Open Access Article
Daniel
Arribas
a,
Víctor
Villalobos-Vilda
a,
Ezequiel
Tosi
ab,
Paolo
Lacovig
b,
Alessandro
Baraldi
bc,
Luca
Bignardi
b,
Silvano
Lizzit
b,
José Ignacio
Martínez
a,
Pedro Luis
de Andres
ad,
Alejandro
Gutiérrez
ef,
José Ángel
Martín-Gago
a and
Pablo
Merino
*a
aInstituto de Ciencia de Materiales de Madrid, c/Sor Juana Inés de la Cruz, 3, Spain. E-mail: pablo.merino@csic.es
bElettra-Sincrotrone Trieste S.C.p.A., S.S. 14 km 163.5, Trieste, Italy
cPhysics Department, University of Trieste, Via Valerio 2, 34127 Trieste, Italy
dOn leave of absence at nanoteeche@surfaces Laboratory, Swiss Federal Laboratories for Materials Science and Technology, 8600-Dübendorf, Switzerland
eApplied Physics Department, Universidad Autónoma de Madrid, Calle Francisco Tomás y Valiente, 7, 28049 Madrid, Spain
fInstituto Nicolás Cabrera, Cantoblanco, Calle Francisco Tomás y Valiente, 7, 28049 Madrid, Spain
First published on 4th July 2023
The catalytic dehydrogenation of alkanes constitutes a key step for the industrial conversion of these inert sp3-bonded carbon chains into other valuable unsaturated chemicals. To this end, platinum-based materials are among the most widely used catalysts. In this work, we characterize the thermal dehydrogenation of n-octane (n-C8H18) on Pt(111) under ultra-high vacuum using synchrotron-radiation X-ray photoelectron spectroscopy, temperature-programmed desorption and scanning tunneling microscopy, combined with ab initio calculations. At low activation temperatures, two different dehydrogenation stages are observed. At 330 K, n-C8H18 effectively undergoes a 100% regioselective single C–H bond cleavage at one methyl end. At 600 K, the chemisorbed molecules undergo a double dehydrogenation, yielding double bonds in their carbon skeletons. Diffusion of the dehydrogenated species leads to the formation of carbon molecular clusters, which represents the first step towards poisoning of the catalyst. Our results reveal the chemical mechanisms behind the first stages of alkane dehydrogenation on a platinum model surface at the atomic scale, paving the way for designing more efficient dehydrogenation catalysts.
Efforts have been made to investigate the on-surface alkane dehydrogenation processes on transition metals using theoretical calculations10–13 and techniques such as temperature-programmed desorption, infrared spectroscopy, or X-ray photoelectron spectroscopy.14–19 Such studies have made it possible to determine statistical and thermodynamic parameters of the dehydrogenation reactions on these substrates (reactivity, selectivity, or activation energy, among others). Most of them are based on theoretical simulations to study dehydrogenation since the experimental data at the atomic scale are relatively scarce in the literature. Moreover, the focus is generally placed on the dehydrogenation products of lighter alkanes (less than six carbon atoms), which require high temperatures for dehydrogenation, favoring the poisoning of the catalyst. Alternatively, for long unbranched alkanes (n-C44H90 and n-C32H66), scanning probe microscopy studies have been carried out on Au(110),20–22 Pt(110)20 and Cu(111),23 in which linear polymerization, C–C bond cracking and cascade dehydrogenation were respectively reported.
In this work, we investigate the dehydrogenation products from n-octane (n-C8H18) deposited on Pt(111) under ultra-high vacuum (UHV) conditions from an atomic-scale perspective. We take advantage of the high intensity and energy resolution of synchrotron radiation for the in situ characterization of the on-surface reactions by X-ray photoelectron spectroscopy (XPS). Temperature-programmed desorption (TPD) experiments enable us to follow the release of hydrogen from the system. Scanning tunneling microscopy (STM) images allow the morphological characterization of dehydrogenated products with submolecular resolution. The complementary information provided by the experimental data combined with ab initio calculations allows us to determine in detail the mechanism of dehydrogenation reactions and the corresponding energy barriers. Our results show that n-octane undergoes two initial stages of dehydrogenation on the Pt(111) surface. In the first one, a terminal C(sp3)–H bond is cleaved and a C–Pt bond is formed to stabilize the radical. In the second one, a double dehydrogenation occurs at the third and fourth C atoms, which, together with the diffusion and aggregation of adsorbates, results in the formation of carbon clusters that represents the first step towards catalyst poisoning.
We performed our synchrotron radiation XPS and TPD experiments at the SuperESCA beamline of the Elettra Synchrotron facility (Italy). High-resolution XPS spectra were acquired at normal emission with the sample held at 77 K. The spectra shown in this work correspond to the normalized sum of 30 sequential scans. The photon energy was set to 400 eV to maximize high surface sensitivity and photoionization cross-sections for both Pt-4f7/2 and C-1s core levels. We observed no beam-induced damage to the samples during spectra acquisition. For the analysis of the spectra, we subtracted a Shirley background and fitted each individual contribution with a Gaussian–Lorentzian sum that approximates a Voigt function. The free parameters for each peak were the binding energy, the area, the full width at half maximum (FWHM, typically about 400 meV), and the Gaussian–Lorentzian ratio. No further restrictions were imposed to the parameters, apart from requiring consistent values for all peaks in all spectra. For clarity, our XPS fits do not include vibrational profiles for each peak; rather we consider the vibrations later in the analysis of the intensities and for the assignment of each peak.
For TPD experiments (heating rate of 2 K s−1), the mass spectrometer head, equipped with a quartz glass cylinder with a diameter similar to that of the sample, was approached to the single crystal so that the background signal was minimized. The measurable masses ranged from 1 to 100 atomic mass units. We acquired several TPD spectra to ensure reproducibility. A custom population dynamics model was employed to fit the experimental TPD results. Alkane dehydrogenation and rehydrogenation processes and hydrogen recombination to yield molecular H2 were described with rate constants by following the Arrhenius equation. The pre-exponential factors and the activation energies were used as free parameters and H2 was assumed to desorb immediately after its formation.25 Simulated TPD spectra were convoluted with a Gaussian envelope (σ = 15 K) to account for the observed experimental broadening.
STM measurements were performed in an in-house UHV system after n-octane deposition on the Pt(111) substrate. n-Octane coverages were high enough to saturate the surface, typically after a dosage of 1–5 Langmuir. To reveal the different chemical species that appear on the surface upon thermal dehydrogenation, we carried out measurements after briefly annealing the samples to the corresponding temperatures and allowing them to cool to room temperature. Images were acquired in constant current mode with a chemically etched tungsten tip conditioned under UHV by standard procedures and post-processed using the WSxM software.26
FIREBALL was employed to simulate the theoretical STM images based on the established interfacial geometries along the different dehydrogenation stages.31 FIREBALL uses a local-orbital formulation to achieve self-consistency on the orbital occupation numbers, calculated using the orthonormal Löwdin orbitals. In our calculations, we used a basis set of optimized single and double numerical atomic orbitals: sp3d5 for C and Pt and ss* for H. Tunneling currents for the STM images were computed using the Keldysh–Green function formalism, together with the first-principles tight-binding Hamiltonian obtained from the FIREBALL code, as explained in full detail elsewhere.32,33 Our STM theoretical simulation approach includes a detailed description of the electronic properties of both the W-tip and the sample simultaneously. All computed theoretical STM images were obtained under constant-current scanning conditions by moving the W-tip perpendicularly to the sample in each STM scanning step to search for a pre-selected fixed value of the tunnel current, mimicking the experimental procedure.
According to the experimental data, three different chemical stages can be established in their corresponding temperature ranges, representing different degrees of dehydrogenation of the molecules. The temperature limits for each range were defined as the temperatures of the main maxima in the TPD spectrum (stage 1: 100 K–330 K; stage 2: 330 K–600 K; stage 3: 600 K–900 K, depicted in blue, yellow and green, respectively in Fig. 1a). We fit the TPD spectrum with an Arrhenius-based dynamic rate model (Fig. 1a red line). In this model, dehydrogenation reactions are governed by first-order dynamic equations in which the constants follow the Arrhenius equation,
, where k corresponds to Boltzmann's constant, Ai corresponds to the pre-exponential factors, and Ei to the activation energies, of each reaction.14,25 Dehydrogenation populates the surface with atomic hydrogen, whose recombination and desorption yielding gas phase molecular hydrogen proceed as a second-order reaction. In our model, we assume that diffusion of atomic hydrogen is faster than recombination over the measured temperature range (for further details, see the ESI, section 1†). The good agreement of the model with the TPD experimental data demonstrates that the observed chemical changes can be correlated with well-defined activation energies that quantitatively compare with barriers obtained from DFT calculations of the most probable energy path (see the last paragraph of this section and Fig. 5).
The chemical evolution with the temperature and the proposed division into three temperature ranges is also supported by temperature-resolved XPS (Fig. S3†). Fig. 1b–d show high-resolution XPS spectra obtained within each of the three temperature regions. Within each stage, the line shape of the XPS C-1s peak remains almost unchanged, while there are clear differences in the line shape between the different stages. Additionally, we introduce a detailed analysis of the XPS spectra by deconvolution into basic components. The stark differences evidence strong chemical changes in the molecules. In the next sections, we show that the combination of XPS, TPD, STM and DFT results allows us to determine the atomic-scale configuration of the three stages and assign them to (i) physisorption of intact n-octane molecules in the range of 100 K–330 K, (ii) dehydrogenation of a methyl termination in the range of 330 K–600 K, and (iii) fast dehydrogenation of two adjacent carbons, formation of a double bond and molecular aggregation, in the range of 600 K–900 K.
0] high symmetry directions (see schematics in Fig. 2c and Fig. S9†). The measured dimensions and the morphology suggest that the observed features correspond to single, intact, non-dehydrogenated n-octane molecules adsorbed on the surface (Fig. 2d and e). No apparent morphological modifications (i.e., cis–trans isomerization or on-surface fragmentation) were observed. The surface density amounts to approximately 0.5 molecules per nm2. We notice that tip scanning can induce, in some cases, small displacement of the molecules around their adsorption positions but no tip- or thermal-induced diffusion was detected. The absence of self-assembled molecular islands suggests a repulsive lateral interaction between adsorbates.
To further understand the adsorption geometry in this temperature range, we performed DFT simulations of the minimum energy configuration of an n-octane molecule on a Pt(111) surface. The calculations showed that the molecules physisorb horizontally following the high-symmetry directions in an all-trans configuration with the plane defined by the carbon zig-zag skeleton tilted 14° with respect to the platinum surface. Such a result slightly differs from the common assumption that n-alkane molecules lie flat on the (111) surface, with the zig-zag skeleton parallel to the surface plane,35,36 and showcases the need for advanced calculations for interpreting the atomistic structures of small hydrocarbons on surfaces. The DFT-based STM simulations are in good agreement with the experimental observations (Fig. 2f). In this tilted physisorbed configuration, one of the two H atoms of each methylene group lies closer to the surface while the other protrudes to the vacuum side. In our DFT simulations, we observed neither significant bond deformation nor torsion in agreement with previously reported infrared (IR) experiments.35,37,38
High-resolution XPS confirms this physisorption pattern. Spectra taken at 290 K (Fig. 1b) can be deconvoluted into four different contributions at 283.10 eV, 283.50 eV, 283.95 eV and 284.35 eV, which can be interpreted in terms of a weak interaction of n-octane with Pt(111). n-Octane has two types of non-equivalent carbon atoms: two carbon atoms at the methyl-ends (–CH3) and six internal carbon atoms (–CH2–).39 In addition, the XPS spectra of hydrocarbons show a characteristic vibrational profile due to the simultaneous excitation of quantum vibrational states of the core-excited state (i.e. the final state).40–46 This vibrational excitation produces a loss of kinetic energy of the photoelectrons, giving rise to satellite peaks at higher binding energies. The main excited vibrational mode is the C–H stretching mode (2950 cm−1 ≈ 0.36 eV). Therefore, each C-1s carbon peak is expected to have at least one vibrational replica at higher energies (around +0.36 eV). Taking this into account, we can assign the highest binding energy peak (284.30 eV, blue dashed curve in Fig. 1b) to a vibrational replica of the most intense peak at 283.95 eV (blue curve), the latter assigned to the most abundant internal –CH2– carbon atoms. Similarly, up to 25% of the intensity at this energy (283.95 eV) is caused by the vibrational replica of the peak at 283.50 eV (ref. 40) (green curve), which is assigned to the methyl carbon atoms at the end of the chains (see section 2 of the ESI† for further details). Finally, we assigned the peak at 283.10 eV (and a small part of that at 283.50 eV due to vibrations) to the remnants of a denser monolayer phase, detectable at lower temperatures, that desorbs mainly between 240 and 280 K (see Fig. S3†).
Fig. 1c shows the high-resolution C-1s XPS spectrum of a dehydrogenated sample after flash-annealing at 375 K. In addition to the previously mentioned contributions, a new narrow component appears at 283.7 eV (grey curve in Fig. 1c). The fitted FWHM of the new component is 0.3 eV, which is narrower than the rest of the components (around 0.45 eV). We assign this component to the C–Pt bond at the methyl end of the chemisorbed molecules.47 Compared to the C-1s spectrum of physisorbed molecules, the intensity of the component associated with the methyl end (283.5 eV, green peak) is halved, leaving the ratio between the C–Pt bond and the methyl components at approximately 1
:
1. In addition, a small shift of 0.1 eV towards lower binding energies was observed upon chemisorption. The peaks associated with the internal carbon atoms (283.95 and 284.35 eV, solid and dashed blue curves in Fig. 1c) underwent minimal intensity changes as expected for the unaltered methylene-based backbone. The contribution of the component associated with the low-temperature layer remnants (283.10 eV) becomes negligible because desorption of this weakly bonded species is more efficient at higher temperatures. The almost complete disappearance of this component explains the 10% decrease in the global intensity of the C-1s XPS signal with respect to the sample annealed at 290 K.
C bonds makes the interaction possible between molecules and the subsequent aggregation of some of them.
To determine the nature of the structures observed in the STM images, we performed an analysis of the high-resolution XPS data. The C-1s XPS spectrum obtained after thermal treatment at 675 K (Fig. 1d) reveals the changes in the relative intensities of the different components. The contributions assigned to the methylene groups (–CH2–), at 283.95 eV and 284.35 eV (solid and dashed blue peaks in Fig. 1d), decreased by 47%, while the components at 283.4 eV and 283.7 eV (light green and grey curves) increased by 53% and 25%, respectively. We assigned the latter component to the C–Pt bonds, whose increase implies a larger interaction between carbon atoms and platinum, hindering adsorbate desorption even at temperatures higher than 675 K (see Fig. S3 and S5†). The increase of this component suggests that further dehydrogenation is taking place, which would result in the formation of more C–Pt bonds and eventually lead to possible structural changes of n-octane, such as fragmentation or branching of the molecule. Regarding the component at 283.4 eV, it has been reported that the XPS contribution corresponding to ethylene di-σ-bonded to the surface appears at this particular energy,48 so this component probably arises from σ-related interaction between molecules containing C
C and the Pt surface. DFT simulations show that a double dehydrogenation at the third and fourth carbon atoms is energetically favored, explaining the decrease in the intensity of the XPS peak assigned to –CH2– carbon atoms. A total intensity reduction of 20% was observed in the overall XPS C-1s signal with respect to the sample treated at 375 K, most likely due to the direct desorption of some chemisorbed molecules.
The first dehydrogenation step involves the exothermic activation of a methyl-end C–H bond with an energy barrier of 0.90 eV. The second dehydrogenation occurs at the third carbon atom in an endothermic reaction presenting a barrier of 1.11 eV followed by an endothermic C–H bond cleavage at the fourth carbon atom yielding a double bond between the third and fourth carbon atoms, with an energy barrier of 0.90 eV. The energy barriers of the effective dehydrogenation processes (Fig. 5c) are in good agreement with the activation energies extracted from the measured TPD spectrum (Fig. 1a). The activation energy for the main peak in TPD appearing at 320 K is 0.85 eV and the one associated to the shoulder appearing at 380 K is 1.06 eV (see Table S1†), compared with that of the calculated first transition state (TS1) of 0.9 eV. We attribute the main peak to molecules adsorbed on the terraces, while the shoulder would correspond to molecules adsorbed on defective surface regions (e.g., steps). At temperatures above 550 K, dehydrogenation continues. The second peak in the TPD spectrum at 600 K requires an activation energy of 1.74 eV and the corresponding shoulder of 2.02 eV. These values are consistent with simultaneous dehydrogenation at the third and fourth carbon atoms leaving two adsorbed hydrogen atoms, for which the activation energy can be estimated as the addition of the barriers calculated for the second and third dehydrogenation. The total intensity of the TPD contributions at 600 K and 700 K is just a third of that of the first peak, implying that not all the molecules doubly dehydrogenate following this path. Therefore, isomerization, cyclization, and aggregation processes without hydrogen desorption or desorption of hydrocarbon fragments may also contribute to the morphological changes observed by STM (Fig. 4a and b). We propose that the formation of unsaturated bonds, such as the double bonds formed in the second dehydrogenation stage above 600 K, is instrumental in cluster formation and further adsorbate aggregation, representing a preliminary step for the catalytic poisoning of the surface and coke formation.
C bonds that tend to cluster. We have modeled the TPD spectrum using population dynamics equations, which is in good agreement with the experimental activation energies and the energy barriers calculated by DFT. Our investigations of surface-catalyzed alkane dehydrogenation provide new insights into the atomistic mechanisms of such industrially relevant reactions.
Footnote |
| † Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d3nr02564k |
| This journal is © The Royal Society of Chemistry 2023 |