Next Article in Journal
Colorimetric Method for Sensitive Detection of Microcystin-LR Using Surface Copper Nanoparticles of Polydopamine Nanosphere as Turn-On Probe
Previous Article in Journal
Rotational Dynamics of Linkers in Metal–Organic Frameworks
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Interaction of Ethylene with Irn (n = 1–10): From Bare Clusters to γ-Al2O3-Supported Nanoparticles

1
School of Material Engineering, Shanghai University of Engineering Science, 333 Longteng Road, Songjiang District, Shanghai 201620, China
2
Institute of Physical Chemistry, University of Innsbruck, Innrain 80-82, Innsbruck A-6020, Austria
*
Author to whom correspondence should be addressed.
Nanomaterials 2019, 9(3), 331; https://doi.org/10.3390/nano9030331
Submission received: 18 January 2019 / Revised: 13 February 2019 / Accepted: 20 February 2019 / Published: 2 March 2019

Abstract

:
Comprehending the bond nature of ethylene-metal clusters at the atomic level is important for the design of nanocatalysts and their applications in the fields of fine chemistry and petroleum refining. The growth of Irn (n = 1–10) on γ–Al2O3(110) and ethylene adsorption on bare and γ–Al2O3(110)-supported Irn (n = 1–10) clusters were investigated using the density functional theory (DFT) approach. The mode stability of ethylene adsorption on the bare Irn clusters followed the order π > di-σ > B-T, with the exception of Ir8 where the π structure was less stable than the di-σ configuration. On supported Irn (n = 4–7 and 10) the stability sequence was π > di-σ > di-σ′ (at interface), while on supported Irn (n = 2, 3, 8, and 9) the sequence changed to di-σ > π > di-σ′ (at interface). Two-thirds of ethylene adsorption on the supported Irn clusters were weaker than its adsorption on the bare Irn clusters. The pre-adsorbed ethylene at the interface was found to facilitate the nucleation from the even-sized supported Irn to odd-sized Irn clusters, but hindered the nucleation from the odd-sized Irn to even-sized Irn clusters.

1. Introduction

Interaction of light alkenes with metal clusters has been widely studied in the past due to its important applications in the fields of fine chemistry and petroleum refining. As the smallest alkene and one of the most common probe molecules, the adsorption of ethylene on supported metal catalysts has been widely investigated [1,2,3,4,5]. Keppeler et al. [3] studied ethylene hydrogenation on the NaY and KL zeolite-supported Pt13 cluster and found ethane was the only product.
Recently, many experimental works [6,7,8] have investigated the properties of nanosized metal clusters on alumina support. D’Ippolito et al. [8] found that the addition of iridium to SiO2–Al2O3 enhanced the decalin conversion in the selective ring opening reaction of decalin, while adding HCl barely affected the reaction. Argo et al. [9,10] found that iridium clusters (Ir4 and Ir6) on γ-Al2O3 rearranged slightly to adapt reactive intermediates in the reaction of ethylene hydrogenation, but remained intact using extended X-ray absorption fine structure (EXAFS) study.
Many efforts have been carried out to investigate the adsorption and nucleation of transition metals on alumina using first-principles calculations [11]. Wang et al. [12] studied the growth of Irn (n = 1–5) clusters on the dehydrated and hydrated γ-Al2O3 surfaces. They found the surface hydroxyl hindered the adsorption but facilitated the nucleation of Irn clusters. Chen et al. [13] investigated the nucleation of Irn (n = 2–10) clusters on dehydrated γ-Al2O3(001) and MgO(100) surfaces using density functional theory approach. They found that the growth of Irn (n = 2–10) on these two surfaces was always exothermic. To our best knowledge, however, the study of ethylene interaction with bare Irn metal clusters and γ-Al2O3-supported Irn clusters is still unexplored.
The present work will focus on the effect of the support, which influences not only the adsorption performance of ethylene on Irn, but also the structural features and cluster stability of the nanosized iridium. For this purpose, we use the DFT calculations to explore the growth of Irn (n = 1–10) clusters on γ-Al2O3 and discuss the interaction of ethylene with the bare and γ-Al2O3-supported Irn (n = 1–10) clusters. According to the previous work [14,15,16,17], the γ-Al2O3 surface probably is covered by the hydroxyl groups and the corresponding coverage of surface hydroxyl depends on the preparation temperature. In the reaction of ethylene hydrogenation, the γ-alumina-supported Ir catalysts are typically calcined at 598–673 K during the pretreatment process [9,10]. To correctly describe the surface structures under experimentally relevant conditions, we employ the surface structure constructed by Digne et al. [16] to perform our studies.

2. Materials and Methods

The hydrated γ-Al2O3(110) surface with a hydroxyl coverage of 5.9 OH/nm2 at 673 K was constructed according to [16]. As shown in Figure 1, the covering hydroxyl groups and the surface Al and O atoms on this hydrated (110) surface formed a distinct valley. To prevent the interaction between the neighboring Irn clusters, a large slab of 2 × 2 unit cell (surface area is ~272 Å2) with four layers of thickness containing 224 atoms was employed. Only the bottom two layers were fixed during the geometry optimizations. For the bare Irn systems, all atoms were relaxed.
We used the the Vienna ab initio simulation package (VASP) [18,19] to perform the spin-polarized DFT calculations. The Perdew–Wang exchange-correlation functional [20,21] and projector augmented wave (PAW) method [22] were employed with a cutoff energy of 400 eV. The force threshold of geometry optimization was set to 0.03 eV/Å. A gamma K-point for optimizations and 4 × 4 × 1 K-points for electronic structure calculations were used. All parameters in our calculations were carefully tested, with the change of the calculated Eads smaller than 2%.
The adsorption energy Eads of adsorbates species on substrate was calculated using
Eads = E(Ads/Sub) − E(Sub) − E(Ads)
E(Ads/Sub), E(Sub), and E(Ads) are the total energies of the energy minimized substrate with adsorbates, bare substrate, and gas-phase adsorbates, respectively.
The deformation energy of the adsorbate between its equilibrium structure in the gas phase and the adsorbed state was calculated by
Edef(Ads) = E(Ads′) − E(Ads)
Here E(Ads′) is the total energy of adsorbate species in the gas phase employing the structure displayed in the adsorbed state. Similarly, the deformation energy of the substrate Edef(Sub) was calculated by
Edef(Sub) = E(Sub′) − E(Sub)
E(Sub′) is the total energy of the substrate retaining the adsorbed geometry, but with the adsorbates removed. The adsorbate–substrate interaction energy Eint was calculated by
Eint = E(Ads/Sub) − E(Ads′) − E(Sub′)
For ethylene adsorption, the corresponding adsorption Gibbs free energy ∆Gads(T, P) was defined using
∆Gads(T, P) = Eads − Gɵ(T) − RTln(PC2H4/Pɵ)
Gɵ(T) contains the thermodynamical items of translation, vibration, and rotation of ethylene molecules in the gas phase. PC2H4 is the partial pressure of ethylene. We include zero-point vibrational energy in our present work. A detailed description of Equation (5) can be found in our previous work [23].

3. Results and Discussion

3.1. Gas-Phase Clusters

The structures of gas-phase Irn (n = 2–10) clusters, which are critical to comprehending the nucleation of Irn clusters on γ-Al2O3, have been well studied before [12,13]. Thus, although we considered different gas-phase Irn structures, only the most favorable geometries with the lowest energy are summarized in Table S1. The energetically preferred geometries for Irn (n = 3–8) were linear (D∞h), square planar (in D4h), square pyramid (C4v), triangular prism (D3h), side-face-capped triangular prism (C2v), and cubic structure (Oh), respectively. The Ir9 cluster presented a Cs point group with one Ir atom bridged on two neighboring edge Ir atoms of the cube. The most stable Ir10 geometry yielded the configuration with the Ir2 dimer capping on one face of the cube. The Ir–Ir distance of the Irn (n = 2–10) clusters was in the range of 2.18–2.51 Å, which was shorter than the bulk Ir–Ir distance of 2.74 Å. A similar bond contraction was observed for the bare Rh clusters without ligands experimentally [24] and theoretically [25].

3.2. Small Irn Clusters on Hydrated γ-Al2O3(110)

For Ir adsorption, we considered a series of adsorption sites including seven top (O(A), O(B), O(C), O(D), O(F), Al(1), and Al(2)), ten bridge (O(A)-O(C), O(B)-O(D), O(D)-O(F), O(B′)-O(F), Al(1)-Al(2), Al(1′)-Al(2), O(A)-Al(1), O(B)-Al(1), O(C)-Al(2), and O(D)-Al(2)), and three hollow sites (O(A)-O(C)-Al(3), O(D)-O(F)-Al(2), and O(B′)-O(E′)-O(F)) (Figure 1). The most energetically favorable structures and corresponding adsorption energies Eads are summarized in Figure 2 and Table S2, respectively.
As shown in Figure 2, a single Ir atom preferred to bond to surface Al(1), Al(2), and O(A) atoms yielding Eads of −2.58 eV. For Ir2 adsorption, both Ir atoms bonded to the surface, forming three Ir–Al bonds and two Ir–O bonds with Eads of −2.53 eV. Two Ir atoms bonded to the same surface Al center.
Unlike the case in the gas phase, the adsorbed Ir3 preferred the triangular configuration over the linear one. The adsorbed triangular structure presented Eads of −3.67 eV with respect to the gas-phase triangular trimer, which was 1.14 eV lower in energy than the linear structure, while in the gas phase the former was 0.26 eV higher than the latter. The adsorbed triangular trimer presented a stand-up configuration with two Ir atoms binding to the surface and one Ir atom pointing to the air. While in the adsorbed linear trimer, each Ir atom bonded to the surface.
For Ir4 adsorption, all attempts to obtain the square planar Ir4 cluster (the most stable configuration in gas phase) converted to the geometry with a bent rhombus on the surface. The reconstruction of small deposited metal clusters on the support was also observed for Pdn clusters (n = 1–7) adsorption on α–Al2O3 (0001). Nigam and Majumder [26] found that the Pd4 deformed from the tetrahedral configuration in the gas phase into the bent rhombus in the adsorbed state on the Al2O3 surface. The most stable structure was a tetrahedron with three Ir atoms bonding to the surface and resulting in Eads of −4.12 eV. Our results agree with the earlier experimental and theoretical reports. Using EXAFS spectroscopy, Argo et al. [9] found an Ir–Ir first-shell coordination number in γ-Al2O3-supported Ir4 nanoparticles of about 3. Our previous study showed that Rh4 also prefers the tetrahedral frame on γ-Al2O3 surfaces [25].
The adsorbed Ir5 retained the square pyramid geometry upon adsorption on the surface, with four Ir atoms bonding to the substrate and yielding Eads of −3.74 eV.
Analogous to the case in the gas phase, the most energetically preferred Ir6 geometry was a triangular prism with Eads of −3.92 eV. In the adsorbed state, four Ir atoms made contact with the substrate and two Ir atoms kept away from the surface. The most stable octahedral structure of adsorbed Ir6 was less stable by 1.76 eV (higher in energy) than the triangular prism on the γ-Al2O3 support, and in the gas phase the energy difference between each was 0.70 eV. The coordinates RMSD (root-mean-square deviation) between the triangular prism configuration and the octahedral structure for the Ir6 cluster were ~0.94 in the gas phase and ~0.98 for the supported case, respectively, which is consistent with the energy difference between the two. The octahedral adsorbed Ir6 cluster bonded to the substrate via two Ir atoms (Figure 2h). This was inconsistent with the experimental observation [9,26,27], but in agreement with the previous calculations [13,28,29]. Experimentally, the octahedral Ir6 commonly exists in the form of Ir6 complexes [9]. Argo et al. [9] used the octahedral frame of [Ir6(CO)15]2− upon its decarbonylation to obtain an Ir6 cluster on the γ-Al2O3 support and found that Ir6 maintained the octahedral frame after decarbonylation using EXAFS measurements. Here, the CO ligands helped it to keep the octahedral frame, while several theoretical studies [13,28,29] found that the bare gas-phase Ir6 cluster without any ligand favored a triangular prism with D3h symmetry over the Oh octahedral structure.
For Ir7 adsorption, the energetically preferred structure yielded a similar configuration to the gas-phase cluster with Eads of −3.38 eV. The additional Ir atom to the adsorbed triangular prism Ir6 cluster bonded to the oxygen atom in a surface hydroxyl. A surface distortion (Figure 2i) was observed upon the adsorption of the Ir7 cluster, where the bonded hydroxyl moved upwards to make contact with the Ir atom. The similar adsorbate-induced support rearrangement was found for triangular Ir3 adsorption on γ-Al2O3 (Figure 2c).
The most stable configuration for Ir8 adsorption on alumina was a cubic Ir8 cluster with four Ir atoms bonding to the surface and resulting in Eads of −2.16 eV (Figure 2j). The most favorable structures for Ir9 and Ir10 adsorption exhibited Eads of −2.74 and −2.98 eV, respectively. As shown in Figure 2k,l, one Ir atom in the adsorbed Ir9 and Ir10 cluster bonded to the oxygen center in a surface hydroxyl group.
Similar to our previous finding for Rh adsorption on the hydrated γ-Al2O3(110) surface [25], Irn clusters preferred to adsorb in the valley of the surface O and Al sites instead of on the covering hydroxyls layer, indicating the nanosized metal cluster can retard the transformation of γ-Al2O3 to AlOOH by pre-adsorption on the transformation site. This is consistent with the previous experimental [30] and theoretical [25] observations, confirming that it may be a general effect for a large number of nanosized transition metals.
In summary, Irn preferred to adsorb on the no-hydroxyls-covered area. When cluster diameter (to simplify, the cluster diameter is defined by the longest distance of M–M in the cluster) was smaller than 4.11 Å (e.g., n = 1–6 and 8), the Irn cluster bonded to the surface O and Al atoms only. When cluster diameter was larger than 4.21 Å (n = 7, 9, and 10), besides bonding to the surface O and Al atoms, the Irn cluster bonded to the oxygen of surface hydroxyl as well. The γ–Al2O3 (110) support changed the morphological features and the cluster stability of Ir3 and Ir4 upon their adsorption on the support. The Ir–Ir distance of adsorbed Irn (n = 2–10) in the basal plane underwent an elongation compared with the gas-phase Irn clusters.

3.3. Adsorption of Ethylene on Bare Irn

For ethylene adsorption on the bare Irn cluster, three adsorption modes were considered: di-σ mode, with two carbon ends of ethylene binding to two substrate atoms; π mode, with two carbon ends of ethylene binding to one substrate atom; and a bridge-top (B-T) mode, with one carbon bridging two substrate atoms and the other carbon binding to one substrate atom.
Ethylene adsorption on atomic Ir yielded the largest Eads of −2.96 eV, suggesting the strongest ethylene binding of all considered Irn (n = 1–10) clusters (Table 1). Ethylene adsorption on the Ir2 cluster (Figure 3a) via π-bound mode yielded an adsorption energy Eads of −1.79 eV, which was very close to the di-σ-bound mode that featured Eads of −1.75 eV. This suggests very similar adsorbate binding for both modes at 0 K.
Geometry optimization of ethylene adsorption on the bare Ir3 cluster yielded three different local-energy minima, and those with the lowest total energy for each mode are shown in Figure 3b. The Eads, −1.82 eV (π-bound mode) and −1.75 eV (di-σ-bound mode), differ only slightly. The adsorption of ethylene induced a strong deformation of the Ir3 cluster, where Ir3 changed from the linear structure in the free phase to the bent configuration with the adsorbate. The deformation energy Edef(Ir3) reached as high as 0.23 (0.15) eV for the π (di-σ)-bound mode.
For ethylene adsorption on the square Ir4 cluster (Figure 3c), the calculations found two local equilibrium geometries for each mode. The π-bound state was more favorable than the di-σ-bound mode, where Eads of the π-bound mode was −2.46 eV and the di-σ-bound state yielded Eads of −1.89 eV. To compare with the supported Ir4 cluster, ethylene adsorption on the bent-rhombus Ir4 cluster was studied as well. Ethylene adsorption yielded four di-σ and two π local minima, and the energetically preferred structure for each mode is provided in Figure 3d. The di-σ mode was less stable than the π mode by 0.24 eV (higher in energy).
Ethylene adsorption on Ir5 (two local equilibrium geometries for each mode are found) preferred the π-bound state to the di-σ-bound mode, which accounts for Eads of −2.30 eV (π) and −1.80 eV (di-σ).
Ethylene stabilizes on the triangular-prism Ir6 cluster, and the most favorable adsorption geometries are shown in Figure 3f. Ethylene adsorption on the triangular-prism Ir6 preferred the π state to the di-σ mode. The corresponding adsorption energies were −2.06 and −1.84 eV for the π-bound state and di-σ-bound mode, respectively. Similarly, ethylene adsorption on the octahedral Ir6 cluster (Figure 3g) via the π-bound state was more stable than it was via the di-σ-bound mode, with an energy difference of 0.46 eV.
Geometry optimization of ethylene adsorption on the Ir7 cluster yielded three π and four di-σ local-energy minima. The most stable adsorption geometries of ethylene on the Ir7 cluster for each mode are sketched in Figure 3h. Our results show that the π structure yields Eads of −1.94 eV, which is more favorable than the di-σ structure by 0.09 eV (lower in energy).
The energetically preferred adsorption geometries of ethylene on the Ir8 cluster are sketched in Figure 3i. The calculations yield Eads of −1.63 eV (π) and −1.76 eV (di-σ), suggesting stronger binding for the di-σ structure compared with the π state.
For ethylene adsorption on Ir9, the calculations found three π and five di-σ local-energy minima structures, and the most stable one is shown in Figure 3j. The most stable π and di-σ structures yielded Eads of −2.11 and −1.82 eV, respectively, indicating a stronger stability of the π mode.
The most stable adsorption geometry of ethylene on the bare Ir10 cluster via π-bound mode (of three local equilibrium geometries found) is sketched in Figure 3k. The stability of ethylene adsorption decreased in the sequence of π > di-σ > B-T. Among the obtained three local equilibrium geometries, the most stable π structure yielded Eads of −1.88 eV, which was 0.11 eV lower than the most stable di-σ structure (of seven local equilibrium geometries). Two di-σ structures yielded identical Eads of −1.77 eV. The B-T configuration (Figure 3k) was the least stable, with much smaller Eads of −0.02 eV. Note that the B-T structure was only available on Ir10. All attempts to obtain the B-T configuration on the other Irn clusters resulted in either the π structure or the di-σ mode.
In summary, the stability of ethylene adsorption on the bare Irn clusters decreased in the sequence of π > di-σ > B-T with one exception of Ir8 where the di-σ structure was energetically preferred over the π structure.
To analyze the adsorption energy in more details, we divided it into three contributions according to Eads = Edef(C2H4) + Edef(Irn) + Eint. From the data summarized in Table 1, we can see that the ethylene deformation energies for the di-σ and π structures were within the energy range of 1.53–1.85 and 0.45–0.53 eV, respectively. The deformation energy of ethylene accompanied the adsorption of ethylene along with the C−C bond elongation. The di-σ structure always induced a greater elongation regarding the gas phase than the π mode in our studies, which is in agreement with previous theoretical studies [31,32]. The C−C bond distance in the di-σ mode enlarged to ~1.51Å from 1.33 Å in the gas phase, while the π mode caused a smaller C−C bond extension (~1.43 Å).
Further analysis shows that the deformation of the adsorbate was much stronger than that of the substrate. As the deformation energies of the adsorbed ethylene were larger than 0.45 eV, the deformation energies of Irn clusters were rather small, below 0.23 eV. The energy cost for the deformation could have been compensated by the interaction energy between the adsorbates and the substrate. Despite the cluster–ethylene interaction energy in the di-σ mode always being larger than that in the π mode, it could not compensate for the energy cost difference of the deformation between two modes on most of the Irn clusters (excluding Ir8). As a result, ethylene preferred to adsorb on the bare clusters via the π mode, except for the Ir8 cluster. The reverse preference of adsorption mode on Ir8 was the same with Ir(111) [33], where the di-σ mode was more favorable than the π mode.

3.4. Adsorption of Ethylene on Al2O3(110)-Supported Irn

Next, we studied the adsorption of ethylene on hydrated γ-Al2O3(110)-supported Irn clusters. For ethylene adsorption on Irn/γ-Al2O3, besides the three scenarios described on the bare Irn cluster, one more scenario was considered: the di-σ′ mode at the interface with one carbon atom on the Irn cluster and one carbon atom on the γ-Al2O3 support. Therefore, for ethylene adsorption on Irn/γ-Al2O3, we considered four possible adsorption geometries, including three modes on the supported Irn cluster (π, B-T, and di-σ) and one mode at the interface (di-σ′). The most stable configuration for each mode and their corresponding energies are summarized in Figure 4 and Figure 5 and Table 2.
Ethylene adsorption yielded almost the identical adsorption energy of −2.95 eV both on Ir1/γ-Al2O3(110) and on atomic Ir. The di-σ′ mode at the interface, where one carbon atom of ethylene bonds to the Ir atom and the other carbon end bonds to the oxygen site of the surface hydroxyl group, was less stable by 2.42 eV in energy.
For ethylene adsorption on Ir2/γ-Al2O3(110), the di-σ structure was energetically more favorable than the π state (Figure 4b). The adsorption energy of the di-σ mode was −2.26 eV, while the π mode gave Eads of −2.10 eV. As was the case for Ir1/γ-Al2O3(110), the di-σ′ structure at the interface resulted in much higher Eads of −0.48 eV, indicating this structure was less favorable than the di-σ and π structures on the supported Ir2 cluster thermodynamically.
Three π and three di-σ structures were obtained for ethylene adsorption on Ir3/γ-Al2O3(110) through geometry optimization. In the most stable π mode (Figure 4c), ethylene preferred to adsorb on the upper Ir site while in the most favorable di-σ mode, two carbon ends bonded to the bottom Ir centers. The adsorption energies of the di-σ structure and the π mode were −1.99 and −1.75 eV, respectively, indicating the preference of the di-σ structure. The most stable di-σ′ structure at the interface (of three obtained structures) yielded Eads of −0.58 eV.
Our calculations obtained five di-σ, six π, three monodentate (M) with one carbon end adsorbing on a metal site (Figure S1a), and two di-σ′ structures at the interface for ethylene adsorption on Ir4/γ-Al2O3. The most stable π mode yielded Eads of −1.94 eV. While the most stable di-σ structure yielded Eads of −1.79 eV (Figure 4d). The di-σ′ structure at the interface with one carbon end at a bottom Ir site and one carbon end at the oxygen O(G) site in the hydroxyl group of the γ-Al2O3(110) surface yielded Eads of −0.44 eV. The most stable M mode yielded Eads of –0.66 eV. The stability of the ethylene adsorption mode decreased in the order of π > di-σ > M > di-σ′ (at interface).
For ethylene adsorption on Ir5/γ-Al2O3, ethylene binded to one bottom Ir atom resulting in Eads of −1.84 eV in the most stable π mode (of four local equilibrium geometries found). Ethylene adsorbed on the substrate through its two C atoms bonding to one bottom Ir and one upper Ir (di-σ) yielding Eads of −1.59 eV. Two di-σ′ structures at the interface yielded very close Eads, with an energy difference of 0.08 eV (the more stable one is shown in Figure 4e).
On Ir6/γ-Al2O3, the most stable π-bound ethylene yielded Eads of −1.71 eV (Figure 5a). It adsorbed on a bottom Ir atom. The most stable di-σ-bound structure yielded an adsorption energy of −1.54 eV and bridged one top and one bottom Ir atoms. The most stable di-σ′ structure at the interface (of two local equilibrium geometries) on Ir6/γ-Al2O3(110), with one carbon atom on a bottom Ir atom and one carbon atom on the oxygen site of surface hydroxyl, yielded Eads of −0.56 eV.
Although the most stable supported Ir6 exhibited a triangular-prism configuration and the octahedral structure was less stable by 1.76 eV (higher in energy), as discussed in Section 3.2, the supported octahedral Ir6 cluster was observed by the experiments. Therefore, the adsorption of ethylene on the most stable supported octahedral Ir6 cluster was studied for comparison. On Ir6oct/γ-Al2O3, the most stable π-bound ethylene yielded Eads of −2.19 eV (Figure 5b). It adsorbed on a top Ir atom with two Ir–C bond lengths both of 2.11 Å. The most stable di-σ-bound structure yielded an adsorption energy of −1.91 eV. It bridged one top Ir atom and one middle Ir atom. The present results agree with the previous experimental observation [9] and theoretical results [34,35]. Qi et al. [34] and Valero et al. [35] discovered that the π adsorption mode was more stable than the di-σ mode for ethylene adsorption on Ir4-C/γ-Al2O3(110) and Pd4/γ-Al2O3(110) catalyst. Argo et al. [9] observed that the π adsorption mode was directly relevant to the hydrogenation reaction on Irn/γ-Al2O3 (n = 4 and 6). The most stable M structure (of three local-energy minimum structures, Figure S1b) yielded Eads of −0.62 eV. The di-σ′ (at interface) structure yielded Eads of +0.82 eV suggesting the adsorption was meta-stable and strongly endothermic.
For ethylene adsorption on Ir7/γ-Al2O3, five local equilibrium geometries for both the π and di-σ modes were found. The most stable π state resulted in Eads of −1.96 eV (Figure 5c). The most stable di-σ structure was less stable than the π state, with Eads of −1.79 eV. The di-σ′ structure at the interface yielded Eads of −0.56 eV. Therefore, the stability of these structures decreases in the order π > di-σ > di-σ′ (at interface).
We found three π and four di-σ-bound local-energy minimum structures for ethylene adsorption on Ir8/γ-Al2O3. Analogous to the case on the bare Ir8 cluster, the di-σ configuration was energetically preferred over the π state by 0.18 eV (lower in energy). For the most stable π and di-σ structures, ethylene preferred to adsorb on the upper Ir atoms away from the interface (Figure 5d). The most stable di-σ′ structure at the interface among two obtained local minima yielded an adsorption energy of −0.25 eV, suggesting less stability than the other two modes.
For ethylene adsorption on Ir9/γ-Al2O3, geometry optimization obtained four π and three di-σ structures. In the most stable π and di-σ structures, ethylene preferred to adsorb on the Ir atom, capping on the face of the cube (Figure 5e). Unlike the case on the bare Ir9 cluster, the π structure was less stable than the di-σ state. The corresponding adsorption energies Eads were −1.61 and −1.74 eV for the π and di-σ structures, respectively. For ethylene adsorption at the interface, the adsorption was nearly neutral, with Eads of −0.02 eV.
For ethylene adsorption on Ir10/γ-Al2O3, the most stable π (of five local minima) and di-σ (of four local equilibrium geometries) structures are provided in Figure 5f. Ethylene adsorbed on the upper Ir atom of the cube in the most stable π mode. In the most stable di-σ mode, ethylene adsorbed on the top surface away from the interface where it bridged one Ir atom of the cube and one caped Ir atom. The most stable π structure yielded an adsorption energy of −1.59 eV, 0.27 eV lower in energy than the most stable di-σ state and suggesting the preference of π structure. Similar to the case on supported Ir9, ethylene adsorption at the interface of Ir10/γ-Al2O3, with one carbon bonding to one surface oxygen atom and the other carbon bonding to one bottom Ir site, was slightly endothermic, with Eads of +0.07 eV.
As shown in Table 1 and Table 2, the di-σ mode (either at the interface or on the Irn cluster) caused a stronger distortion of the adsorbed ethylene associated with the larger deformation energy Edef(C2H4) than the π state. This was due to the fundamental nature of the bonds. The bonding in these modes involved a rehybridization of the carbon centers. In Table 1 and Table 2, we provide the mean hybridization value according to the work of [35]. According to this work, the mean hybridization value of the gas-phase ethylene—whose carbon center exhibits sp2 hybridization—was 2, and the carbon center with sp3 hybridization (e.g., C in ethane) yielded the mean hybridization value of 3. We found the mean hybridization value of the di-σ′ mode at the interface was about 3, indicating a complete rehybridization of the carbon centers from sp2 in the gas phase to sp3 in the adsorbed state. The di-σ structure at the interface exhibited the same structure with the gas-phase ethane where two H atoms of ethane were substituted by one Ir atom and one surface O (see Figure 3a). The mean hybridization values of the di-σ and π states on the Irn cluster were ~2.45 and ~2.85, indicating a weaker rehybridization of the carbon centers in the π state than the di-σ structures regarding the gas-phase ethylene (the mean hybridization value was 2). These results reveal an electron transfer from the support to π* orbitals of ethylene. A linear relationship between the deformation energy of ethylene and the mean hybridization value of the carbon centers can be observed in Figure S2.
It is noted that Ir6oct/γ-Al2O3 exhibited the largest deformation energy upon ethylene adsorption, suggesting the strongest reconstruction among all considered Irn/γ-Al2O3. Because of the steric hindrance effect, which limits space at the interface, the supported octahedral Ir6 rearranged itself to accommodate the adsorbed ethylene molecule. Meanwhile the ethylene-support interaction energy was not large enough to balance the deformation energies, resulting in the largest positive Eads of +0.82 eV and indicating the adsorption at the interface was the weakest among all the considered configurations and strongly meta-stable.

3.5. Thermodynamics

The adsorption Gibbs free energy ∆Gads(T, P) is shown in Figure 6. According to the previous work [36,37,38,39], typical molecular adsorption enthalpies for ethylene on silica-supported metal (Pt and Pd) surfaces are about 1.20–1.40 eV in absolute size at 300 K. Our calculation showed that standard adsorption Gibbs free energy at 300 K, with the partial pressure of ethylene at 1 atm, ∆Gɵads(300 K) for ethylene adsorption on the Ir(111) surface [33], fell within the same thermodynamic window as ∆Gɵads(300 K) of 1.32–1.45 eV in absolute size at 1/3 monolayer coverage. Meanwhile, for ethylene adsorption on γ-Al2O3-supported Irn clusters, the calculated ∆Gɵads(300 K) for the most stable π and di-σ structures yielded an energy range between −2.07 and −3.70 eV (Figure 6). This suggest that the ethylene–Irn interactions were much stronger than the ethylene–Ir(111) interactions. And from the thermodynamic view, ethylene adsorption on the bare and supported Irn clusters were much more favorable than those on the Ir(111) surface.
The di-σ′ structure at the interface gave an energy range between +0.09 and −1.38 eV. Due to the limited space at the interface for the supported octahedral Ir6oct cluster, ethylene adsorption at the interface via di-σ′ yielded a positive ∆Gɵads(300 K) value, indicating ethylene adsorption at the interface at 300 K with partial pressure of ethylene of 1 atm is thermodynamically unfavorable.

3.6. Analysis of Electronic Properties

To know more about the charge redistribution upon the adsorption, we examined the local charge flow for adsorbed monomer, Ir4, and Ir10 systems using electron density difference maps. The electron density difference (∆ρ) was calculated by
ρ = ρ(Ads/Sub) − ρ(Ads)fixρ(Sub)fix
where ρ(Ads/Sub) is the total electron density of the adsorbates/substrate complex, ρ(Ads)fix, and ρ(Sub)fix are the electron densities of the isolated adsorbates and substrate in the same geometry as the adsorbed state, respectively.
In the electron density difference maps (Figure 7), some d orbitals of Ir were depleted upon Irn cluster adsorption on the surface, which was associated with the charge redistribution of the formed Ir–O and Ir–Al bonds. Oxygen atoms, which bond to Ir atoms, lose electrons during the adsorption process, causing decreased electron density along the Ir–O bond, while in the regions of the Ir–Al bond, electron density increases. Similar phenomena have been observed for Pd [40] and Rh [25] cluster adsorption on γ-Al2O3. The depletion of Rh (Pd) d orbitals during its adsorption on γ-Al2O3 was balanced by increased electron density along the Rh(Pd)–Al bond.
As shown in Figure 7, electrons previously accumulated along Ir–Al bonds in Irn/γ-Al2O3 were transferred to the Irn/C2H4 part upon C2H4 adsorption on γ-Al2O3-supported Irn. This suggests that the adsorption of ethylene on γ-Al2O3-supported Irn influences the charge distribution at the metal–alumina interface. The following projected density of states (PDOS) analysis further confirms this statement.
The PDOS are summarized in Figure 8 for the Ir1/γ-Al2O3 and Ir10/γ-Al2O3 systems before and after ethylene adsorption. O(A)-Ir(1) and O(D)-Ir(4) were chosen to analyze the metal–support interaction for Ir1/γ-Al2O3 and Ir10/γ-Al2O3, respectively.
Before ethylene adsorption, the d band states of Ir monomer on the alumina support were quite localized and showed an energy gap of ~0.8 eV, while for the supported Ir10 cluster, broad and delocalized d states (either Ir(4) or Ir(8) atoms) showed up and exhibited a finite density of states at the Fermi level. After ethylene adsorption, the d band states became a little smoother for the supported Ir monomer but still localized at the metal cluster, while for the supported Ir10 cluster the d band states became sharper upon ethylene adsorption. We note that the Ir atom in contact with the support but away from the adsorbed ethylene—Ir(4) as labeled—yielded sharper d band states after ethylene adsorption compared with the case before ethylene adsorption. This means the adsorbed ethylene induced charge redistribution at the iridium–alumina interface. The p orbitals of oxygen (carbon) atoms strongly mixed with the low-energy (typically below −3 eV) d states of Ir atoms.

3.7. Effect of Adsorbed Ethylene on Nucleation of Irn Clusters on γ-Al2O3

We notice that the di-σ′ adsorption mode at the interface of Irn(n = 1–8)/γ-Al2O3 occurred at the same place. This raises the question of whether the pre-adsorbed ethylene at the interface would have affected the growth of Irn clusters on the support. To answer this question, we calculated the nucleation energy according to the following equations.
The nucleation or growth energy of Irn clusters from a combination of an adsorbed monomer and an Irn−1 was defined by
Enuc = E(Irn/γ-Al2O3) + E(γ-Al2O3) − E(Irn−1/γ-Al2O3) − E(Ir1/γ-Al2O3)
For pre-adsorbed ethylene at the interface (di-σ′ mode) of Irn/γ-Al2O3, the nucleation energy Enuc was obtained by
Enuc = E(C2H4-Irn/γ-Al2O3) + E(γ-Al2O3) − E(C2H4-Irn−1/γ-Al2O3) − E(Ir1/γ-Al2O3)
For the nucleation of the gas-phase Irn clusters, the nucleation energy Enuc was calculated using
Enuc = E(Irn) − E(Irn−1) − E(Ir1)
For pre-adsorbed ethylene on the gas-phase Irn clusters, the nucleation energy was defined by
Enuc = E(C2H4/Irn) − E(C2H4-Irn−1) − E(Ir1)
Only the most favorable structure is considered for each cluster size.
As shown in Figure 9, the nucleation energies for all Ir cluster sizes we considered were negative, indicating the critical cluster size for Ir growth is 2. Each nucleation step was exothermic, indicating the nucleation process is thermodynamically favorable. In other words, Ir atoms prefer to grow into nanosize Irn clusters atom by atom both in the gas phase and on the γ-Al2O3 surface, which is consistent with the experimental observation. Yentekakis et al. [41] observed the Ir particle agglomeration on γ-Al2O3 in the methane dry reformation reaction using transmission electron microscopy.
An Enuc comparison between hydrated (110) and non-hydrated (001) γ-Al2O3 surfaces revealed that except for Ir3 and Ir6 clusters, the nucleation of Irn clusters were less favorable on the hydrated γ-Al2O3 (110) than the non-hydrated γ-Al2O3 (001). For Ir3 and Ir6 clusters, the trend was reversed. The corresponding Enuc for Ir3 and Ir6 on the hydrated (110) surface was 0.26 and 0.86 eV lower than those on the non-hydrated (001) surface, respectively. It should be pointed out that Ir6 exhibited different adsorption configurations on two surfaces. Chen et al. [13] found the most stable Ir6 adsorption configuration was an octahedron structure on the non-hydrated γ-Al2O3(001) surface, while on the hydrated γ-Al2O3(110) we found the octahedron structure of Ir6 was less stable by 1.76 eV (higher in energy) than the triangular prism (which was the most stable in the gas phase also).
A comparison of the nucleation energies of Irn on the hydrated γ-Al2O3(110) and pre-adsorbed ethylene at the interface (di-σ′ mode) of Irn/γ-Al2O3(110) suggests that the pre-adsorbed ethylene facilitated the nucleation from the even-sized supported Irn to the odd-sized Irn clusters, but hindered the nucleation from the odd-sized Irn to the even-sized Irn clusters. For ethylene pre-adsorbed Irn in the gas phase via the π mode, the pre-adsorbed ethylene hindered the nucleation of Irn (n = 2, 5, and 6), facilitated the nucleation of Ir4 from the Ir3 cluster, and had no effect on the nucleation of Ir3 from Ir2.

4. Conclusions

Understanding the adsorption properties of ethylene on supported metal clusters at the atomic level is of great significance for the design of nanocatalysts and their applications in fine chemistry and petroleum refining. The interaction of ethylene with the bare and hydrated γ-Al2O3 (110)-supported Irn (n = 1–10) clusters was systematically studied by DFT calculations using periodic models. We first identified the most favorable configurations of Irn (n = 1–10) clusters on the γ-alumina support. For Irn (n = 1–6 and 8) with a cluster diameter smaller than 4.11 Å, the Irn cluster adsorbed on the surface O and Al sites only, while for Irn (n = 7, 9, and 10) with a cluster diameter larger than 4.21 Å, besides binding to the surface O and Al atoms, the Irn cluster binded to the oxygen of surface hydroxyl as well. The γ–Al2O3 (110) support changed the morphological features and the cluster stability of Ir3 and Ir4 upon their adsorption on the support.
The stability of ethylene adsorption on the bare Irn clusters decreased in the sequence of π > di-σ > B-T, with an exception of Ir8 where a preference of the di-σ structure over the π structure was found. Compared to ethylene adsorption on the bare Irn clusters, the γ-Al2O3 support reversed the stability of π and di-σ modes on the supported Irn (n = 2, 3, and 9) but kept the same for the other bare and supported Irn (n = 4–8 and 10) clusters. For supported Irn (n = 2, 3, 8, and 9), the stability of the ethylene adsorption mode decreased in the order di-σ > π > di-σ′ (at interface) while, on the supported Irn (n = 4–7 and 10), the sequence changed to π > di-σ > M > di-σ′ (at interface). M mode was only available on the supported Ir4 and Ir6oct clusters. The carbon centers of the adsorbed ethylene completely rehybridized at the interface from sp2 in the gas phase to sp3 in the adsorbed state, while for adsorptions on Irn, the orbital hybridization of the carbon centers in adsorbed ethylene was between sp2 and sp3.
Among 21 pairs (for example, the π mode on the bare Ir1 and supported Ir1 clusters counts as 1 pair), 6 ethylene adsorption modes on the supported Irn clusters were stronger than on the bare ones, including π and di-σ on Ir2, di-σ mode on Ir3, π and di-σ on Ir6oct, and π mode on Ir7. One pair—π mode on the supported and bare Ir1—showed similar stability. For the remaining 14 pairs, ethylene adsorption on the supported Irn clusters was weaker than on the bare ones.
Thermodynamic analysis showed that ethylene adsorption on the bare and supported Irn clusters was much more favorable than on the Ir(111) surface. The interface between the Irn clusters and γ-Al2O3 support provided a new adsorption mode di-σ′ (at interface), which was the weakest among all adsorption modes.
The pre-adsorbed ethylene at the interface was found to facilitate the nucleation from the even-sized supported Irn to odd-sized Irn clusters, but hindered the nucleation from the odd-sized Irn to even-sized Irn clusters.
The electronic analysis shows that the adsorbed ethylene induced charge redistribution between the support and metal clusters. The d band states became a little smoother for the supported Ir monomer, but still localized to the metal cluster upon ethylene adsorption, while for the supported Ir10 cluster, the d band states became sharper after ethylene adsorption.

Supplementary Materials

The following are available online at https://www.mdpi.com/2079-4991/9/3/331/s1, Figure S1: M configuration of ethylene adsorption. Figure S2: Linear fitting of ethylene deformation energy Edef(C2H4) (eV) and mean hybridization value of the carbon center in adsorbed ethylene. Table S1: Geometry, magnetic moment (M), and energy of gas phase Irn (n = 2–10) Table S2: Adsorption energy Eads (eV) and Nucleation energy Enuc (eV) for Irn clusters on γ-Al2O3 surfaces.

Author Contributions

DFT simulations, X.-R.S.; data analysis, S.Z., Y.Z., W.G., P.M., N.L., and X.-R.S.; writing—original draft preparation, X.-R.S.; writing—review and editing, W.G., N.L., P.M., and X.-R.S.; supervision, X.-R.S.

Funding

This research was funded by the National Natural Science Foundation of China, grant number 21703137, and the Shanghai Pujiang Program, grant number 17PJ1403100.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Henry, R.; Komurcu, M.; Ganjkhanlou, Y.; Brogaard, R.Y.; Lu, L.; Jens, K.-J.; Berlier, G.; Olsbye, U. Ethene oligomerization on nickel microporous and mesoporous-supported catalysts: Investigation of the active sites. Catal. Today 2018, 299, 154–163. [Google Scholar] [CrossRef]
  2. Quesada, J.; Arreola-Sánchez, R.; Faba, L.; Díaz, E.; Rentería-Tapia, V.M.; Ordóñez, S. Effect of Au nanoparticles on the activity of TiO2 for ethanol upgrading reactions. Appl. Catal. A Gen. 2018, 551, 23–33. [Google Scholar] [CrossRef]
  3. Keppeler, M.; Bräuning, G.; Radhakrishnan, S.G.; Liu, X.; Jensen, C.; Roduner, E. Reactivity of diatomics and of ethylene on zeolite-supported 13-atom platinum nanoclusters. Catal. Sci. Technol. 2016, 6, 6814–6823. [Google Scholar] [CrossRef] [Green Version]
  4. Guo, Z.; Liu, Y.; Liu, Y.; Chu, W. Promising SiC support for Pd catalyst in selective hydrogenation of acetylene to ethylene. Appl. Surf. Sci. 2018, 442, 736–741. [Google Scholar] [CrossRef]
  5. Aho, A.; Eränen, K.; Lemus-Yegres, L.J.; Voss, B.; Gabrielsson, A.; Salmi, T.; Murzin, D.Y. Ethylene epoxidation over supported silver catalysts—Influence of catalyst pretreatment on conversion and selectivity. J. Chem. Technol. Biotechnol. 2018, 93, 1549–1557. [Google Scholar] [CrossRef]
  6. Sikora, P.; Augustyniak, A.; Cendrowski, K.; Nawrotek, P.; Mijowska, E.; Sikora, P.; Augustyniak, A.; Cendrowski, K.; Nawrotek, P.; Mijowska, E. Antimicrobial Activity of Al2O3, CuO, Fe3O4, and ZnO Nanoparticles in Scope of Their Further Application in Cement-Based Building Materials. Nanomaterials 2018, 8, 212. [Google Scholar] [CrossRef] [PubMed]
  7. Bruk, L.; Titov, D.; Ustyugov, A.; Zubavichus, Y.; Chernikova, V.; Tkachenko, O.; Kustov, L.; Murzin, V.; Oshanina, I.; Temkin, O. The Mechanism of Low-Temperature Oxidation of Carbon Monoxide by Oxygen over the PdCl2–CuCl2/γ-Al2O3 Nanocatalyst. Nanomaterials 2018, 8, 217. [Google Scholar] [CrossRef] [PubMed]
  8. D’Ippolito, S.A.; Ballarini, A.D.; Pieck, C.L. Influence of Support Acidity and Ir Content on the Selective Ring Opening of Decalin over Ir/SiO2–Al2O3. Energy Fuels 2017, 31, 5461–5471. [Google Scholar] [CrossRef]
  9. Argo, A.M.; Odzak, J.F.; Gates, B.C. Role of cluster size in catalysis: Spectroscopic investigation of gamma-Al2O3-supported Ir4 and Ir6 during ethene hydrogenation. J. Am. Chem. Soc. 2003, 125, 7107–7115. [Google Scholar] [CrossRef] [PubMed]
  10. Argo, A.M.; Odzak, J.F.; Goellner, J.F.; Lai, F.S.; Xiao, F.S.; Gates, B.C. Catalysis by oxide-supported clusters of iridium and rhodium: Hydrogenation of ethene, propene, and toluene. J. Phys. Chem. B 2006, 110, 1775–1786. [Google Scholar] [CrossRef] [PubMed]
  11. Wang, J.; Yu, H.; Geng, L.; Liu, J.; Han, L.; Chang, L.; Feng, G.; Ling, L. DFT study of Hg adsorption on M-substituted Pd(111) and PdM/γ-Al2O3(110) (M = Au, Ag, Cu) surfaces. Appl. Surf. Sci. 2015, 355, 902–911. [Google Scholar] [CrossRef]
  12. Wang, Y.; Su, Y.; Kang, L. Stability and nucleation of Irn (n = 1–5) clusters on different γ-Al2O3 surfaces: A density functional theory study. Phys. Lett. A 2016, 380, 718–725. [Google Scholar] [CrossRef]
  13. Chen, Y.; Huo, M.; Chen, T.; Li, Q.; Sun, Z.; Song, L. The properties of Irn (n = 2–10) clusters and their nucleation on γ-Al2O3 and MgO surfaces: From ab initio studies. Phys. Chem. Chem. Phys. 2015, 17, 1680–1687. [Google Scholar] [CrossRef] [PubMed]
  14. Heemeier, M.; Frank, M.; Libuda, J.; Wolter, K.; Kuhlenbeck, H.; Bäumer, M.; Freund, H.J. The influence of OH groups on the growth of rhodium on alumina: A model study. Catal. Lett. 2000, 68, 19–24. [Google Scholar] [CrossRef]
  15. Libuda, J.; Frank, M.; Sandell, A.; Andersson, S.; Brühwiler, P.A.; Bäumer, M.; Mårtensson, N.; Freund, H.J. Interaction of rhodium with hydroxylated alumina model substrates. Surf. Sci. 1997, 384, 106–119. [Google Scholar] [CrossRef]
  16. Digne, M.; Sautet, P.; Raybaud, P.; Euzen, P.; Toulhoat, H. Use of DFT to achieve a rational understanding of acid-basic properties of γ-alumina surfaces. J. Catal. 2004, 226, 54–68. [Google Scholar] [CrossRef]
  17. Digne, M.; Sautet, P.; Raybaud, P.; Euzen, P.; Toulhoat, H. Hydroxyl Groups on γ-Alumina Surfaces: A DFT Study. J. Catal. 2002, 211, 1–5. [Google Scholar] [CrossRef]
  18. Kresse, G.; Furthmüller, J. Efficiency of ab-initio total energy calculations for metals and semiconductors using a plane-wave basis set. Comput. Mater. Sci. 1996, 6, 15–50. [Google Scholar] [CrossRef]
  19. Kresse, G.; Furthmüller, J. Efficient iterative schemes for ab-initio total-energy calculations using a plane-wave basis set. Phys. Rev. B 1996, 54, 11169–11186. [Google Scholar] [CrossRef]
  20. Perdew, J.P.; Chevary, J.A.; Vosko, S.H.; Jackson, K.A.; Pederson, M.R.; Singh, D.J.; Fiolhais, C. Atoms, molecules, solids, and surfaces: Applications of the generalized gradient approximation for exchange and correlation. Phys. Rev. B 1992, 46, 6671–6687. [Google Scholar] [CrossRef] [Green Version]
  21. Perdew, J.P.; Wang, Y. Accurate and simple analytic representation of the electron-gas correlation energy. Phys. Rev. B 1992, 45, 13244–13249. [Google Scholar] [CrossRef]
  22. Kresse, G.; Joubert, D. From ultrasoft pseudopotentials to the projector augmented-wave method. Phys. Rev. B Condens. Matter Mater. Phys. 1999, 59, 1758–1775. [Google Scholar] [CrossRef]
  23. Shi, X.-R.; Wang, S.-G.; Wang, J. Chemisorption of oxygen and subsequent reactions on low index surfaces of β–Mo2C: Insights from first-principles thermodynamics and kinetics. J. Mol. Catal. A Chem. 2016, 417, 53–63. [Google Scholar] [CrossRef]
  24. Fulton, J.L.; Linehan, J.C.; Autrey, T.; Balasubramanian, M.; Chen, Y.; Szymczak, N.K. When is a Nanoparticle a Cluster? An Operando EXAFS Study of Amine Borane Dehydrocoupling by Rh4–6 Clusters. J. Am. Chem. Soc. 2007, 129, 11936–11949. [Google Scholar] [CrossRef] [PubMed]
  25. Shi, X.-R.; Sholl, D.S. Nucleation of Rhn (n = 1–5) Clusters on γ-Al2O3 Surfaces: A Density Functional Theory Study. J. Phys. Chem. C 2012, 116, 10623–10631. [Google Scholar] [CrossRef]
  26. Nigam, S.; Majumder, C. Adsorption of Small Palladium Clusters on the alfa-Al2O3(0001) Surface: A First Principles. Stud. J. Phys. Chem. C 2012, 116, 2863–2871. [Google Scholar] [CrossRef]
  27. Kulkarni, A.; Lobo-Lapidus, R.J.; Gates, B.C. Metal clusters on supports: Synthesis, structure, reactivity, and catalytic properties. Chem. Commun. 2010, 46, 5997–6015. [Google Scholar] [CrossRef] [PubMed]
  28. Du, J.; Sun, X.; Chen, J.; Jiang, G. A theoretical study on small iridium clusters: Structural evolution, electronic and magnetic properties, and reactivity predictors. J. Phys. Chem. A 2010, 114, 12825–12833. [Google Scholar] [CrossRef] [PubMed]
  29. Pawluk, T.; Hirata, Y.; Wang, L. Studies of iridium nanoparticles using density functional theory calculations. J. Phys. Chem. B 2005, 109, 20817–20823. [Google Scholar] [CrossRef] [PubMed]
  30. Ravenelle, R.M.; Copeland, J.R.; Kim, W.-G.; Crittenden, J.C.; Sievers, C. Structural Changes of γ-Al2O3-Supported Catalysts in Hot Liquid Water. ACS Catal. 2011, 1, 552–561. [Google Scholar] [CrossRef]
  31. Neurock, M.; van Santen, R.A. A First Principles Analysis of C−H Bond Formation in Ethylene Hydrogenation. J. Phys. Chem. B 2000, 104, 11127–11145. [Google Scholar] [CrossRef]
  32. Heard, C.J.; Siahrostami, S.; Grönbeck, H. Structural and Energetic Trends of Ethylene Hydrogenation over Transition Metal Surfaces. J. Phys. Chem. C 2016, 120, 995–1003. [Google Scholar] [CrossRef]
  33. Shi, X.R.; Kong, H.; Wang, S.; Wang, H.; Qin, Z.; Wang, J. Mechanistic Insights into Ethylene Transformations on Ir(111) by Density Functional Calculations and Microkinetic Modeling. ChemPhysChem 2017, 18, 906–916. [Google Scholar] [CrossRef] [PubMed]
  34. Qi, K.; Zhao, J.-M.; Wang, G.-C. A density functional theory study of ethylene hydrogenation on MgO- and γ-Al2O3-supported carbon-containing Ir4 clusters. Phys. Chem. Chem. Phys. 2015, 17, 4899–4908. [Google Scholar] [CrossRef] [PubMed]
  35. Valero, M.C.; Raybaud, P.; Sautet, P. Interplay between molecular adsorption and metal-support interaction for small supported metal clusters: CO and C2H4 adsorption on Pd4/gamma-Al2O3. J. Catal. 2007, 247, 339–355. [Google Scholar] [CrossRef]
  36. Natal-Santiago, M.A.; Podkolzin, S.G.; Cortright, R.D.; Dumesic, J.A. Microcalorimetric studies of interactions of ethene, isobutene, and isobutane with silica-supported Pd, Pt, and PtSn. Catal. Lett. 1997, 45, 155–163. [Google Scholar] [CrossRef]
  37. Shen, J.; Hill, J.M.; Watwe, R.M.; Spiewak, B.E.; Dumesic, J.A. Microcalorimetric, Infrared Spectroscopic, and DFT Studies of Ethylene Adsorption on Pt/SiO2 and Pt−Sn/SiO2 Catalysts. J. Phys. Chem. B 1999, 103, 3923–3934. [Google Scholar] [CrossRef]
  38. Shen, J.; Hill, J.M.; Watwe, R.M.; Podkolzin, S.G.; Dumesic, J.A. Ethylene adsorption on Pt/Au/SiO2 catalysts. Catal. Lett. 1999, 60, 1–9. [Google Scholar] [CrossRef]
  39. Podkolzin, S.G.; Alcala, R.; de Pablo, J.J.; Dumesic, J.A. Monte Carlo Simulations of Reaction Kinetics for Ethane Hydrogenolysis over Pt. J. Phys. Chem. B 2002, 106, 9604–9612. [Google Scholar] [CrossRef]
  40. Valero, M.C.; Raybaud, P.; Sautet, P. Nucleation of Pdn (n = 1–5) clusters and wetting of Pd particles on gamma-Al2O3 surfaces: A density functional theory study. Phys. Rev. B 2007, 75, 45427. [Google Scholar] [CrossRef]
  41. Yentekakis, I.V.; Goula, G.; Panagiotopoulou, P.; Katsoni, A.; Diamadopoulos, E.; Mantzavinos, D.; Delimitis, A. Dry Reforming of Methane: Catalytic Performance and Stability of Ir Catalysts Supported on γ-Al2O3, Zr0.92Y0.08O2−δ (YSZ) or Ce0.9Gd0.1O2−δ(GDC) Supports. Top. Catal. 2015, 58, 1228–1241. [Google Scholar] [CrossRef]
Figure 1. Top view of hydrated γ-Al2O3(110) surface with 5.9 OH/nm2. Atoms in the first layer and other layers are displayed in sphere and line forms, respectively. White: H, red: O, and pink: Al.
Figure 1. Top view of hydrated γ-Al2O3(110) surface with 5.9 OH/nm2. Atoms in the first layer and other layers are displayed in sphere and line forms, respectively. White: H, red: O, and pink: Al.
Nanomaterials 09 00331 g001
Figure 2. Side (top) and top views (bottom) of the most favorable Irn (n = 1–10) adsorption structures with the corresponding adsorption energies Eads (in eV) on the hydrated γ-Al2O3(110) surface. Largest spheres: Ir; the other color/label scheme is identical to Figure 1.
Figure 2. Side (top) and top views (bottom) of the most favorable Irn (n = 1–10) adsorption structures with the corresponding adsorption energies Eads (in eV) on the hydrated γ-Al2O3(110) surface. Largest spheres: Ir; the other color/label scheme is identical to Figure 1.
Nanomaterials 09 00331 g002
Figure 3. The most stable structure for each ethylene-binding mode on bare Irn. The energy difference (eV) with respect to the most stable configuration (energy set to 0) is labeled at the top of the figure.
Figure 3. The most stable structure for each ethylene-binding mode on bare Irn. The energy difference (eV) with respect to the most stable configuration (energy set to 0) is labeled at the top of the figure.
Nanomaterials 09 00331 g003
Figure 4. Side view (top) and top view (bottom) of most stable structure for each ethylene-binding mode on γ-Al2O3(110)-supported Irn (n = 1–5). The energy difference (eV) with respect to the most stable configuration (energy set to 0) is labeled at the top of the structure. For the di-σ′ mode at the interface on supported Ir monomer, x and y are the different views of the same structure along two directions. The largest spheres are Ir, and the other color/label scheme is identical to Figure 1.
Figure 4. Side view (top) and top view (bottom) of most stable structure for each ethylene-binding mode on γ-Al2O3(110)-supported Irn (n = 1–5). The energy difference (eV) with respect to the most stable configuration (energy set to 0) is labeled at the top of the structure. For the di-σ′ mode at the interface on supported Ir monomer, x and y are the different views of the same structure along two directions. The largest spheres are Ir, and the other color/label scheme is identical to Figure 1.
Nanomaterials 09 00331 g004
Figure 5. Side (top row) and top views (bottom row) of the most stable structure for each ethylene-binding mode on γ-Al2O3(110)-supported Irn (n = 6–10). The energy difference (eV) with respect to the most stable configuration (energy set to 0) is labeled at the top of the figure. The color/label scheme is identical to Figure 4.
Figure 5. Side (top row) and top views (bottom row) of the most stable structure for each ethylene-binding mode on γ-Al2O3(110)-supported Irn (n = 6–10). The energy difference (eV) with respect to the most stable configuration (energy set to 0) is labeled at the top of the figure. The color/label scheme is identical to Figure 4.
Nanomaterials 09 00331 g005
Figure 6. Standard adsorption Gibbs free energy at 300 K; ∆Gqads(300 K) of ethylene adsorption on (a) bare Irn clusters and (b) hydrated (110) γ-Al2O3-supported Irn clusters.
Figure 6. Standard adsorption Gibbs free energy at 300 K; ∆Gqads(300 K) of ethylene adsorption on (a) bare Irn clusters and (b) hydrated (110) γ-Al2O3-supported Irn clusters.
Nanomaterials 09 00331 g006
Figure 7. Electron density difference map for hydrated γ-Al2O3(110)-supported (a) Ir atom, (b) Ir4, and (c) Ir10 systems. Left column: Irn cluster adsorbed on γ-Al2O3; right column: ethylene adsorbed on Irn/γ-Al2O3. Depletion regions: blue; accumulation region: yellow.
Figure 7. Electron density difference map for hydrated γ-Al2O3(110)-supported (a) Ir atom, (b) Ir4, and (c) Ir10 systems. Left column: Irn cluster adsorbed on γ-Al2O3; right column: ethylene adsorbed on Irn/γ-Al2O3. Depletion regions: blue; accumulation region: yellow.
Nanomaterials 09 00331 g007
Figure 8. Density of states for the supported Ir1 (upper row) and Ir10 (bottom row) clusters on the hydrated γ-Al2O3 (110) surface projected on the bonded Ir atom, surface O, and C in the adsorbed ethylene (π mode) before and after ethylene adsorption. (Dotted lines: p states; solid lines: d states.).
Figure 8. Density of states for the supported Ir1 (upper row) and Ir10 (bottom row) clusters on the hydrated γ-Al2O3 (110) surface projected on the bonded Ir atom, surface O, and C in the adsorbed ethylene (π mode) before and after ethylene adsorption. (Dotted lines: p states; solid lines: d states.).
Nanomaterials 09 00331 g008
Figure 9. Nucleation energies Enuc of the gas-phase Irn cluster; ethylene pre-adsorbed Irn in the phase via the π mode; Irn on hydrated γ-Al2O3(110); and ethylene pre-adsorbed at the interface of Irn/γ-Al2O3 (110). For comparison, nucleation energy for Irn on γ-Al2O3(001) [13] is also included.
Figure 9. Nucleation energies Enuc of the gas-phase Irn cluster; ethylene pre-adsorbed Irn in the phase via the π mode; Irn on hydrated γ-Al2O3(110); and ethylene pre-adsorbed at the interface of Irn/γ-Al2O3 (110). For comparison, nucleation energy for Irn on γ-Al2O3(001) [13] is also included.
Nanomaterials 09 00331 g009
Table 1. Adsorption energy Eads (eV), standard adsorption Gibbs free energy at 300 K with the partial pressure of ethylene at 1atm ∆Gɵads(300 K) (eV), ethylene deformation energy Edef(C2H4) (eV), Irn cluster deformation energy Edef(Irn) (eV), interaction energy Eint (eV), carbon–carbon bond distance dC-C (Å), and mean hybridization value (hyd.) for ethylene adsorption on bare Irn (n = 1–10) clusters.
Table 1. Adsorption energy Eads (eV), standard adsorption Gibbs free energy at 300 K with the partial pressure of ethylene at 1atm ∆Gɵads(300 K) (eV), ethylene deformation energy Edef(C2H4) (eV), Irn cluster deformation energy Edef(Irn) (eV), interaction energy Eint (eV), carbon–carbon bond distance dC-C (Å), and mean hybridization value (hyd.) for ethylene adsorption on bare Irn (n = 1–10) clusters.
nmodeEads∆Gɵads(300 K)Edef(C2H4)Edef(Irn)EintdC-Chyb.
1π−2.96−3.700.630−3.591.442.50
2π−1.79−2.530.520.02−2.331.442.44
di-σ−1.75−2.491.850.15−3.751.532.89
3π−1.82−2.560.520.23−2.571.432.45
di-σ−1.75−2.491.530.15−3.431.512.84
4squπ−2.46−3.200.480.02−2.961.432.42
di-σ−1.89−2.631.700.11−3.701.522.84
4benπ−2.25−2.990.500.12−2.871.432.42
di-σ−2.01−2.751.600.11−3.721.512.84
5π−2.30−3.040.520.02−2.841.432.44
di-σ−1.89−2.631.710.04−3.641.512.89
6triπ−2.06−2.800.550.04−2.651.442.45
di-σ−1.84−2.581.650.01−3.501.512.84
6octπ−2.02−2.760.570.06−2.651.442.47
di-σ−1.56−2.301.750.14−3.451.522.86
7π−1.94−2.680.540.01−2.491.442.42
di-σ−1.85−2.591.680.17−3.701.512.86
8π−1.63−2.370.540.17−2.341.442.46
di-σ−1.76−2.501.740.08−3.581.522.84
9π−2.11−2.850.520.05−2.681.442.44
di-σ−1.82−2.561.720.02−3.561.522.85
10π−1.88−2.620.450.15−2.481.432.43
di-σ−1.77−2.511.710.13−3.611.522.85
Table 2. Adsorption energy Eads (eV), standard adsorption Gibbs free energy at 300 K with the partial pressure of ethylene at 1atm ∆Gɵads(300 K) (eV), ethylene deformation energy Edef(C2H4) (eV), substrate deformation energy Edef(Irn/γ-Al2O3) (eV), interaction energy Eint (eV), carbon–carbon bond distance dC-C (Å), and mean hybridization value (hyd.) for ethylene adsorption on hydrated γ-Al2O3(110)-supported Irn (n = 1–10) clusters.
Table 2. Adsorption energy Eads (eV), standard adsorption Gibbs free energy at 300 K with the partial pressure of ethylene at 1atm ∆Gɵads(300 K) (eV), ethylene deformation energy Edef(C2H4) (eV), substrate deformation energy Edef(Irn/γ-Al2O3) (eV), interaction energy Eint (eV), carbon–carbon bond distance dC-C (Å), and mean hybridization value (hyd.) for ethylene adsorption on hydrated γ-Al2O3(110)-supported Irn (n = 1–10) clusters.
nmodeaEads∆Gɵads(300 K)Edef(C2H4)Edef(Irn/γ-Al2O3)EintdC-Chyb.
1π−2.95−3.690.610.04−3.601.442.51
di-σ′−0.53−1.273.751.49−5.771.523.00
2π−2.10−2.840.530.19−2.821.442.45
di-σ−2.26−3.001.700.73−4.691.512.80
di-σ′−0.48−1.223.871.06−5.411.512.98
3π−1.75−2.490.550.11−2.411.442.46
di-σ−2.15−2.891.910.69−4.751.522.93
di-σ′−0.58−1.323.731.08−5.391.512.96
4π−1.94−2.680.480.32−2.741.422.45
di-σ−1.79−2.531.550.15−3.491.512.84
di-σ′−0.44−1.183.721.08−5.241.482.97
5π−1.84−2.580.520.02−2.381.432.48
di-σ−1.59−2.331.650.08−3.321.512.84
di-σ′−0.64−1.383.920.87−5.431.512.96
6triπ−1.71−2.450.480.20−2.391.422.44
di-σ−1.54−2.281.710.06−3.311.522.87
di-σ′−0.56−1.303.750.99−5.301.512.96
6octπ−2.19−2.930.580.24−3.011.442.47
di-σ−1.91−2.651.760.27−3.941.522.86
di-σ′+0.820.083.691.58−4.451.512.95
7π−1.96−2.700.570.19−2.721.442.48
di-σ−1.79−2.531.420.48−3.691.492.84
di-σ′−0.56−1.303.720.85−5.131.522.99
8π−1.52−2.260.550.09−2.161.442.46
di-σ−1.67−2.411.750.07−3.491.522.85
di-σ′−0.25−0.993.680.90−4.831.512.96
9π−1.61−2.350.540.19−2.341.432.49
di-σ−1.74−2.481.490.38−3.611.492.78
di-σ′−0.02−0.763.671.09−4.781.512.99
10π−1.59−2.330.530.18−2.301.432.48
di-σ−1.32−2.061.570.13−3.021.502.82
di-σ′+0.07−0.673.661.10−4.691.513.00
1 π and di-σ: two C atoms bond to Ir atom(s); di-σ′: di-σ mode at the interface where one C atom binds to O of the support and the other C atom binds to Ir.

Share and Cite

MDPI and ACS Style

Shi, X.-R.; Zhang, Y.; Zong, S.; Gu, W.; Ma, P.; Lu, N. Interaction of Ethylene with Irn (n = 1–10): From Bare Clusters to γ-Al2O3-Supported Nanoparticles. Nanomaterials 2019, 9, 331. https://doi.org/10.3390/nano9030331

AMA Style

Shi X-R, Zhang Y, Zong S, Gu W, Ma P, Lu N. Interaction of Ethylene with Irn (n = 1–10): From Bare Clusters to γ-Al2O3-Supported Nanoparticles. Nanomaterials. 2019; 9(3):331. https://doi.org/10.3390/nano9030331

Chicago/Turabian Style

Shi, Xue-Rong, Yajing Zhang, Shibiao Zong, Wen Gu, Pan Ma, and Na Lu. 2019. "Interaction of Ethylene with Irn (n = 1–10): From Bare Clusters to γ-Al2O3-Supported Nanoparticles" Nanomaterials 9, no. 3: 331. https://doi.org/10.3390/nano9030331

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