Floris van den Bosch,
Nick Gerrits
and
Jörg Meyer
*
Leiden Institute of Chemistry, Leiden University, PO Box 9502, Leiden, 2300 RA, The Netherlands. E-mail: f.van.den.bosch@lic.leidenuniv.nl; n.gerrits@lic.leidenuniv.nl; j.meyer@chem.leidenuniv.nl
First published on 11th July 2025
Plasma catalysis offers a promising alternative to current ammonia production processes, due to the combination of high selectivity of heterogenous catalysis and efficient activation of nitrogen in the plasma. However, the theoretical understanding of how various plasma processes contribute to efficiency improvements remains limited. The pioneering work of Metha et al. (Nat. Catal., 2018, 1, 269) extended the standard formulation of transition state theory by making it vibrational state-specific through the use of the Fridman–Macheret α model. The resulting microkinetic model accounted for vibrational contributions under the non-equilibrium conditions of a plasma reactor. In this work, we critically examine the prototypical chemical process of activated N2 reactivity on ruthenium through explicit rate coefficient calculations using state-of-the-art molecular dynamics, based on a potential energy surface previously validated against molecular beam experiments. Our findings reveal that vibrational activation is significantly more effective in promoting surface reactivity than predicted by the Fridman–Macheret α model, which fails to capture the full complexity of state-specific contributions. Furthermore, our calculations indicate that vibrational activation is also the primary driver of highly activated thermal catalytic reactions. These results provide a valuable benchmark to guide the development of future state-specific microkinetic models for heterogeneous and plasma catalysis.
Broader contextPre-activating reactants by a plasma holds great potential to boost heterogeneously catalysed processes, especially in scenarios where dissociative chemisorption (DC) reactions are rate-limiting. However, the fundamental atomic-scale understanding and quantification of this enhanced performance are still lacking. The pioneering work on plasma-enabled ammonia synthesis for DC of N2 introduced the Fridman–Macheret (FM) α model, which is widely used to extend standard micro-kinetic modelling to account for vibrational excitations in plasma catalysis. While this model is computationally inexpensive, it relies on assumptions based on gas-phase chemistry. In this article, we scrutinise the FM-α model for N2 dissociation on Ru(0001) using explicit molecular dynamics simulations on an accurate potential energy surface. We find that vibrational excitation plays a far more dominant role in N2 dissociation than the FM-α model predicts, and the reasons behind this discrepancy are analysed. This is a crucial step towards developing accurate effective models to replace the FM-α model. |
Vibrational excitation of N2 is known to increase the dissociative chemisorption (DC) rate.24–27 The ground-breaking work of Mehta et al.28 for plasma-enhanced ammonia production has made the DC of N2 a prototypical reaction and an important showcase for the importance of vibrational excitations in plasma catalysis.20 This notion has been reached based on simulations for the turnover frequency (TOF) for ammonia synthesis under realistic catalytic reaction conditions. The underlying microkinetic models (MKMs) rely on the Fridman–Macheret (FM) α model29 to quantify the effect of vibrational excitation on the reaction rate coefficient of the DC of N2.28,30 The FM α model is rooted in gas-phase reactions, where it models the vibrational contribution in diatom–atom reactions by computing the relative ratio of the forward and backward barrier heights. This model has become the workhorse approach for modelling vibrational excitation in plasma catalysis.31,32
Recently, questions have arisen regarding the relative importance of vibrational excitation compared to other plasma effects.21,33 For instance, the concentration of vibrationally excited molecules is typically much lower compared to the concentration of radicals in dielectric barrier discharges, which are the plasmas most commonly used in plasma catalysis.21 It is important to note that the FM α parameter cannot be directly measured experimentally. Instead, it requires models fitted to experimental data or theoretical calculations to validate its value. While a simple gas-phase reaction may be reasonably well described by a model like TST and the FM α, molecule–metal surface reactions typically involve significantly more complex potential energy surfaces (PESs) and necessitate high-dimensional models. The applicability of the FM α model to these complex reactions has already been questioned by Kedalo et al.34 for N2 + Ru(0001) based on comparison with MD simulations. While their study should be considered with caution (see Section 3.3 and Section S2.2 in the ESI†), they found that the effect of vibrational excitation exceeds the FM α model's predictions but is not as significant as the effect of translational energy. Recently, some of us have suggested that the FM α approach performs poorly, both qualitatively and quantitatively, for DC in general. This is based on an extensive analysis of theoretical molecular dynamics (MD) studies available in the literature.35 For instance, the effect of the vibrational excitation of polyatomic molecules on the reactivity depends on the specific vibrational mode(s) being excited.35–40 For methane, excitation of the vibrational stretch modes is more effective than exciting the bend modes.36–38 Similarly, exciting overtones generally leads to complex distributions of near-degenerate vibrational states that again yield considerably different reaction probabilities.39,40 Furthermore, the FM α model seems to significantly underestimate the effect of vibrational excitation for a variety of DC reactions.35 In late barrier systems, the bobsleigh effect can make molecules with high incidence energy deviate from the minimum energy path (MEP), and consequently they experience much higher barrier heights, lowering reactivity40–42 (also often referred to as Polanyi's rules43). Additionally, sterical hindrance in the MEP combined with dynamical effects can have a major influence on the effect of rotational and vibrational excitation of reactants.44 Unfortunately, the FM α model is not able to capture these effects, since it only considers the forward and backward reaction barrier heights of the ground state PES and the absolute vibrational energy.
MD approaches are able to capture these complex effects of vibrational excitation, offering an intriguing alternative to the FM α model.45 In this study, we explore how and why the FM α model and some simple extensions deviate considerably from full-scale MD simulations. For the latter, we build on our previous work that established an accurate MD model for N2 dissociation on a Ru(0001) surface, meticulously validated against the gold standard in gas-surface dynamics provided by molecule-beam experiments.11,12,46 This model employs quasi-classical trajectory (QCT) calculations using a machine-learned high-dimensional potential energy surface (PES) based on DFT calculations, including all relevant degrees of freedom of both the metal surface and the molecule. We utilise this model to compute the effect of vibrational excitation on the reaction probabilities and uncover a nontrivial relationship between vibrational excitation and reactivity. Crucially, we show that vibrational excitation plays a significantly more important role in plasma catalysis than previously anticipated, influencing the predicted TOF of ammonia in plasma catalysis. If the rate-controlling reaction step is highly activated, our findings suggest that the same can hold for conventional thermal catalysis. This could have serious implications for the modelling of heterogeneous catalysis; to the best of our knowledge, vibrational excitation is consistently neglected in MKMs for thermal catalysis. Finally, we analyse the reaction dynamics to elucidate the disparity in reactivity between conventional TST methods and our dynamics-based approach, concluding that dynamical effects play an important role in plasma catalysis.
(1) A ‘ground-state-only’ distribution (ν = 0), as typically used to model thermal catalysis.
(2) A Boltzmann distribution to account for the population of vibrational states in thermal equilibrium at a given vibrational temperature Tvib that is the same as the gas temperature Tgas.
(3) A Treanor distribution47
![]() | (1) |
Throughout this work, we use this MKM to compute relative TOFs of ammonia synthesis, which are independent of pressure. This highlights the fact that different reaction rate coefficients for the N2 dissociation step cause the same relative differences for the TOFs for both the industrial high-pressure (100 atm for the Haber–Bosch process) and the typical low-pressure conditions in plasmas (1 atm). We assume a vibrational temperature of Tvib = 3000 K for the plasma, like in the work of Mehta et al.28 As mentioned in Section 1, other plasma-induced effects (e.g., the presence of radicals) can also play a role, which we neglect here, in order to focus on the effect of vibrational excitation.
![]() | (2) |
In this equation, kB is the Boltzmann constant, Ea is the activation barrier, including the effects of zero-point energy, and A is a frequency factor related to the entropy barrier and pressure. All activation barriers Ea in the work of Mehta et al.28 are determined by scaling relationships (SR), i.e., using fitted trends between adsorption energy and reaction barriers across many metal surface species.50 Focussing on N2 dissociation in this work, we opt to use a more accurate activation barrier of ETSa = 1.83 eV on the basis of a DFT-based transition state (TS) search11,51 for this rate-limiting step. This is a substantially higher barrier than the ESRa = 1.48 eV barrier from the SR. For all other elementary reaction steps, we keep the SR-based values as determined by Mehta et al.28 for terraces on the Ru(0001) surface unchanged.
![]() | (3) |
Here, H is the Heaviside step function, and most importantly, the FM α quantifies how effectively the vibrational energy Eν (excluding zero-point energy) reduces the effective barrier height. The FM α model aims to enforce the Polanyi rules43—which state that, in a late-barrier system, vibrational excitation has a greater impact on reactivity—by linking the ratio between the forward and backward reaction barriers to the “lateness” of the barrier and, consequently, to α as follows:
![]() | (4) |
Here, the forward and backward reactions are the DC and desorption of N2, respectively, and can be obtained with quantum chemistry calculations, such as DFT, either directly through a transition state (TS) search, or indirectly using the SR. Using scaling relationship activation energies results in an FM αSR of 0.37 and using the DFT activation energy for the forward dissociation reaction (keeping the adsorption energy equal) a slightly larger αDFT = 0.39 is obtained. We refer to this rate calculation method as the ‘TST + FM’ level of theory, appending ‘@SR’ or ‘@TS’ when using the activation energy of N2 dissociation from scaling relationships or a DFT transition state search, respectively.
In the field of gas-surface dynamics, vibrational efficacies (VEs)
![]() | (5) |
kν = A〈Pν〉(Tgas), | (6) |
![]() | (7) |
In this integral, finc(Einc; Tgas) is the probability density of the 1D Maxwell–Boltzmann distribution for the incidence energy Einc according to the gas temperature Tgas given by:
![]() | (8) |
Here, , where m is the molecular mass of an N2 molecule and vinc is its center of mass's velocity component perpendicular to the surface. We calculate the N2 dissociation probability Pν(Einc) using MD simulations, as detailed in Section 2.3.3. Rate coefficients obtained with this method are referred to as the ‘MD’ level of theory in the following.
![]() | (9) |
Here, the total number of trajectories is the sum of dissociated, scattered and trapped MD trajectories, respectively, at each Einc and vibrational quantum number ν (only explicitly denoted on the left-hand side of eqn (9) for brevity).
We have compared different approaches to numerically evaluate the integral in eqn (7) using the results from our MD calculations (see Section S2 in the ESI† for details). The most robust method uses the carefully constructed fitting function
![]() | (10) |
The HDNNP was designed with an emphasis on configurations where the total initial energy of the N2 molecule (both translational and vibrational components) and the Ru surface does not exceed 15 eV. Consequently, the chance increases that the HDNNP needs to extrapolate towards parts of the PES that are less-well covered by the original DFT data set for MD trajectories at higher initial energies. We account for these uncertainties in ν(Einc) and propagate them in the form of 95% confidence bounds. Using the exact Clopper–Pearson method,65 we compute the 95% confidence intervals for the estimators. For the lower bound, we include the number of non-conclusive trajectories terminated because of extrapolation Nextr in the total as if they are non-reactive (N+total = Ndiss + Nscat + Ntrap + Nextr), such that the lower bound is from the 95% confidence interval of a binom(Ndiss, N+total) distribution. Likewise, for the upper bound, we also include the inconclusive extrapolative trajectories in the reactive trajectories N+diss = Ndiss + Nextr, such that the upper bound is from the 95% confidence interval of a binom(N+diss, N+total) distribution. This allows both the uncertainty in the estimator due to finite sampling and the outcome of the trajectories due to the extrapolation of the HDNNP to be accounted for (see Fig. S1 in the ESI†). These errors are dominated by the former (latter) at low (high) incidence energies. Uncertainties in the resulting reaction rate coefficients kν and TOFs from the MKM calculations are accounted for by distinct fits of eqn (10) to the lower bound and upper bound of the estimators, giving us effective 95% confidence intervals (p− and p+) for a given fit parameter p. Additionally, we account for the uncertainty introduced by the least-squares procedure σ± at the confidence bounds p± by expanding the individual bounds as follows:
![]() | (11) |
Here, z = 1.96 corresponds to 95% confidence intervals of the least-squares uncertainties and we assume that the fit uncertainty is independent of the input data uncertainty allowing us to sum the variances associated with the confidence intervals. This results in 3 sets of parameters for eqn (10) per vibrational state: a lower bound, an estimator and an upper bound (see Section S2.1 in the ESI†), which are used to compute the error bars for results at the ‘MD’ level of theory. Although the least-squares uncertainties are substantial for high vibrational quantum numbers, they ultimately hardly affect the uncertainties of the derived kν and TOFs.
![]() | (12) |
This method considers only a single point of the reaction probability curves, which needs to be chosen carefully. In previous work for N2 dissociation on Ru(0001), P has been chosen between 0.1 and 0.5, resulting in ≈ 1.6 for ν ≤ 3.10,12
Here, we extend the original approach by aligning and incorporating the reaction probability curves for different vibrational states over a broader range of incidence energies. To achieve this, we express the numerator in eqn (5) as a sum of shifts between the successive vibrational states, . The δEνinc values are obtained as optimal shifts by minimising the total square difference integrals
![]() | (13) |
![]() | (14) |
Finally, the mean value is obtained as the slope of a linear fit through the points (ΔEνvib and ΔEνinc). For the details of this procedure and estimation of its uncertainty, see Section S4 in the ESI.†
![]() | ||
Fig. 1 Relative turnover rates for ammonia production on Ru(0001) under industrial reaction conditions for different levels of theory using the ‘plasma-off’ result from the study by Mehta et al.28 as the reference. For each level of theory, we report the TOF for the vibrational ground state (Tvib = 0 K), a thermal gas (Tvib = 673 K) where the vibrational states are Boltzmann distributed, and a plasma (Tvib = 3000 K) where the vibrational states are Treanor distributed. Tgas is 673 K in all three cases. The levels of theory are (i) SR activation energies with the FM α model as described in ref. 28, (ii) DFT-calculated activation energy11,51 with the FM α model, (iii) DFT[-calculated] activation energies with the mean VE ![]() |
At this level of theory, Mehta et al.28 demonstrated that the plasma vibrational distribution produces a significantly higher TOF—four orders of magnitude greater—compared to the ground-state-only result. Here, we show that the TOF increases by only 3% when a thermal distribution of N2 vibrational states is considered, as opposed to only the ground state. In other words, TST + FM@SR predicts that vibrationally excited species contribute minimally to ammonia production under thermal conditions—consistent with common expectations. As mentioned above, the first refinement of the reaction rate (i.e., TST + FM@TS) involves computing the barrier height directly instead of relying on scaling relationships. In general, the TOFs for all vibrational distributions have decreased roughly by the same amount (almost 3 orders of magnitude), which is directly caused by the 0.35 eV increase in activation energy. Although this increase of Eforwarda also increases the FM α (|Eforwarda| ≫ Ebackwarda in eqn (4)), the vibrational enhancement of the reactivity is negligible in the ground state and for thermal conditions, and only slightly larger under plasma conditions compared to TST + FM@SR.
At the next higher level of theory, TST + @TS, we observe a qualitative shift, with the vibrational ground state and thermal distributions now differing by more than 5 orders of magnitude. This suggests that, when the FM α = 0.39 is replaced by the much higher
= 1.8, vibrationally excited species dominate the TOF in the MKM not only under plasma conditions but also under thermal conditions. Given the increased contributions of vibrationally excited N2, one might expect a larger relative difference in TOFs between thermal and plasma distributions. However, the effect remains similar, with the TOF increasing by 5 orders of magnitude compared to 4 orders of magnitude previously. In Section 3.2, we will show that, at higher vibrational levels, the reaction becomes effectively ‘barrierless’, leading to a saturation of the reaction rate.
Finally, at our highest level of theory—rate coefficients from MD—the difference between the ground state and thermal distributions is larger than in the FM α-based results but remains considerably smaller than the -based results. Conversely, the gap between the plasma distribution and the ground state or the thermal distribution is immense, spanning 10 orders of magnitude—a difference unlikely to be observed in experiments. However, we argue that the MD simulations capture a realistic trend, as discussed in Section 3.2. Instead, the likely overestimation of the population of highly excited vibrational states by the Treanor distribution, as also reported in ref. 30, may account for this discrepancy. This underscores how the partitioning of energy can have a substantial impact on the predicted reaction rates in plasma catalysis. Finally, the ground state TOF at the MD level of theory is significantly lower than at the TST-based levels, indicating a higher effective activation energy for N2 dissociation. In Section 3.4, we will show that this effective increase arises from dynamical effects.
![]() | ||
Fig. 2 Thermally averaged reaction probabilities per vibrational level at different levels of theory: (i) activation energies from the SR28 with the FM α model29 (red stars), (ii) activation energy for N2 dissociation from a DFT TS search11,51 with the FM α model (green triangles), (iii) the same activation energies with the mean VE from MD (orange squares), and (iv) reaction probability explicitly computed using MD. |
When the FM α is substituted with , the slope is notably increased, reaching 〈Pν〉 = 1 already for vibrational states ν ≥ 4. In other words, according to the FM +
@TS level of theory, N2 DC on Ru(0001) is effectively barrierless and thus is guaranteed to occur starting from the 4th vibrationally excited state. For both TST + FM@SR and TST + FM@TS, the first vibrationally excited state ν = 1 is only one order of magnitude more reactive than ν = 0, much less than the decrease in population between the two states for a thermal Boltzmann distribution. On the other hand, for TST +
, ν = 1 is four orders of magnitude more reactive, more than enough to allow this state dominate reactivity over the vibrational ground state despite its lower population. The individual contributions of each vibrational state to the total reaction probability is further discussed in Section S3 of the ESI.† All of this explains why the increase in TOF from a thermal to a plasma distribution in Fig. 1 for TST +
@TS is smaller than expected a priori: the reaction rate saturates at high vibrational quantum numbers, which is precisely where the plasma distribution differs most significantly from the thermal distribution in terms of vibrational state populations. Nevertheless, the substantial impact of the plasma on the ammonia TOF is maintained, as the total number of vibrationally excited species remains higher in the plasma distribution compared to the thermal one.
Finally, MD-based reaction probabilities are the only ones that strongly deviate from a linear dependence on the vibrational level. While they initially exhibit a slope similar to , this slope gradually decreases with increasing vibrational quantum number, appearing to asymptotically approach 〈Pν〉 = 1. This change in slope explains the smaller difference between the TOFs for the ground state and thermal distribution, as only the first few vibrational quantum numbers contribute significantly to the N2 dissociation rate (see Section S3 in the ESI†). The fact that the TOFs in the vibrational ground state at the TST + FM@TS and TST +
@TS levels of theory are more than three orders of magnitude higher than those from MD is a direct consequence of the corresponding decrease in reaction probability at ν = 0 by a similar amount.
Going beyond the alignment of entire reaction curves, our molecular dynamics simulations enable us to delve deeper into the dependence of ην on the incidence energy according to eqn (5). This analysis allows us to examine the vibrational efficiencies in greater detail. Fig. 3a illustrates that the ν-specific vibrational efficacies are indeed significantly smaller at low incidence energies, exhibiting an increase with each vibrational level. Notably, ην(Einc) rapidly increases until approximately Einc ≈ 3 eV, where they attain a steady value of ην ≈ 2 for each vibrational quantum number. This observation suggests that the thermally averaged state-specific vibrational efficacies
![]() | (15) |
![]() | ||
Fig. 3 (a) VE as a function of incidence energy per vibrational state, with the inset zooming into 0 eV ≤ Einc ≤ 0.5 eV. (b) Comparison of the vibrational state independent FM αFM (eqn (4), orange line) and mean VE ![]() |
Previously, Kedalo et al.34 proposed an approach to extract FM α values from the reaction probability curves generated through MD simulations. This involves solving for α in eqn (3), which yields
![]() | (16) |
We have included the results from ref. 34 in Fig. 3(b) (black crosses) for αMDν(Tgas = 1000 K) for N2 dissociation on Ru(0001). In addition, we have also applied eqn (16) to the thermally averaged reaction probability obtained from our own MD simulations according to eqn (7) for the same temperature (orange circles in Fig. 3b). The results for αMDν(Tgas = 1000 K) from the two different sets of calculations differ quite notably from each other. Most importantly, as ν increases, our results do not approach αFM but rather converge to zero. We think that systematic shortcomings in the methodology of Kedalo et al.34 are causing these differences: firstly, their barrier height for N2 dissociation on Ru(0001) computed with the PBE density functional (1.9 eV) is larger than our value underlying our HDNNP computed with the RPBE density functional (1.84 eV11). This is in large disagreement with the SBH17 database,51 where the PBE and RPBE barrier heights have been computed as 1.49 eV and 1.95 eV, respectively. The values from the SBH17 follow the common trend that RPBE yields larger barrier heights for molecule–metal surface reactions than PBE. The small deviation of the aforementioned RPBE values (0.1 eV) is caused by slight differences in the DFT setup used in both studies, whereas the much larger difference between the PBE values (0.4 eV) is clearly beyond this error margin. Secondly, the root mean square error (RMSE) for the forces of the neural network fit for the potential energy surface reported by Kedalo et al.34 (0.2–0.4 eV Å−1) is an order of magnitude larger than what is usual (0.02–0.03 eV Å−1).40 Thirdly, in their MD simulations, Kedalo et al.34 shift the Maxwell–Boltzmann distribution for the velocities to considerably higher energies, effectively assuming that 80% of the total (translational) velocity of N2 molecules is distributed into the component perpendicular to the surface (see Fig. S4 in the ESI†). Furthermore, their MD simulation yields higher N2 dissociation probabilities for the flat Ru(0001) surface than that for the stepped Ru(113) surface—opposite to what has been obtained in other computational work.7,66,67
In light of these concerns regarding the DFT calculations, machine-learned potential and MD simulations of Kedalo et al.,34 we now focus our comparison of αMDν with 〈ην〉 on the results obtained from our own MD simulations (orange circles and blue squares in Fig. 3(b)). Starting from αMD1 ≈ 1.4, there is a monotonous decrease with vibrational quantum number ν, which still leaves αMD2 significantly larger than the FM α given by eqn (4). The fundamentally opposite trends for 〈ην〉 and αMDν with respect to ν arise from the different order of operations. For 〈ην〉, one first calculates the difference between the reaction probability curves and then integrates this difference over the velocity distribution. The definition of 〈ην〉 in eqn (15) consistently generalises the standard definition of vibrational efficacies used in gas-surface dynamics, completely analogous to 〈Pν〉 in eqn (7). However, we do not consider 〈ην〉 as a suitable substitute to introduce ν-dependence to the FM α, as this would effectively double-count statistical effects in the TST and MD descriptions. On the other hand, for αMDν, one first integrates Pν(Einc) over the Maxwell–Boltzmann distribution and then calculates the difference between the thermally averaged reaction probabilities 〈Pν〉. Consequently, vibrational-state-related effects are not blended with a combination of (different) statistical treatments at the TST and MD levels. By construction, if the results for αMDν were to be substituted for the constant in eqn (3), the resulting TOFs would closely align with the MD reference results shown in Fig. 1. Clearly, αMD1(1000 K) > 1 is the key enhancement that cannot be achieved with the FM α model.
Altogether, we present compelling evidence that capturing the impact of vibrational excitation on the reaction rate coefficient using the FM α approach, which currently stands as the state-of-the-art in plasma catalysis modelling, presents substantial challenges. According to the Polanyi rules,43 this effect can be estimated based on the potential energy surface. Early (late) barriers typically exhibit low (high) vibrational enhancement. Therefore, we anticipate similar implications for the efficiency of plasma excitation for other highly activated late barrier reactions, such as CH4,40,68,69 HCl,44,70 and CH3OH,71,72 where high VEs have also been observed.
To analyse these dynamical effects, we extracted statistics from MD trajectories of dissociated N2 molecules in the vibrational ground state at a high incidence energy (Einc = 3.25 eV) above the barrier along the MEP on our PES (ETSa = 1.83 eV). Even at such a high incidence energy the dissociation probability obtained from the MD simulations is less than 100% (see Section S1, ESI†). In other words, not all molecules are efficiently steered towards the MEP and can thus encounter barriers higher than ETSa. In our statistical analysis, we focus on the coordinates depicted in the insets of Fig. 4 that are traditionally not included in the reaction coordinate: the angular orientation of the molecular axis of N2 by the polar angle θ and its lateral position (x and y) relative to the surface unit cell, all based on the N2 centre of mass as the origin. As described in Section 2.3.2, our trajectory ensemble is initially uniformly distributed over both orientations and lateral positions. According to TST, all reactive trajectories will cross the same minimal energy barrier along the MEP, which implies that there should be no preferential initial conditions. In Fig. 4, we show the distributions for the initial orientation θ and xy position of the subset of dissociated N2 molecules, which have been rescaled such that their maxima coincide with those of the distributions for all trajectories. A sharper distribution for reacted trajectories indicates a degree of inaccessibility of the MEP for initial conditions further from the MEP, leading to a lower reaction probability compared to TST. By computing the ratio between the integrals over these (rescaled) distributions and their original counterparts for all trajectories, we can estimate the reduction in reactivity in particular degrees of freedom due to dynamical effects.
![]() | ||
Fig. 4 Distributions for initial (a) orientation θ and (b) xy position of N2 molecules above the Ru(0001) surface for Einc = 3.25 eV (>ETSa = 1.83 eV) and ν = 0. The coordinates are depicted in the insets. (a) The initial distribution for all trajectories (both reacted and non-reacted, black dashed line, eqn (S6) in the ESI†) is compared to the distribution of the subset comprised by the reacted trajectories (blue bars). To guide the eye, the latter has been fitted to the initial distribution multiplied by a Gaussian centred at θ = 90° (orange line, eqn (S7) in the ESI†). (b) The distribution for the initial xy-position of reacted trajectories in the irreducible wedge of the unit cell, as depicted in the inset. a denotes the surface lattice constant. Individual reacted trajectories are in red, the histogram is in greyscale, and the fitted bimodal Gaussian distribution is the contour lines. For better comparison and to quantify dynamical steering effects, the reacted distributions have been rescaled such that their maxima coincide with those of the distributions for all trajectories. |
As shown in Fig. 4a and b, respectively, both the θ and xy distributions for the reacted molecules deviate considerably from the uniform distribution of all molecules. The very narrow distributions in Fig. 4(b) illustrate that dissociation dominantly occurs for molecules that are initially placed over the bridge site, in agreement with previous findings.10,11 This significant narrowing of the distribution for xy due to surface corrugation leads to a substantial reduction of reactivity by more than a factor of 100. Fig. 4a shows that the distribution for θ narrows similarly for the reacted molecules, but the reactivity reduction in this case is only a factor of 2–3×. In the ESI,† we show that the effect of the azimuthal orientation of the molecular axis on the reactivity is even smaller (see Fig. S7, ESI†). There we also demonstrate that xy-related reactivity reduction strongly depends on the incidence energy and ranges from a 50× at low incidence energy to a 10× reduction at high incidence energy.
Altogether, assuming that the aforementioned angular and translational degrees of freedom are independent, the total reduction in reactivity due to dynamical effects is a factor of 20–250× compared to TST, depending on the incidence energy (see Section S5 in the ESI†). This accounts for more than half of the 103–104× reduction in the reaction probability shown in Fig. 2. Recrossing events in our MD simulations, i.e., molecules which do not dissociate despite crossing the barrier, could be responsible for a further reduction of the reaction probability compared to TST. Our recrossing analysis reveals that this is not the case (see Section S6 in the ESI†). Consequently, the bobsleigh effect is very likely responsible for the remaining reduction in reactivity due to dynamical effects, since it is closely related to the movement along the remaining molecular degree of freedom, the z-coordinate.40–42 In principle, MEP deviations and barrier recrossing could be accounted for in TST by including entropic contributions and transmission factors in the reaction rate coefficients for the vibrational ground. However, accounting for these effects requires additional computational efforts, which is why we have not further investigated this here.
For plasma catalysis, we found that the effects of vibrational excitation are much more pronounced than previously anticipated based on the basis of the FM α model. Some of us have recently observed that, in general, the vibrational efficacies for DC reactions scale with the absolute barrier height.35 In comparison to other metal surfaces, Ru(0001) has a relatively low barrier height for the DC of N2. Other metals thus likely exhibit even higher vibrational efficacies than described by the FM α model. Consequently, the predicted optimal catalyst and operating conditions might change considerably, depending on the level of theory employed to account for the vibrational enhancement of DC reactions.
Similarly, it is often unclear which kind of plasma-induced effects are dominant.20 Using different plasma catalytic models that try to disentangle the various effects might lead to qualitatively different conclusions. For example, the Langmuir–Rideal reaction mechanism (also, incorrectly, known as Eley–Rideal75), where a reaction takes place directly between a gas phase reactant (in plasma catalysis typically a radical) and a surface adsorbate, is often studied along vibrationally excited reactants. For the plasma-catalytic non-oxidative coupling of methane, it has been predicted that the importance of radicals and vibrational excitation of reactants depends on the binding strength of the catalyst surface.30,76,77 Similarly, for CO2 hydrogenation, radicals were predicted to be much more important.31 In a kinetic investigation of ammonia synthesis on Fe, adsorption of radicals, Langmuir–Rideal reactions, and vibrational excitation were all found to be important.73 Bayer et al.,33 on the other hand, concluded on the basis of a combination of simulations and experiments that the loss of vibrationally excited N2 on Fe catalysts was due to vibrational relaxation (i.e., non-reactive energy loss), and not DC. Furthermore, for dry reforming of methane on Ni, it was predicted that the Langmuir–Rideal reaction mechanism dominates, compared to vibrational excitation.32 Generally, it has been concluded that Langmuir–Rideal reactions and adsorption of radicals are much more important than vibrational excitation of reactants. However, in all cases, the FM α approach was employed. Based on our work, it can be expected that vibrational excitation plays a significantly more important role than has previously been accounted for by catalytic models. Furthermore, it is likely that the competing Langmuir–Rideal reaction mechanism occurs less often than typically modelled, especially when dynamical effects are considered.78 This could have positive implications for the practical application of plasma catalysis. As discussed above, it is often believed that radicals are the driving force behind the enhanced chemical reactivity. However, designing a catalyst and a reactor in such a way that radicals are not ‘quenched’, which reduces the availability of radicals for desired reaction steps, is not facile. But if vibrational excitation is much more important than thus far thought, practical plasma catalysis might be more easily achieved. In short, the way dynamical effects are modelled in plasma catalysis can alter predictions considerably.
In this work, we have also observed that scaling relationships can lead to a considerable underestimation of the barrier height, necessitating explicit TS search calculations instead. Similarly, it can be expected that, for the majority of catalytically important reactions, standard DFT calculations will underestimate the barrier height, even when the TS is directly computed instead of employing the SR.79 Self-interaction errors are likely the main culprit and are most prominent when the charge transfer at the TS is large.80,81 Future improvements in electronic structure calculations are crucial for obtaining more reliable PESs, including barrier heights for MKMs. Fortunately, for N2 + Ru(0001), the employed PES yields very good agreement with currently available experimental data from well-defined molecular beam experiments.11,46
We also acknowledge that the employed QCT method may have certain limitations, such as neglecting nuclear quantum effects. Although tunnelling is likely insignificant for a molecule as heavy as N2, zero-point energy violation does occur in our simulations (see Fig. S5 in the ESI†), leading to a slight overestimation of the reaction rate of lower vibrational states. In the future, by substituting the QCT approach with ring polymer MD, nuclear quantum effects can be simulated approximately, while keeping the simulations tractable.82 For the DC of methane on a moving Pt(111) surface, this approach has been demonstrated to accurately reproduce results from experiments, even for low incidence energies and high vibrational temperatures.83 However, to the best of our knowledge, ring polymer MD simulations cannot yet be performed for specific vibrational states, necessitating methodological advancements to simulate vibrational state populations found in plasma catalysis.
Despite all methodological challenges, dynamical effects remain a crucial important aspect in determining reaction rates of molecules on metal surfaces. For instance, in the case of heterogeneous catalytic cracking of ammonia on Fe surfaces, which is the reverse of the Haber–Bosch process investigated in this work, dynamical effects play a significant role not only in the diffusion coefficient of ammonia but also in reaction steps involving surface adsorbates.84 Therefore, we conclude that dynamical effects cannot be ignored in determining the reaction rates of molecules on metal surfaces. In the context of plasma catalysis, it appears that dynamical effects are even more important than in thermal catalysis, which unfortunately cannot be fully captured by TST models (at present). Accurate MD calculations for gas-surface dynamics are challenging to perform, as evidenced by the discrepancies between the present work and the work by Kedalo et al.34 (see Section 3.3 and Section S2.2 in the ESI†). In the meantime, if available in the literature, vibrational state-resolved reaction probabilities measured or calculated for molecular beams can be utilised to guide catalytic modelling. Unfortunately, accurate data are scarce, particularly for higher vibrational states and very low reaction probabilities.
Accurate determination of the vibrational enhancement of DC reactions, as exemplified in this work, requires the computation of numerous reaction probability curves, each requiring many trajectories. Therefore, the remaining discrepancy between the most advanced TST model and the MD simulations for N2 dissociation on Ru(0001) prompts further research to develop computationally efficient alternatives to account for the impact of vibrational excitation and dynamical effects on reaction rate coefficients. Although the mean VE does improve the FM α model by reproducing at least the initial trend of the MD results, there is room for improvement as the mean VE does overestimate the effect of vibrational excitations. The sudden vector projection model85 is an alternative to performing full dimensional MD simulations, coming at a much lower computational cost. Unfortunately, this model still struggles with the same assumption as the FM α model that vibrational energy cannot be more effective than translational energy in promoting reactivity, and that dynamical effects such as the bobsleigh effect are not explicitly accounted for. Nevertheless, some of us have recently shown that a combination of the absolute barrier height, equilibrium and transition state bond lengths, and the sudden vector projection model manages to reproduce VEs of various molecule–metal surface reactions reasonably well at a similar computational cost as the FM α model.35 Furthermore, it was observed that the curvature features of the MEP could significantly enhance the prediction of VEs with simplified static models.
Even under typical thermal catalytic conditions, the small fraction of vibrationally excited N2 molecules dominate the rate coefficient for N2 dissociation due to the very high vibrational efficacy. Consequently, typical MKMs for thermal catalysis, which only consider the vibrational ground state, prove insufficient for a highly activated chemical process like ammonia synthesis. Most importantly, we predict that the reactivity gap between thermal catalysis and plasma catalysis due to vibrational excitation is substantially larger than previously anticipated. MKMs used in plasma catalysis do commonly account for vibrational excitation with the FM α approach, which unfortunately leads to a large underestimation of the reactivity of vibrationally excited N2. Employing vibrational efficacies determined from MD simulations yields a much larger impact of vibrational excitation on the reaction rate and improves the agreement between the TST model and the MD approach. Complex features from the PES, such as corrugation and anisotropy, along with dynamical effects like the bobsleigh effect, are responsible for the remaining discrepancy between the two approaches. It remains to be seen if a simple and computationally efficient modification of the Arrhenius equation can account for these effects. Our results emphasise the significance of how energy is partitioned and modelled in a chemical system, as it can significantly alter predictions. Since the energy distributions in plasma catalysis can be highly out of equilibrium, it is probable that dynamical effects play a substantial role not only in the DC of vibrationally excited reactants but also in other plasma-induced effects (e.g., Langmuir–Rideal reactions). For the time being, we conclude that the gold standard for determining the reaction rate coefficients for DC reactions in heterogeneous catalysis and plasma catalysis involves performing dynamical simulations.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d5ey00132c |
This journal is © The Royal Society of Chemistry 2025 |