.; Liu, D.; Li, R.; Lu, X.; Zhao, W.; Chi, Y.; Ren, H.; et al. Density Functional Theory Study of Methanol Steam Reforming on Pt3Sn(111) and the Promotion Effect of a Surface Hydroxy Group. Nanomaterials 2024, 14, 318. https://doi.org/10.3390/nano14030318

AMA Style

He P, Zhu H, Sun Q, Li M, Liu D, Li R, Lu X, Zhao W, Chi Y, Ren H, et al. Density Functional Theory Study of Methanol Steam Reforming on Pt3Sn(111) and the Promotion Effect of a Surface Hydroxy Group. Nanomaterials. 2024; 14(3):318. https://doi.org/10.3390/nano14030318

Chicago/Turabian Style

He, Ping, Houyu Zhu, Qianyao Sun, Ming Li, Dongyuan Liu, Rui Li, Xiaoqing Lu, Wen Zhao, Yuhua Chi, Hao Ren, and et al. 2024. "Density Functional Theory Study of Methanol Steam Reforming on Pt3Sn(111) and the Promotion Effect of a Surface Hydroxy Group" Nanomaterials 14, no. 3: 318. https://doi.org/10.3390/nano14030318

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