Open Access Article
Yang Wei†
a,
Yang Wang†a,
Yibo Changa,
Simone Lang
b,
Shiru Lin
*b and
Junwei Lucas Bao
*a
aDepartment of Chemistry, Boston College, Chestnut Hill, Massachusetts 02467, USA. E-mail: lucas.bao@bc.edu
bDepartment of Chemistry and Biochemistry, Texas Woman's University, Denton, Texas 76204, USA. E-mail: slin6@twu.edu
First published on 13th October 2025
Dry reforming of methane (DRM) both mitigates greenhouse gases (CH4, CO2) and produces syngas (H2, CO) for fuels and chemicals. Designing efficient DRM catalysts requires quantitative links between reaction barriers and adsorption/reaction energies. Traditional Brønsted–Evans–Polanyi (BEP) scaling, although widely assumed, may fail to predict quantitative kinetics, motivating the discovery of nonlinear relations that better couple thermodynamics and kinetics. Here, we investigate CH4 and CO2 decomposition on 12 doped Ni nanoparticles (M–Ni) using density functional theory. We show that BEP breakdown in DRM arises not only from binding-site preferences but also from the intrinsic electronic properties of the dopant, as analyzed by diabatic state calculations; kinetic modeling further amplifies these deviations. To uncover governing connections, we employ symbolic regression to derive physically interpretable expressions linking dopant descriptors to kinetic and thermodynamic parameters. Compared with linear scaling, these relations capture nonlinear effects, reveal hidden correlations, and provide predictive formulas that accelerate high-throughput screening. Validation on 12 previously unexplored doped nanoparticles confirms transferability: the best candidate, Mo–Ni, shows overall reaction energy and rate-determining barrier predictions within 0.05 eV of DFT. This integrated mechanistic and data-driven framework offers design rules for highly active, coke-resistant DRM nanocatalysts beyond BEP limitations.
The reaction involves the cleavage of strong C–H and C
O bonds, requiring the use of catalysts to lower the activation energy. By choosing suitable catalyst materials and optimizing their properties through doping, it is possible to enhance the catalytic activity and selectivity for CH4 and CO2 decomposition. Ni metallic nanoparticle is a commonly used catalyst in DRM due to its high activity and cost-effectiveness.2,3 The doping of other metals onto Ni (M–Ni) has been shown to improve catalytic performance and alleviate coke formation issues associated with Ni catalysts.4–7 Our previous studies have demonstrated the effectiveness of doping other metals onto Ni in enhancing catalytic performance and mitigating coke formation.8,9
Understanding the factors that affect the reaction barriers is crucial in the context of CH4 and CO2. This knowledge enables rapid assessment and facilitates accelerated catalyst screening with reduced computational demands. For an elementary reaction, the Brønsted–Evans–Polanyi (BEP) correlation (eqn (1)) directly establishes a linear correlation between the change in the reaction barrier (Eact) to the change in reaction energy, ΔErxn, for different surfaces:10–13
| Eact = αΔErxn + β | (1) |
However, the applicability of BEP relationships and volcano plots becomes limited in the case of doped nanoparticle catalysts. For example, Nwaokorie and Montemore illustrated the breakdown of linear scaling relations between adsorption energies and reaction barriers in steam reforming of methane that attempt to circumvent the traditional trade-offs imposed by these relations.16 On alloy surfaces, such deviations often arise from variations in coordination environments and site-specific bonding, which disrupt the uniform electronic behavior assumed in classical BEP scaling. For instance, intermediates like CH3 and C may adsorb on different atomic sites, each with distinct reactivity. In certain alloys, CH3 preferentially adsorbs on a specific metal dopant, while C interacts with multiple metal atoms, leading to stronger CH3 adsorption without over-stabilization of C.
A recent DFT study by Zhou et al. supports this observation through an investigation of C–H bond activation in methane over 12 single-atom alloy (SAA) catalysts with Cu and Ag as hosts.17 The authors reported substantial deviation from the linear BEP correlation and revealed that the breakdown is mainly due to the differentiated adsorption behaviors of the H atom and the CH3 group products across different SAAs.
These findings highlight the potential of engineering nanoparticle catalysts to break the linear BEP relationship, thereby offering a promising strategy to surpass the intrinsic activity limits defined by volcano plots. Nevertheless, the fundamental mechanistic origins of BEP breakdown remain insufficiently understood, warranting further systematic investigation. Given the complexity of reaction mechanisms and the interplay of surface phenomena, accurate prediction in such systems demands an advanced computational framework beyond conventional linear scaling.
Machine learning (ML) is increasingly utilized in materials science as a powerful tool for identifying quantitative relationships between reaction barriers and properties, accelerating materials design. However, conventional ML models are often criticized for their “black box” nature and overfitting, as they tend to generate highly complex mathematical expressions, making it challenging to derive straightforward physical interpretations from the model.
As an interpretable alternative, symbolic regression (SR) simultaneously identifies the optimal mathematical expression that describes underlying physical relationships, thus offering direct insights for materials design.18 For instance, Weng and co-workers used SR to derive the expression that is used to guide the design of oxide perovskite catalysts for the oxygen evolution reaction (OER).19 From a feature set containing electronic parameters as well as structural parameters of 23 selected oxide perovskites, SR identified that the octahedral factor over the tolerance factor provided the best balance between complexity and accuracy among 43
200
000 candidates. The descriptor exhibited a linear correlation with the potential VRHE, revealing that the activity of the OER is related to the structural parameters of catalysts. Based on these insights, five new oxide catalysts were synthesized, and four of them exhibited higher experimental activities than previously reported perovskite oxide catalysts.
In addition to SR, other models such as the least absolute shrinkage and selection operator (LASSO) regression20 have also been employed to derive interpretable correlations. For example, the existing linear scaling relationships between the adsorption energy of metal adatoms on various supports and the metal adatom's oxide formation enthalpy suggest that the interactions of adatoms with surface oxygen atoms largely determine the strength of their interfacial bond with the metal oxides. O'Connor and co-workers utilized LASSO regression to identify key property descriptors that predict the interaction strengths between 13 single metal atoms and 7 oxide supports.21 Their analysis revealed that the ratio of the oxide formation enthalpy of the metal adatom to the oxygen vacancy formation energy consistently appeared in the predicted models with low RMSE, indicating its strong correlation. This descriptor clearly suggests that metal atoms exhibit stronger binding to surfaces when they have higher oxide formation enthalpies and weaker binding to surfaces with increased oxygen vacancy formation energies.
A comparative study of multiple ML models was presented by Liu and co-workers, who aimed to predict the adsorption energies of CO on nine different transition metal planes.22 Four models, including adaptive boosting (Ada), decision tree (DT), LASSO regression, and random forest (RF), were trained with 20 selected features. Among them, RF exhibited the best performance with an R2 value of 0.885 and identified the valency of adsorbate and boiling point of the catalyst element as the most important features that affect the adsorption. Feature engineering, which involved identifying highly correlated feature pairs and removing low-importance features, only slightly improved the R2 of the RF model by 0.01. In comparison, the utilization of SR enabled the construction of a new composite feature derived from three less important variables (≥10 among all features), which are the surface energy of the substrates (Es), the coordination number of the adsorption sites (CN), and the interplanar crystal spacing (D). The resulting expression (CN + D)/Es replaced the original variables in the RF model and improved the R2 to 0.921, while the mean absolute error (MAE) decreased by 0.048 eV.
These examples highlight the potential of ML models for building accurate and interpretable expressions. In this study, we employed symbolic regression in Python (PySR)23 to derive mathematical expressions that reveal the underlying complex relationships between the target properties and the input features of M–Ni catalysts.
The features of 12 well-studied dopants, including Ni (the prinstine system), Cr, Fe, Mn, Co, Pt, Rh, Ru, Ti, V, Y, and Zn, obtained from quantum-level calculations, were used to train PySR models for regression, as shown in the dataset box in Fig. 1. PySR identifies one or two of the most important features and effectively uses simple formulas to correlate them with (i) the overall reaction energy of CH4 decomposition (excluding the coke formation step, i.e., the 4th step initiated with C(ads) formation), which is denoted as ΔE1–3(CH4): CH4(ads) → CH(ads) + 3H(ads), (ii) the rate-determining barriers for the CH4 decomposition (V‡rds(CH4)), (iii) the CO2 direct decomposition energy (ΔECO2), and (iv) its barriers
. The inner loop of PySR follows a classic evolutionary algorithm, where each population evolves through tournament selections, mutations, and crossovers for generating new formulas. The process begins with the generation of an initial population with selected features as variables, followed by the random selection from the population. In each tournament, the parent formulas are evaluated based on the fitness function, and the formula with the lowest loss undergoes a random crossover, which exchanges variables with other individuals, and mutation, which alters operators, to create offspring. The parents and the offspring then compete, and the formula that best balances the models' complexity and loss at a certain complexity is saved. If the termination criteria are met, the saved formula is printed along with its MAE and complexity. Otherwise, a copy of the winner formula replaces the weakest member in the original population, ensuring iterative improvement through selection and variation. Subsequently, the ranking metric, normalized discounted cumulative gain (nDCG),24 kinetic Monte Carlo (kMC) simulation, and the interpretation of the formula were used to assess the most meaningful formula generated by PySR. The selected formula is then chosen to screen 12 candidate dopants from untrained transition metals, including Sc, Mo, Zr, Nb, Tc, Pd, Cd, Hf, Ta, W, Os, and Ir, followed by the validation through DFT calculations.
We model and compare the reaction kinetics of DRM using DFT-calculated energy, PySR-predicted nonlinear correlation, and BEP-based linear correlation. By comparing the results from these approaches, we showcase the critical importance of accounting for nonlinear effects in scaling relations for accurate prediction of reaction kinetics and efficient catalyst design. The PySR framework not only enabled us to uncover the intricate correlations between these variables but also offers a predictive tool in catalysis research, reactor design, and the development of sustainable chemical processes.
For single-atom doped M–Ni supercell optimization consisting of 125 atoms in five layers, we used a 2 × 2 × 1 Monkhorst–Pack k-point mesh.31,32 All computations were spin-polarized with an energy cutoff of 400 eV for the plane wave basis set. Local minimum structure was achieved using a force convergence tolerance of 0.023 eV Å−1, while transition-state searches employed a tolerance of 0.05 eV Å−1. To prevent unphysical interactions between adjacent layers, the M–Ni surfaces were oriented in the xy-plane with the z-direction perpendicular to the layer plane, adopting a vacuum spacing of over 15 Å and applying a z-direction dipole correction.33
The adsorption energy of CH4 (ΔECH4(ads)) was calculated using eqn (2),
| ΔECH4(ads) = ECH4(ads on M–Ni) − EM–Ni − ECH4(GS,G) | (2) |
The CH4 dissociation energy (ΔECH4) is computed by
| ΔECH4 = ECH4(ads on M–Ni) − ECH3(ads on M–Ni) − EH(ads on M–Ni) | (3) |
Although the subsurface layers may contribute to reactivity,38,39 we chose to use cluster-based models for performing the CDFT-CI analysis, aiming to elucidate qualitative diabatic energy trends across dopants. To evaluate the validity of our cluster-based approach, single-point Kohn–Sham density functional theory (KS-DFT) calculations were performed using Q-Chem,40 employing the PBE-D3BJ functional with def2-TZVP basis set,41 to determine the most favorable spin state by using the barrier and reaction energies from periodic calculations as the reference. As shown in Tables S1 and S6, the trends predicted by KS-DFT using the top-layer cluster model are in strong agreement with those obtained from periodic VASP calculations. This consistency suggests that the top-layer cluster is sufficient to capture the key local electronic reorganizations responsible for BEP deviations. In the CDFT-CI calculations, the total spin multiplicity was fixed between the reactant state and the product states. Although the constraint eliminates the spin-crossing effects observed in the periodic calculations, the fixed-spin diabatic model provides a controlled framework to isolate the effect of geometric and electronic structure variations on the BEP correlation. DFT-D3(BJ) dispersion correction and def2-ECP effective core potential (ECP) were used to account for the effects on heavy atoms. Due to the small HOMO–LUMO gaps for the metal slab, the relaxed constraint algorithm (RCA)42 and level-shifting method43,44 were applied to facilitate SCF convergence. Population analyses were carried out using the Becke scheme to obtain atomic charges and spin densities, while natural bond orbital (NBO)45 analysis was conducted to further investigate changes in the valence orbital character along the reaction pathway.
The Becke charges and spins from the initial and final states along the reaction path were compared to compute atomic charge and spin differences. Based on these differences, the system was partitioned and reordered into three fragments, which are the donor region of the slab, the acceptor region of the slab, and the molecular moiety (CH4 or CH3). We tested alternative fragmentation schemes, including slab/dopant/CH4 and spin- or charge-constrained multi-atom partitions. These either failed to converge because the constraints were unstable or produced unphysical diabatic profiles without a crossing. Increasing the number of fragments also substantially raised computational cost and hindered convergence. We therefore chose the donor/acceptor/CH4 partitioning for its physical relevance and numerical stability.
The diabatic states were constructed by constraining the total charge and spin of the three fragments defined above using the Becke partitioning scheme, and fragment populations were taken directly from the KS-DFT results. The CI was performed between the two constrained diabatic states to compute state energies and the state couplings along the reaction path.
In each tournament, 15 formulas were considered, and the MAE was used as the loss function in this study:
![]() | (4) |
000 in this study.
The features selected for training the PySR in this study were based on the Pearson correlation coefficient between each pair of features from our previous work.9 These features along with their corresponding physical interpretations are summarized in Table 1.
| Features | Explanations |
|---|---|
| d | The number of d electrons in the dopant (M) |
| qM | The atomic Bader charges (in e) of the dopant in the clean surface of M–Ni (without adsorbates)46,47 |
| IEM | The first ionization energy of the dopant (in eV) |
| ΔEf(M–C) | Carbide formation energy of the dopant (in eV) |
| ΔEf(M–H) | Hydride formation energy of the dopant (in eV) |
| ΔEf(M–O) | Oxide formation energy of the dopant21,48 (in eV) |
The X–M formation energy ΔEf(M–X) (X = C, O, H) of the dopant metal, which indicates the thermodynamic feasibility for M to bind with atom X, can be calculated by
| ΔEf(M–X) = (EX–C − nXEX − nMEM)/(nX + nM) | (5) |
The values of all features across various dopants are provided in the SI (see Table S8) and were computed at the PBE/6-311+G** level of theory27,49,50 using the Gaussian 16 program.51
![]() | (6) |
![]() | (7) |
To solve the rate equations of the kinetic model, we employed the Gillespie kinetic Monte Carlo (kMC) method,52 revealing substantial discrepancies between linear and nonlinear correlation models. The rate constants associated with CH4 and CO2 dissociations are obtained by
![]() | (8) |
The kMC simulations were performed using GillesPy2,54 with a total simulation time of 10 ns, at which point the concentration of all species had reached steady state.
| M–Ni | ΔECH4(ads) | ΔECH3(ads) | ΔEC(ads) | ΔECH4(ads)−CH3(ads) | V‡rds |
|---|---|---|---|---|---|
| Ni | −0.20 | −2.27 | −7.08 | 2.07 | 0.78 |
| Cr | −0.35 | −2.52 | −7.19 | 2.17 | 0.73 |
| Fe | −0.22 | −2.27 | −7.06 | 2.05 | 0.87 |
| Mn | −0.25 | −2.30 | −6.88 | 2.05 | 0.84 |
| Co | −0.21 | −2.29 | −7.11 | 2.08 | 0.83 |
| Pt | −0.18 | −2.00 | −7.06 | 1.81 | 0.90 |
| Rh | −0.24 | −2.23 | −7.18 | 1.98 | 0.66 |
| Ru | −0.26 | −2.34 | −7.40 | 2.07 | 0.57 |
| Ti | −0.41 | −2.48 | −7.13 | 2.08 | 0.86 |
| V | −0.32 | −2.54 | −6.94 | 2.22 | 0.86 |
| Y | −0.35 | −2.09 | −7.31 | 1.73 | 1.43 |
| Zn | −0.18 | −2.24 | −7.07 | 2.06 | 1.30 |
| R2 | 0.0024 | 0.1745 | 0.0007 | 0.2689 |
Upon plotting V‡rds against ΔECH4(ads) and ΔECH3(ads) we observed a relatively compact clustering of points within the range of −0.41 eV to −0.18 eV for ΔECH4(ads) and −2.54 eV to −2.00 eV for ΔECH3(ads). We turned our attention to correlating the adsorption energies of ΔEC(ads) with V‡rds, which has been reported to obey the scaling relation on pure metal surfaces.16 Although the resulting distribution was relatively scattered from −7.40 eV to −6.88 eV, we found an even weaker linear correlation with an R2 value of 0.0007 eV (Fig. 2c). Given a poor correlation between individual intermediate adsorption energies and V‡rds, we explored the energy difference between ΔECH4(ads) and ΔECH3(ads) (ΔECH4(ads)–CH3(ads)). Yet this approach failed to distinguish between the different M–Ni alloys in terms of CH4 decomposition barriers, with an R2 value of 0.2689 (Fig. 2d). We further tested the square and cubic adsorption energy of ΔECH3(ads) and ΔECH4(ads)–CH3(ads) due to their relatively large R2. While the adsorption energy for M–Ni alloys began to separate more as the degree of these expressions increased, linear fitting remained inadequate, with the highest R2 value reaching only 0.2559 (Table S3).
We sought to identify a BEP correlation between the reaction energy and the V‡rds, as reported in Table S4, to assess the capability of a dopant to promote reaction activation. Our screening results reveal that dopants that show large deviation from BEP scaling in the rate-determining step (Fig. 3a) also exhibit significant deviation in the overall dissociation energy (Fig. 3b). Notably, in both absorption energy and reaction energy analysis, Ru, Rh, Y, and Zn show the largest deviation from the expected BEP-derived linear trend. Specifically, Ru and Rh are below the best-fitting line, indicating an enhanced catalytic activity with lower actual activation barriers than predicted by the BEP linear relations. These observations suggest that violations of the BEP relationship in the rate-determining step often coincide with deviations in the overall CH4 dissociation energy, indicating a possible coupling between kinetic and thermodynamic scaling violations.
As all the elementary-step dissociation barriers of CH4 were higher than those of CO2 (Table S2), this demonstrates that the dissociation of CH4 should be the RDS on M–Ni. The detailed information for CO2 is listed in the SI. Similarly to CH4 decomposition, no good correlation was observed for CO2 decomposition. Among the tested descriptors, CO adsorption energy showed the most notable degree of correlation, with an R2 value of only 0.58, indicating the complexity of this relationship, as shown in Table S5.
As shown in Fig. 4a–c, Mn, Fe, and Co show a similar structure for CHx upon CH4 dissociation, where at the reactant state, CH4 is located above the dopant (d1), and the CH3 fragment is tri-coordinated with two Ni atoms and one dopant atom at the product state. However, at the product state of Fe and Co, the H atom is bound to three Ni atoms with a distance of about 3.0 Å away from the dopant (d2), while for Mn, the H atom is tri-coordinated with two Ni atoms and one Mn atom at a distance of 2.91 Å. Despite the very close V‡rds among Mn (0.84 eV), Fe (0.87 eV), and Co (0.83 eV), notable geometric distinctions are observed. Interestingly, Rh, which exhibits a significantly lower V‡rds of 0.66 eV, displays a similar binding geometry and absorption site to Fe and Co, as shown in Fig. 4d. These observations suggest that the adsorption site geometry alone does not fully account for the violation of the BEP rules.
To further examine the factors underlying the bonding pattern variations, we performed CDFT-CI calculations on four systems. By imposing constraints on the wavefunction (spin and charge distribution), we constructed diabatic states to gain insight into the electronic structure. Along the same diabatic state, the wavefunction character remains unchanged over the entire reaction coordinate. Two diabatic states were generated: one by constraining the wavefunction at the reactant state while varying the geometry toward the product, and the other by constraining the wavefunction at the product state while varying the geometry from reactant to product. Adiabatic states, in contrast, fully relax the wavefunctions to their respective eigen-solutions and can thus be viewed as constructed from diabatic states. Generally speaking, the crossing point of the two diabatic states marks a significant change in wavefunction character at a specific geometry (an image) on its adiabatic state.
The resulting diabatic and adiabatic potential energy surfaces reveal distinct crossing behaviors depending on the dopant. For Mn, Fe, and Co (Fig. 5a–c), the diabatic crossing point appears slightly before or coincident with the ground-state adiabatic transition state (TS) at image 4, image 4, and image 3, respectively (the starting image is labeled as image 1). In contrast, for the off-line metal Rh (Fig. 5d), the adiabatic TS occurs at image 3, while the diabatic crossing point is shifted beyond the TS, appearing near image 4. A similar behavior was observed in Ni–Ni and Ru–Ni (Fig. S2), where Ni follows the BEP relationship while Ru deviates from it.
NBO analysis provides a complementary perspective on electronic structure changes along the reaction coordinate, revealing detailed variations in valence orbital occupations for selected atoms in the different doped-Ni systems (Table S7). For Mn–Ni, the NBO electron configuration of the dopant and the dissociated H atom exhibits abrupt occupancy changes at the TS (Table 3), with the dopant's 4s and 3d orbitals' occupations changing by ∼0.1 and ∼0.2, respectively, and the H 1s orbital increasing by ∼0.2. In contrast, the corresponding changes before the TS (image 3) and after the TS (image 5) are much smaller. This observation, consistent with the localized nature of 3d orbitals and the overlap of diabatic and adiabatic transition points, indicates that in these dopants the electronic configuration is highly sensitive to geometric distortions along the reaction pathway. A similar behavior is found for Fe and Co, where electron configuration changes at the TS are robust.
| Image 1 | Image 2 | Image 3 | Image 4 | Image 5 | Image 6 | Image 7 | Image 8 | Image 9 | ||
|---|---|---|---|---|---|---|---|---|---|---|
| Mn | Dopant | 4s0.88 | 4s0.86 | 4s0.83 | 4s0.70 | 4s0.70 | 4s0.66 | 4s0.64 | 4s0.62 | 4s0.62 |
| 3d5.82 | 3d5.86 | 3d5.91 | 3d6.09 | 3d5.99 | 3d5.98 | 3d5.92 | 3d6.00 | 3d5.91 | ||
| H | 1s0.78 | 1s0.78 | 1s0.80 | 1s1.02 | 1s1.20 | 1s1.20 | 1s1.20 | 1s1.17 | 1s1.20 | |
| Fe | Dopant | 4s0.90 | 4s0.89 | 4s0.85 | 4s0.71 | 4s0.70 | 4s0.66 | 4s0.65 | 4s0.74 | |
| 3d6.74 | 3d6.76 | 3d6.81 | 3d7.02 | 3d6.97 | 3d6.91 | 3d6.87 | 3d6.77 | |||
| H | 1s0.78 | 1s0.78 | 1s0.81 | 1s1.05 | 1s1.21 | 1s1.21 | 1s1.23 | 1s1.19 | ||
| Co | Dopant | 4s0.91 | 4s0.87 | 4s0.67 | 4s0.66 | 4s0.63 | 4s0.68 | 4s0.72 | ||
| 3d7.92 | 3d7.96 | 3d8.23 | 3d8.19 | 3d8.09 | 3d7.99 | 3d7.98 | ||||
| H | 1s0.78 | 1s0.80 | 1s1.01 | 1s1.21 | 1s1.22 | 1s1.18 | 1s1.19 | |||
| Rh | Dopant | 5s0.83 | 5s0.78 | 5s0.68 | 5s0.69 | 5s0.64 | 5s0.60 | 5s0.67 | ||
| 4d8.45 | 4d8.53 | 4d8.66 | 4d8.58 | 4d8.58 | 4d8.57 | 4d8.47 | ||||
| H | 1s0.78 | 1s0.79 | 1s0.90 | 1s1.14 | 1s1.15 | 1s1.18 | 1s1.20 |
In contrast, Rh shows a more gradual redistribution, with the dopant 4d orbital occupations evolving smoothly from 8.45, 8.53, 8.66, to 8.58 from image 1 to 4 (Table 3), reflecting greater delocalization. Notably, the dissociated H atom on Rh undergoes its most significant orbital occupancy change only after passing the adiabatic TS at image 4, as the 1s orbital increases from 0.90 to 1.14. This change is larger than the corresponding variations at the TS (0.113) and at image 5 (0.009) (Table S7). The delayed occurrence of this major electronic reorganization explains why the diabatic crossing is shifted beyond the adiabatic TS. These results underscore a key distinction: while 3d dopants exhibit geometry-driven electronic changes, 4d dopants display electronically driven reorganization. Together, these findings indicate that BEP deviations arise not only from structural differences but also from fundamental differences in the electronic nature and reactivity of the dopants.
| ΔE1–3 = −0.436 eV + 0.00687 eV/(0.596 − qM/e) | (9) |
The V‡rds of CH4 decomposition shows a linear correlation with the summation of IEM and qM, with a MAE value of 0.046 eV (Table 4). These two features also appear in other formulas with complexities 5 and 9 (Table S9), indicating that they are the most sufficient features in predicting the reaction barrier. The descriptor clearly illustrates that a lower IEM, facilitating electron transfers from the dopant to the antibonding orbital of the adsorbed species, and a more electron-rich dopant, which helps to stabilize the product states, promote C–H decomposition. Ru, Rh, and Cr were predicted as the top three candidates, which aligns with the DFT conclusion, as shown in Fig. 6b.
The complete ranking based on PySR equations is listed in Table S10. For ΔE1–3 and V‡rds, the predicted ranking generally aligns with the DFT result with nDCG values of 0.969 and 0.980. Therefore, the PySR provides a reliable prediction for screening promising DRM catalysis activity.
For completeness, we also examined the optimal equation for CO2 decomposition (Table S9). The CO2 decomposition energy is expressed by the equation:
![]() | (10) |
The barrier of CO2 decomposition is expressed by the equation:
![]() | (11) |
of Y and
of Zn show the largest deviation from the DFT-calculated values, with a difference of 0.30 eV and 0.62 eV, respectively (see Table S11). Moreover, BEP prediction identified the CH4 decomposition step (BEP-
) of Zn as the rate-determining step with a barrier of 0.84 eV, which is inconsistent with the DFT results. To test this, we calculated the reaction rate using (1) DFT-calculated (ground truth) adiabatic energy barriers of all steps, (2) DFT-calculated adiabatic energy barriers for non-RDSs, and for the RDS, the nonlinear correlation-derived barrier from PySR, and (3) BEP-predicted adiabatic energy barriers of all steps. For all approaches, the reaction energies of each elementary step were obtained from DFT calculations. Indeed, the underestimated BEP-predicted barriers for the RDS on Y–Ni and Zn–Ni significantly accelerate the decomposition process, resulting in an unrealistically high concentration of the H product, as illustrated by the dashed pink line in the upper panel of Fig. 7a and b. In contrast, the DFT and PySR results show nearly overlapping curves, indicating that PySR provides highly accurate predictions.
These differences span multiple orders of magnitude, which qualitatively affect our understanding of reaction kinetics and pathways. The choice between linear and nonlinear correlations not only impacts the predicted turnover rates but also drastically shifts the relative distribution of intermediates. Importantly, nonlinear correlation models align more closely with the actual DFT results, highlighting the critical role they play in accurately capturing the complexities of catalytic behavior and ensuring reliable kinetic predictions. This emphasizes the limitations of traditional linear approaches and underscores the need for more sophisticated modeling in catalytic studies.
We also perform DFT to calculate the ΔE1–3 of Ta, W, and Ir and the V‡rds of Nb to assess the predicted results due to their high predicted ranking in ΔE1–3 or V‡rds. Although Tc has the highest ranking in V‡rds, its radiation characteristics preclude its industrial application. As a result, the ΔE1–3 of Ta and Ir were found to be −0.4760 eV and −0.4700 eV, which differ by 0.081 eV and 0.047 eV from the predicted value. The V‡rds of Nb is 0.7667 eV, which differs by 0.049 eV from the predicted value.
For unexplored systems, the nonlinear model provides rapid screening to prioritize candidates, but DFT validation remains essential for final selection. The SR model generally agrees with DFT, although discrepancies, such as for W, warrant caution. The relatively large error for W likely reflects limitations of the current training dataset, which focuses on 4th- and 5th-period elements. As a 6th-period transition metal, W may exhibit features not fully captured by the model. Future work will expand the training dataset to include heavier and more diverse elements to improve generalizability.
Overall, the close agreement between PySR predictions and DFT results highlights the potential of symbolic regression in catalyst design, where precise predictions of reaction barriers are essential for optimizing catalytic activity.
Additionally, we utilized PySR to establish correlations between DFT-computed adsorption energies of intermediates on M–Ni and easily obtained physical properties of M–Ni (in separate models) with reaction barriers in the decomposition of CH4 and CO2. Furthermore, we employed PySR to make predictions for the remaining M–Ni systems and selected the most promising systems for further DFT calculations.
Establishing a quantitative and predictive scaling relationship between reaction barriers and reaction/adsorption energies in CH4 and CO2 direct decomposition has proven to be a challenging task. Therefore, using PySR to fit nonlinear relationships between various variables is essential. This approach not only enhances predictability but also offers a simpler and more straightforward method compared to black-box machine learning models with a high complexity. Moreover, we leveraged PySR formulas to predict the barriers for CH4 and CO2 direct decomposition in additional transition metal-doped Ni systems that are out of the training set. Concurrently, we identified the Mo–Ni alloys, with high ranking in both barrier and reaction energy, as the best candidate. DFT calculations were also performed to validate the nonlinear correlation results for other highly ranked doped M–Ni systems with an average error of less than 0.1 eV.
Our findings have significant implications for the design of more efficient DRM catalysts, the reduction of computational costs, and broader applications. These results contribute to advancing sustainable energy solutions, reducing greenhouse gas emissions, and benefiting various industries. By providing a more accurate and efficient method via automated nonlinear correlations for predicting reaction barriers, our work has the potential to accelerate the development of new catalysts and materials for energy-related applications that require a framework that goes beyond BEP linear relationships.
Supplementary information is available. See DOI: https://doi.org/10.1039/d5cy01066g.
Footnote |
| † Equal contribution. |
| This journal is © The Royal Society of Chemistry 2025 |