Pair your accounts.

Export articles to Mendeley

Get article recommendations from ACS based on references in your Mendeley library.

Pair your accounts.

Export articles to Mendeley

Get article recommendations from ACS based on references in your Mendeley library.

You’ve supercharged your research process with ACS and Mendeley!

STEP 1:
Click to create an ACS ID

Please note: If you switch to a different device, you may be asked to login again with only your ACS ID.

Please note: If you switch to a different device, you may be asked to login again with only your ACS ID.

Please note: If you switch to a different device, you may be asked to login again with only your ACS ID.

MENDELEY PAIRING EXPIRED
Your Mendeley pairing has expired. Please reconnect
ACS Publications. Most Trusted. Most Cited. Most Read
My Activity
CONTENT TYPES

Figure 1Loading Img
RETURN TO ARTICLES ASAPPREVQuantum Electronic S...Quantum Electronic StructureNEXT

Strategies to Obtain Reliable Energy Landscapes from Embedded Multireference Correlated Wavefunction Methods for Surface Reactions

Cite this: J. Chem. Theory Comput. 2024, XXXX, XXX, XXX-XXX
Publication Date (Web):July 14, 2024
https://doi.org/10.1021/acs.jctc.4c00558
© 2024 American Chemical Society
  • Subscribed

Article Views

1

Altmetric

-

Citations

-
LEARN ABOUT THESE METRICS
PDF (5 MB)
Supporting Info (2)»

Abstract

Embedded correlated wavefunction (ECW) theory is a powerful tool for studying ground- and excited-state reaction mechanisms and associated energetics in heterogeneous catalysis. Several factors are important to obtaining reliable ECW energies, critically the construction of consistent active spaces (ASs) along reaction pathways when using a multireference correlated wavefunction (CW) method that relies on a subset of orbital spaces in the configuration interaction expansion to account for static electron correlation, e.g., complete AS self-consistent field theory, in addition to the adequate partitioning of the system into a cluster and environment, as well as the choice of a suitable basis set and number of states included in excited-state simulations. Here, we conducted a series of systematic studies to develop best-practice guidelines for ground- and excited-state ECW theory simulations, utilizing the decomposition of NH3 on Pd(111) as an example. We determine that ECW theory results are relatively insensitive to cluster size, the aug-cc-pVDZ basis set provides an adequate compromise between computational complexity and accuracy, and that a fixed-clean-surface approximation holds well for the derivation of the embedding potential. Additionally, we demonstrate that a merging approach, which involves generating ASs from the molecular fragments at each configuration, is preferable to a creeping approach, which utilizes ASs from adjacent structures as an initial guess, for the generation of consistent potential energy curves involving open-d-shell metal surfaces, and, finally, we show that it is essential to include bands of excited states in their entirety when simulating excited-state reaction pathways.

This publication is licensed under the terms of your institutional subscription. Request reuse permissions.

1. Introduction

ARTICLE SECTIONS
Jump To

Traditional correlated wavefunction (CW)-based methods aimed at resolving the electronic structure in strongly correlated systems commonly rely on complete active space self-consistent field (CASSCF) wavefunctions to describe a system’s multiconfigurational character. (1) The first step in such calculations, generally, is the selection of an orbital active space (AS), which, due to the limitations placed upon its size by the exponential scaling of the multiconfigurational wavefunction, commonly relies on approaches based upon chemical intuition and trial-and-error. While some progress has been made in recent years in the development of black box and automated AS selection schemes for CASSCF and other multireference methods, including the use of ASs based on natural orbital (NO) occupation numbers (NOONs), from unrestricted Hartree–Fock (UHF), (2,3) second-order Møller–Plesset perturbation theory, (4) partially converged density-matrix renormalization group, (5) and N-electron valence state theory wavefunctions, (6) as well as automated selection schemes, e.g., based on atomic valence AS, (7) orbital entanglement entropies, (8) or machine learning, (9,10) CASSCF calculations remain far from being black box. More specifically, difficulties arise when constructing ASs for transition metal clusters where meaningfully localized molecular orbitals (LMOs) cannot be obtained using localization schemes designed for molecules due to the large number of orbitals with significant partial occupancy (occupation numbers between 0.1 and 1.9).
Generating a consistent AS along reaction pathways (11,12) is also difficult as the optimal ASs of different atomic structures are not always the same; some active orbitals that are important in a reactant might be insignificant in the product and vice versa. To construct a consistent AS of common size for all structures along the reaction pathway, it is desirable to include all important orbitals for all structures. However, the expansion of the AS along the reaction pathway can quickly become intractable when the reaction is complicated, i.e., involving many degrees of freedom, or when the system is large. As an example, three approaches were used in ref (11), which looked at O–O bond formation from water, catalyzed by a ruthenium (Ru) complex using CASSCF: (i) AS construction for each structure based on their LMOs; (ii) all structures along the reaction path use the same initial guess of active orbitals derived from the transition state (TS); and (iii) from TS to the reactant and from TS to the product, the calculation of a structure along the path (frame) always takes the AS of the previous or next frame as the initial guess (henceforth referred to as the “creeping” strategy), with all three approaches yielding identical results. Examples of successful implementation of the “creeping” strategy include study of ground- and excited-state heterogeneous catalysis in the context of plasmonics for various chemistries and surfaces, e.g., H2 dissociation on Au(111), (13−15) N2 dissociation on Fe- and Mo-doped Au(111), (16,17) NH3 decomposition on Fe- and Ru-doped Cu(111), (18−20) CH4 decomposition on pure and Ru-doped Cu(111), (21) and H2 desorption from Pd(111). (22)
The approach taken in our group for embedded correlated wavefunction (ECW) theory simulations (23−25) of heterogeneously catalyzed reactions generally utilizes NOONs based on unrestricted Kohn–Sham (UKS) density functional theory (DFT) to select orbitals on metal clusters. Toth and Pulay discussed the advantages and disadvantages of the unrestricted NO criterion, where the NOs and their occupation numbers from UHF were used to select active orbitals. (26) Because HF can give qualitatively incorrect descriptions for metals, such as a spuriously vanishing density of states around the Fermi level (27) and incorrect spin and charge density waves, (28) we use the one-electron wavefunctions from UKS instead of UHF as the initial guess for CASSCF calculations. When considering reaction pathways, we obtain the CASSCF orbitals of the isolated adsorbate (closed or open shell) at its geometry as optimized along the reaction path. We then concatenate the UKS orbitals of the isolated metal cluster (structurally fixed in most cases) and the CASSCF NOs of the isolated adsorbate and use the merged orbitals as the initial guess for the embedded CASSCF (emb-CASSCF) wavefunction of metal–adsorbate complexes. This merging approach, which has previously been offered as an alternative to creeping when the latter fails, involves generation of the guess orbitals from the optimized orbitals of each fragment in the system. Merging can improve CASSCF convergence, avoid artificial symmetry breaking, enforce convergence toward the desired states, and prevent unwanted orbitals from entering the AS, among other advantages. For example, Klüner et al. first employed a merging approach to treat the excited states of CO absorbed on Pd(111), (29,30) Hirano and Nagashima used merging to enforce certain symmetry constraints in FeCO molecules, (31) and Cimpoesu et al. used merging to generate good initial guesses for lanthanide complexes. (32)
In this work, we explored why creeping fails to provide consistent ASs for a particularly challenging case, ammonia (NH3) decomposition on palladium(111) (hereafter denoted as Pd(111) + NH3), followed by presentation of guidelines for obtaining reliable energy landscapes using the ECW theory for surface reactions more generally. We initially studied the effects arising from basis set size, basis set superposition error (BSSE), embedded cluster size, and the manner in which the embedding potential is derived (whether from the clean Pd surface or from surfaces optimized in the presence of adsorbates) and find that these factors are not the reasons for discontinuous potential energy curves (PECs). We narrow down the source of error to the inconsistency in the CASSCF AS. We generated smooth PECs from embedded complete AS second-order perturbation theory (33) (emb-CASPT2) by constructing ASs using merging for the preceding emb-CASSCF, which provided the reference wavefunctions for emb-CASPT2. We then compare the NOs in the ASs between merging and creeping to understand why the latter fails for this particular system. Finally, in light of recent success in utilizing ECW theory simulations to resolve photocatalytically and plasmonically driven processes, (34) we also studied the effect of the number of states included in embedded state-averaged CASSCF (emb-SA-CASSCF) calculations on excited-state landscapes.

2. Methods

ARTICLE SECTIONS
Jump To

2.1. Periodic DFT

We performed periodic Kohn–Sham DFT calculations using the Vienna Ab initio Simulation Package (VASP), version 5.4.4. (35−37) We employed standard VASP projector augmented-wave potentials, self-consistently solving for the valence electrons, namely, the H 1s, N 2s,2p, and Pd 4d,5s orbitals, with a kinetic energy cutoff of 700 eV for the plane-wave (PW) basis set, in combination with the Perdew–Burke–Ernzerhof (PBE) generalized gradient approximation exchange–correlation (XC) density functional. (38) Since DFT–PBE predicts a nonzero magnetization ground state for face-centered cubic (fcc) Pd, (39) we suppressed this unphysical excess electronic spin by using spin-restricted (S = SZ = 0) DFT. We utilized a self-consistent dipole-field energy and potential correction orthogonal to the surface to remove the spurious dipole-field interactions between periodic images along that direction. (40) We sampled the Brillouin zone with the Monkhorst–Pack method (41) utilizing Γ-point-centered 16 × 16 × 16, 7 × 7 × 1, and 5 × 5 × 1 k-point meshes for the bulk four-atom unit cell, and the five-layer 3 × 3 and four-layer 5 × 5 supercell periodic slabs, respectively (vide infra). We employed the Methfessel–Paxton method (42) with a width of 0.09 eV to facilitate electronic convergence by smearing the electronic states at and near the Fermi level. We chose these parameters to achieve a total energy convergence of 1.0 meV/atom or better.

2.2. Atomic Models

We optimized the bulk nonmagnetic four-atom fcc Pd unit cell (a = 3.940 Å, −1.3% error (43)) and then generated five-layer 3 × 3 and four-layer 5 × 5 periodic slabs of the most stable (111) surface using the optimized unit cell lattice vectors. A vacuum layer of ∼15 Å was added along the surface normal to separate the slab from its periodic images. The dimensions of the five-layer 3 × 3 and four-layer 5 × 5 supercells are 8.357 Å × 8.357 Å × 24 Å and 13.929 Å × 13.929 Å × 22 Å, respectively.

2.3. Geometry and Minimum Energy Path Optimization

For geometry optimization, we set the maximum absolute total force on each atom to 0.01 eV/Å as the convergence threshold. For five-layer 3 × 3 (Pd45) and four-layer 5 × 5 (Pd100) supercells, we fixed the lattice vectors along with the coordinates of the two bottom-most layers for the former and the very-bottom layer for the latter to emulate a semi-infinite bulk surface while relaxing the atoms in all other layers. We optimized the minimum energy paths (MEPs) for reactions via the climbing image nudged elastic band (CI-NEB) method, (44,45) setting the fictitious spring force constant to 5.0 eV/Å2 and the force convergence threshold to 0.03 eV/Å. We utilized the five-layer 3 × 3 supercells for the CI-NEB calculations involving the dehydrogenation of NH3, NH2, and NH. This supercell size is also adequate for the calculation of geometry-specific embedding potentials for studying the first N–H bond scission step (vide infra). On the other hand, we used the four-layer 5 × 5 supercells for the CI-NEB calculations involving the associative desorption of 2*N to N2 and all other embedding potential optimizations. We define the reaction coordinate (RC) as the cumulative distance between the images on the MEP
RC(n)=i=0n1|Ri+1Ri|
(1)
where RC(n) is the RC value for the nth image and Ri represents the Cartesian coordinate vectors of the atoms in image i.

2.4. Density Functional Embedding Theory with ECW Theory

In density functional embedding theory (DFET), (23,24,46) one divides the total system into a region (or regions) of interest (here, a cluster) and its (or their) environments (if one has multiple embedded subsystems) and derives an embedding potential Vemb to describe the interaction between them. We used the four-layer 5 × 5 supercell described above as the total system, carving out Pd10, Pd12, and Pd14 clusters (displayed in Figure 1; rationales for these choices are provided later) from the aforementioned Pd100 slab, and derived the Vembs using our in-house modified VASP 5.3.3 code, which is now also available for VASP 6. (47) In DFET, one optimizes Vemb by maximizing the extended Wu–Yang functional, (48) W[Vemb]
W[Vemb]=Ecl[ρcl,Vemb]+Eenv[ρenv,Vemb]ρtotVembdr
(2)
where, for the simplest case of two subsystems, Ecl and Eenv are the DFT energies of the cluster and its environment in the presence of the Vemb, respectively, and ρcl, ρenv, and ρtot are the self-consistently optimized electron densities of the cluster, environment, and total system (the latter being simply the periodic DFT reference density), respectively. We determine the optimal Vemb by setting to zero the functional derivative of W with respect to Vemb. Because the functional derivative of the energy with respect to Vemb is formally equal to the density, Vemb convergence is reached when ρtot = ρcl + ρenv, which also makes sense physically: the sum of the embedded subsystem densities, given the physically correct embedding potential, should reproduce the total density at the same level of theory.

Figure 1

Figure 1. (a) Four-layer 5 × 5 Pd(111) (Pd100) periodic slab model. Top-down view of the (b) Pd10 cluster, (c) Pd12 cluster, and (d) Pd14 cluster carved out from the Pd100 periodic slab. Subsurface atoms are faded out. Isosurface plots (yellow: +1.2 V, cyan: −1.2 V) of the optimized embedding potentials generated for the (e) Pd10 cluster in its Pd90 environment, (f) Pd12 cluster in its Pd88 environment, and (g) Pd14 cluster in its Pd86 environment.

To prepare for embedded cluster calculations, we transform Vemb from a real-space grid basis to an atomic Gaussian-type-orbital (GTO) basis using our in-house code. (49) Here, we performed embedded cluster calculations using MOLPRO 2021.2, (50−52) utilizing the matrix operation (MATROP) function as implemented in MOLPRO to add Vemb in the GTO basis to the original one-electron Hamiltonian H0. The following embedded DFT and ECW calculations, therefore, include the interactions between the cluster with its periodic environment at the PW-DFT level, here, within the PBE XC functional approximation.
We chose CASPT2 for carrying out ECW cluster calculations, a method that relies on CASSCF to generate the reference many-body wavefunction. We utilized a 12-electron in 12-orbital (12e, 12o) AS for the emb-CASSCF calculations of N–H bond dissociation in NH3, *NH2, and *NH. The AS comprises three N–X (X = H or Pd) σ-bond orbitals and their corresponding antibonding σ* orbitals and the three highest occupied molecular orbitals (HOMO, HOMO – 1, HOMO – 2) and the three lowest unoccupied orbitals (LUMO, LUMO + 1, LUMO + 2) originating from the Pd clusters. Ground-state emb-CASPT2 calculations were based on the ground-state emb-CASSCF wavefunctions. To obtain excited-state wavefunctions, we applied the emb-SA-CASSCF method, utilizing the ground-state emb-CASSCF wavefunctions as the initial guess. We assigned equal weights to all roots for a balanced treatment of each state and we computed single-state emb-CASPT2 energies for each state based on the emb-SA-CASSCF wavefunctions. Finally, we added the emb-CASPT2 excitation energies onto the emb-CASPT2 ground-state PECs for each image along MEP to get the excited-state PECs. (21,22) For emb-CASPT2 calculations labeled “without 4s4p dynamic correlation”, we included all excitations from all orbitals, except for the core orbitals, which comprise the 28 core electrons of each Pd subsumed into its effective core potential (ECP) and the semicore 4s4p orbitals of Pd and the core 1s orbital of N. For emb-CASPT2 labeled “with 4s4p dynamic correlation,” excitations from these outer core orbitals were included. We applied an IPEA shift (53) of 0.25 hartree to alleviate the intruder state problem in CASPT2 in combination with a level shift (54) of 0.3 hartree to facilitate CASPT2 energy convergence, which are especially critical for metal clusters with a high degree of degeneracies.
Finally, we express the total ECW energy (EtotalECW) as the energy of the total system in PW-DFT-PBE (EtotalPWDFTPBE) plus a correction obtained from the difference between the embedded cluster energy computed using GTO-DFT-PBE (Eemb,clusterGTO-DFT-PBE) and a CW method (Eemb,clusterGTOCW) in the presence of Vemb: (23,24,34,46)
EtotalECW=EtotalPWDFTPBE+(Eemb,clusterGTOCWEemb,clusterGTODFTPBE)
(3)
To demonstrate convergence with respect to cluster size and explore the effects of using a fixed vs structure-specific Vemb, we replaced the CW method with a much simpler electronic structure theory, namely, DFT with local density approximation (LDA) for the XC functional. We simply require the “correcting” method to be distinct from the base method that one used to construct Vemb. Therefore, in Section 3.1, among the various tests we conducted, we calculated the emb-PW-DFT-LDA and emb-GTO-DFT-LDA energies using similar formulas
EtotalembPWDFTLDA=EtotalPWDFTPBE+(Eemb,clusterPWDFTLDAEemb,clusterPWDFTPBE)
(4)
EtotalembGTODFTLDA=EtotalPWDFTPBE+(Eemb,clusterGTODFTLDAEemb,clusterGTODFTPBE)
(5)

3. Results and Discussion

ARTICLE SECTIONS
Jump To

3.1. Effects of Cluster Size (Pd10, Pd12, and Pd14)

First, we established the effects arising from the size of the embedded metal cluster within embedded PW-DFT, i.e., embedded-DFT-in-DFT, which we achieved by performing emb-PW-DFT-LDA, henceforth further abbreviated as emb-LDA, for the dissociative adsorption reaction NH3(gas) → *NH3 → *NH2 + *H. We used Pd10, Pd12, and Pd14 clusters (the structures of these clusters are shown in Figure 1b–d) to investigate cluster size convergence. The structures of all images along the MEP are displayed in the Supporting Information (SI), Figure S1. Figure 2 shows that the emb-LDA energies of images 0–2 (gas-phase structures), 4 (adsorbed structure), and 7–10 (the NH2–H bond breaking) overlap very well with the full PW-DFT-LDA energies regardless of the cluster size, with differences in the cluster size leading to deviations in relative energies of less than 0.05 eV (compare the red line with the blue lines in Figure 2). Figure 2 also shows that PW-DFT-PBE (black line) and PW-DFT-LDA (red line) predict very distinct reaction curves, with the latter exhibiting the typical LDA overbinding. Note that we used the former electronic structure theory to derive the Vemb.

Figure 2

Figure 2. Effect of cluster size for NH3(gas) → *NH3 → *NH2 + *H. We calculate the emb-LDA energies (defined in eq 4) for Pd10, Pd12, and Pd14 clusters. Numerical indices correspond to structures shown in Figure S1.

The minor discrepancies between emb-LDA energies and PW-DFT-LDA energies for images 3, 5, and 6 (respectively, NH3 close to the surface and the tilted adsorbed states) only slightly decrease as the cluster size increases. As it turns out, this energetic discrepancy between emb-LDA and PW-DFT-LDA originates from the fixed-surface approximation, i.e., the cluster atoms are fixed to their optimized positions in the absence of the adsorbate, introduced when constructing Vemb (see also Section 3.4). As proof, we find that emb-LDA PECs almost perfectly reproduce the PW-DFT-LDA PECs when the full reference system PW-DFT-LDA calculations also have all Pd atoms fixed to their clean-surface structure (PW-DFT-LDA-clean, Figure S2a). Figure S2b shows that PW-DFT-PBE is less sensitive to the fixed surface approximation between images 2 and 7. Unlike the total emb-LDA energies (Etotalemb-PW-DFT-LDA), however, the cluster-only emb-LDA energy terms (Eemb,clusterPWDFTLDA) change significantly with respect to the cluster size (Figure S3). This is because embedded cluster energies include the embedding energy contribution (∫Vembρcluster(r) dr), which describes the interaction between the cluster and adsorbate electron densities and Vemb. Changes in the size or shape of the cluster can cause significant changes in said interaction. However, by casting the embedding theory total energy (EtotalembPWDFTLDA) as the total DFT energy plus a regional correction involving the difference between two terms containing the embedding potential, as done in our embedding (DFET/ECW) formalism, the cluster size/change sensitivity quickly damps out.
We also compared the emb-PW-DFT-LDA and emb-GTO-DFT-LDA energies (eqs 4 and 5 and Figure S4) and found that the relative energy discrepancies between the two basis set types are minor and originate from the differences in their size (completeness) and their core electron representation and from BSSEs when using atom-centered GTO basis sets (vide infra).
Our results demonstrate that (1) the fixed surface approximation in deriving Vemb and energy correction leads to negligible and tractable energetic discrepancies although such an approximation requires testing (see also Section 3.4); (2) the minimal Pd10 cluster is sufficiently large to obtain an accurate Vemb and moving to larger Pd12 and Pd14 clusters negligibly affects the energetics for the NH3(gas) → *NH3 → *NH2 + *H reaction, based on the embedding total energy expression; and that (3) transforming Vemb from a PW basis to a GTO basis retains accuracy, with minor discrepancies in part arising from BSSE (vide infra) and not from the ECW theory.

3.2. GTO Basis Set Size Effect on Ground-State PECs within Emb-CASPT2

Following the initial studies on cluster size dependence, we utilized the embedded Pd10 cluster to study the effects of the basis set size on the ground- and excited-state PECs for the NH3 (gas) → *NH3→ *NH2+ *H reaction. The semicore Pd 4s4p dynamic correlation was excluded from the CASPT2 calculations (see Methods). We calculated a DFT-corrected initial adsorption energy (Einit, expt) for NH3 on Pd(111) to serve as a benchmark for the energy of the molecularly adsorbed state of NH3 (image 4, RC ∼ 4.8 Å), using the equation Einit,expt = Einit,PW-DFT + ΔHinit,expt – ΔHinit,PW-DFT. Einit,PW-DFT and ΔHinit,PW-DFT are the PW-DFT-PBE reaction internal energy (−0.76 eV) and enthalpy (−0.67 eV, see ref (55) for details on how this term is calculated) changes at 305 K, respectively, at a coverage of θNH3 = 1/9. ΔHinit,expt (−0.77 eV) is the initial heat of adsorption measured at 305 K obtained from calorimetry and Pd powder. (56)
We tested the following Dunning basis sets: (57−60) cc-pVXZ-PP (hereafter, denoted VXZ), aug-cc-pVXZ-PP (AVXZ), cc-pwCVXZ-PP (WCVXZ), and aug-cc-pwCVXZ-PP (AWCVXZ) with X = D, T to determine the effects of adding higher-order polarization functions (pVX), diffuse functions (aug), and core–valence (CV) correlation. We retrieved all basis sets from the basis set exchange (61) and tested different basis set combinations for Pd and adsorbate (N and H), which are denoted as Pd/adsorbate, e.g., VDZ/VDZ if cc-pVDZ is utilized for Pd, N, and H. The energy curves for the surveyed basis set combinations are displayed in Figure 3. We found that all basis sets yield similar TS barriers, ranging from 1.57 to 1.74 eV, except for the two smallest basis set combinations, VDZ/VDZ and VDZ/AVDZ, which yield over- and under-estimations, with predicted barriers of 1.81 and 1.37 eV, respectively. The larger basis sets give molecular adsorption energies that are in good agreement with the DFT-corrected experimental benchmark Einit,expt (shown as the dashed black line in Figure 3), with AVTZ/AVTZ showing the largest deviation at −0.14 eV. Our data demonstrate that all basis sets for Pd larger than VDZ yield qualitatively similar results, agreeing to within 0.20 eV for the adsorption energy and to within 0.17 eV for the TS barrier.

Figure 3

Figure 3. Basis set effect for NH3(gas) → *NH3 → *NH2 + *H at the emb-CASPT2 level using an embedded Pd10 cluster. Basis sets for Pd and adsorbates are listed as Pd basis/adsorbate basis in the legend. DFT-corrected Einit,expt is the DFT-corrected electronic energy based on the experimental enthalpy of NH3 adsorption on Pd powder. See detailed explanation in the text.

We also investigated the effect of including semicore Pd 4s4p dynamic correlation on the emb-CASPT2 energies (Figure S5; AVTZ/AVTZ results are not available due to prohibitively high memory requirements). Including the semicore Pd 4s4p dynamic correlation that reduced the TS barrier for all tested basis sets and for smaller basis sets without CV correlation, the difference in the TS barrier could be as large as ∼0.3 eV (Figure S5a,b,d). Adding CV correlation, specifically, VDZ to WCVDZ (compare Figure S5a,b vs c) and AVDZ to AWCVDZ (Figure S5d vs e) and higher-order angular momentum functions (VDZ to VTZ) (Figure S5b vs f,g) reduces the effect of including 4s4p dynamic correlation on the barrier to ∼0.1 eV. Consequently, a basis set suitable to capture the CV correlation (as in the WCV basis set family) is needed to properly include the semicore 4s4p dynamic correlation─otherwise it is preferable to not include it in the calculation. This is likely the result of an increase in the 4d95s1 → 4d85s2 excitation energy upon inclusion of 4s4p dynamic correlation, which was found by Peterson et al. (57) to be 3.6 kcal/mol in the complete basis set limit for the Pd atom. Although the inclusion of the semicore 4s4p dynamic correlation has a relatively small effect (∼0.1 eV) on the ground-state and excitation energies of Pd, CV correlation is important for alkali (group 1) and alkaline earth (group 2) metals (62) and first-row transition metals (63) and should be considered in these cases with the proper basis set to capture them. Additionally, Peterson et al. found that the inherent error within the ECP approximation for the Pd atom is small (ΔEECP < 1.1 kcal/mol) at the CCSD(T) level of theory when ignoring the semicore 4s4p dynamic correlation. (57) Here, ΔEECP was defined as the difference between the results using all-electron Douglas–Kroll–Hess (DKH) basis sets of triplet-ζ quality with the DKH Hamiltonian (64,65) and the results using aug-cc-pwCVTZ-PP quality basis sets with a nonrelativistic Hamiltonian. As the error within the ECP approximation for Pd is minimal, we did not consider it here.
We conclude that AVDZ/AVDZ basis sets without evaluating the semicore 4s4p dynamic correlation offer the best balance between computational cost and accuracy, given the considerable expense of adding CV correlation and the attendant 4s4p dynamic correlation. Diffuse functions, which are known to be critical for treating excited states, especially those with Rydberg character, (66,67) are included, providing significant improvements over VDZ/VDZ results, while the larger AVTZ/AVTZ basis set is too computationally intensive and requires prohibitively large amounts of memory. We utilized the AVDZ/AVDZ basis sets for the remainder of the calculations discussed in the following sections.

3.3. BSSE and Counterpoise Correction

BSSEs often arise in weakly bound systems when treated with basis sets of small size as a consequence of borrowing basis functions from adjacent atoms at short interatomic distances. When these distances change, they lead to an apparent strengthening of short-range intermolecular interactions because in the separated case, the atoms can no longer borrow each other’s basis functions. (68−71) The BSSE may be removed by utilizing the so-called counterpoise (CP) correction, performing calculations on the individual fragments utilizing ghost functions (the other fragment’s basis set) to obtain corrections to the interaction energy. (72−74) The calculations of the BSSE and CP corrections are discussed in more detail in the Supporting Information (Note S1).
Our results, shown in Figure 4, demonstrate that BSSE increases as NH3 adsorbs and dissociates on the embedded Pd cluster (higher values of the RC), arising from the adsorbate’s ability to borrow the basis functions of Pd (and vice versa) as the molecule (or its fragments) approaches the surface. The results also confirm that it is essential to utilize a basis set with diffuse functions to minimize BSSE, with the smallest surveyed VDZ/VDZ (red) resulting in BSSEs of up to 1.0 eV in CASPT2 calculations for the NH3(gas) → *NH3 → *NH2 + *H reaction. CASPT2 calculations with AVDZ (yellow), AVTZ (purple), and AWCVDZ/AVDZ (green) basis sets suffer much less BSSE (less than 0.5 eV).

Figure 4

Figure 4. CP corrections for NH3(gas) → *NH3 → *NH2 + *H on embedded Pd10. Combined CP corrections of the adsorbates and the Pd10 cluster are plotted for emb-CASPT2 with various basis sets (denoted as Pd basis/NH3 basis in the legends) and DFT-PBE with the AVDZ basis set.

It is well known that self-consistent mean-field theories, such as DFT, achieve basis set convergence (and hence lower BSSE) faster than non-self-consistent CW theories such as CASPT2. This is consistent with our findings, with CASPT2 AVTZ/AVTZ errors still exceeding those of DFT-PBE at the AVDZ/AVDZ level. Comparing results for different basis sets on the adsorbate, namely, VDZ/AVDZ (pink) to VDZ/VDZ (red) and VTZ/AVTZ (blue) to VTZ/VTZ (light blue), reveals that the addition of diffuse functions on only the adsorbate can worsen BSSE by ∼0.4 eV. These data illustrate the importance of using basis sets of equivalent quality on both adsorbate and the metal cluster. We also posit that AVDZ/AVDZ presents the optimal choice (yellow) of basis sets for emb-CASPT2 simulations, given that simulations utilizing AVTZ basis sets, especially at the SA-CASSCF and CASPT2 levels of theory, are too computationally expensive for routine use.
By comparing the implied needed CP corrections from the estimated BSSEs in Figure 4 and the PECs in Figure 3, both as functions of the basis set, it becomes clear that the CP corrections will be severe overestimations of the errors due to the basis set size. This behavior has been known in the literature, notably for small basis sets. (74−76) In Figure 3, all PECs agree to less than 0.5 eV except for the case of the VDZ/AVDZ basis set, whereas Figure 4 suggests CP corrections more than +0.75 eV relative to AVTZ/AVTZ. This implied correction is much higher in magnitude and in the opposite direction as Figure 3 suggests. Figure 3 shows that AVTZ/AVTZ is among the basis set with the most negative adsorption and dissociation energies and lowest effective barrier relative to NH3(g) with VDZ/AVDZ as an outlier. We therefore additionally recommend that basis set extrapolation to the complete basis set limit (77) should be performed in lieu of CP correction if calculation with multiple basis sets can be afforded, as we have done, e.g., in ref (78).

3.4. Effects of Geometry-Dependent Vemb

Next, we investigated the effect of utilizing geometry-dependent Vembs on emb-CASSCF PECs. Given that many reactions on metal surfaces do not result in significant structural changes of the metal surface, utilizing a fixed geometry for the metal surface is often a reasonable approximation (Section 3.1) where one assumes that the periodic DFT term in the embedding total energy already correctly captures the surface reorganization energy. However, the frozen surface approximation should be tested in cases where the surface metal atoms change significantly. (21) We carried out one such test here, where in the case of adsorbed nitrogen atoms recombining to form N2 gas on Pd(111), the optimized structures along the minimum energy path show fairly large structural changes on the Pd surface.
Up to this point, all results presented above utilized geometry-independent Vembs obtained from a clean Pd surface for all calculations. We now compare the reaction barrier for *N + *N → *[η2-N2] → N2(g) using a geometry-independent Vemb and geometry-dependent Vembs (Figure S6), where *[η2-N2] denotes a molecularly adsorbed N2 molecule in a bidentate (η2) configuration (reaction path discussed in ref (55)). For the geometry-independent Vemb, we obtained Vemb as usual from a clean Pd100 slab (Figure S6a). The metal–adsorbate complex structures then were obtained by translating the optimized reactant and TS structures of the adsorbed N atoms from CI-NEB calculations for the periodic slab onto the clean Pd10 cluster while fixing the internal coordinates of the adsorbates at their PW-DFT-PBE structures. Following this procedure, we projected the Vemb in the PW basis derived from the clean Pd10 cluster to various adsorbate configurations in the GTO basis. The geometry-dependent Vemb instead was derived using the optimized structure of the Pd100 + 2 *N (or *[η2-N2]) periodic slab instead of the clean Pd100 periodic slab, i.e., with the adsorbate present. Figures S6d,e and g,h illustrate the total system and cluster used to derive geometry-dependent Vemb for the reactant and TS, respectively.
We also performed “modified-geometry-dependent” Vemb calculations to determine Vemb from a clean slab without adsorbates and with the clean Pd100 at the geometry optimized with 2 *N or *[η2-N2], i.e., we optimized Vemb without the adsorbates (as we do in the fixed cluster approximation) but the surface Pd atoms are configured as if the adsorbates are present. Emb-LDA reaction barriers derived from geometry-independent, geometry-dependent, and modified-geometry-dependent Vembs are displayed in Table S1. The modified-geometry-dependent Vemb most closely reproduced the full periodic slab DFT-LDA prediction, with an error of only −0.15 eV. The geometry-independent Vemb was slightly less reliable, with an error of 0.36 eV. Note that the errors of geometry-independent Vemb are much smaller for the three dehydrogenation steps than for N2 associative desorption because the adsorbate-induced metal surface distortions are larger in the latter. Interestingly, the geometry-dependent Vemb exhibited the largest error of 0.85 eV, likely the result of significant changes in density upon adsorption, which are larger than those arising from changes in geometry. These findings suggest that it could be worthwhile to use the modified-geometry-dependent Vemb, i.e., still without the adsorbate, when the structure of the metal surface differs significantly from its clean-surface structure; however, optimization of Vemb corresponding to each adsorbate structure along the MEP is computationally demanding and improvements to the accuracy are minor. These results suggest that the use of the geometry-independent Vemb is in most cases sufficient, given the current cost to optimize Vemb. Ongoing work in our group endeavors to reduce this cost so that in the future, the modified-geometry-dependent Vemb might be worth the cost to generate along reaction paths.

3.5. Comparison of Creeping vs Merging for AS Selection in the Reference Emb-CASSCF

Next, we investigated the effects of AS selection on the PECs for the NH3 dehydrogenation steps. Here, we compare the PECs obtained from two commonly utilized approaches for generating initial AS orbital guesses when modeling catalytic pathways, namely, the creeping and merging approaches. The former refers to the strategy of constructing ASs using orbitals of adjacent or successive structures along the reaction path as the initial guess for the CASSCF calculation for the current structure, while the latter involves the generation of new guess orbitals for each structure from the separately optimized orbitals of each fragment in the system. The PECs are displayed in Figure 5 and the NOs comprising the ASs are shown in the Supporting Information (obtained with creeping in Figures S7–S9 and from merging in Figures S10–S12.)

Figure 5

Figure 5. Emb-CASPT2 results using creeping to obtain CASSCF orbitals for (a) NH3(gas) → *NH3 → *NH2 + *H, (b) *NH2 + *H → *NH+ 2*H, and (c) *NH → *N+ *H. The corresponding NOs in the ASs using creeping are given in Figures S7–S9, while the ones from merging are shown in Figures S10–S12 in the Supporting Information. All calculations utilize the Pd10 cluster model.

There are two key indicators of a well-chosen AS: (1) the PEC from creeping is continuous during bond breaking/forming and (2) the orbitals in the AS correspond to the correct bonding pattern for all geometries. Conversely, indicators of a bad AS include: (1) a discontinuous PEC even without bond breaking/forming, (2) flipping of the energy ordering of ground and excited states going from SA-CASSCF to SS-CASPT2, and (3) nonoptimal orbital rotation from inactive to ASs or AS orbital occupations along the path, likely due to new degeneracies associated with different d-orbitals near the Fermi level.
Comparing the results of creeping and merging approaches for the NH3 dehydrogenation steps, we observe that discrepancies usually occur around bond breaking and forming. This is illustrated in Figure 5a (RC ∼ 7 Å), where *NH2 forms a new bond with another Pd atom in the top layer; Figure 5b (RC ∼ 0.5–1 Å), where a N–H bond breaks; Figure 5b (RC ∼ 1.5–2.8 Å), where a new N–Pd bond forms; and Figure 5c (RC ∼ 1.5 Å), where a N–H bond breaks; the largest discrepancies are generally observed around the product and the TS. While the failure of the creeping approach is less obvious for the first and third dehydrogenation steps (Figure 5a,c), where energetic deviations between PW-DFT-PBE and emb-CASPT2 obtained from merging are small, here, we encounter violation of rule (3), i.e., we observe the nonoptimal occupation of metal-based AS orbitals along the path. Comparisons of the AS NOs from creeping and merging approaches show that the (6e, 6o) orbitals primarily associated with the adsorbate are consistent across both, while the (6e, 6o) orbitals primarily from the metal cluster differ (e.g., compare Figures S7 vs S10). In creeping, the metal-based orbitals spatially localize on the same set of Pd atoms along MEPs (Figure S7), which is expected given that each calculation utilizes the orbital coefficients of the previous structure as the initial guess. On the other hand, in the merging approach, metal orbitals may localize on different Pd atoms along MEPs (Figure S10) as the bonding patterns and, consequently, the NOs change when the adsorbate geometry changes along the MEP. Creeping fails to incorporate these changes in the metal orbitals into the AS, falling into local minima during the CASSCF optimization.
Earlier studies by Bao and Carter on NH3 decomposition on Cu(111) (19) and Ru-doped Cu(111) (20) using the creeping strategy delivered consistent results. Comparing the AS NOs of Pd(111) with Cu(111) and Ru-doped Cu(111), we note that the orbitals on Pd are mostly of d-orbital character, while the orbitals on Cu are mostly of s-orbital character. d orbitals are localized and anisotropic, while s orbitals are delocalized and isotropic. This observation may explain why creeping presented a viable approach for pure and doped copper (Cu) (18−21,79) and pure and doped gold (Au) (13−17) in previous studies, while it failed for Pd. In the former examples, either the delocalized sp bands or the well-defined d orbitals of the surface dopants were involved in bonding with the adsorbate, while in the latter case, the nearly degenerate d bands dominated the bonding with the adsorbate, resulting in a plethora of local minima and leading to errors in the PEC when creeping is employed. Merging solves this problem.

3.6. Number of Excited States Included in Emb-SA-CASSCF Calculations

Finally, we investigated the sensitivity of excitation energies and excited-state PECs with respect to the number of states included in SA-CASSCF calculations. Such calculations play a key role in the modeling of plasmonically driven and photocatalytic processes, (25,34) and emb-NEVPT2 and emb-CASPT2 calculations have been utilized successfully to elucidate, e.g., the mechanism of plasmon-induced methane dry reforming to syngas on Ru-doped Cu nanoparticles (21) and the photocatalytic generation of H2 from NH3 utilizing Fe-doped Cu nanoparticles. (18) Here, we consider excited-state photocatalytic decomposition of NH3 via the reactions NH3 (gas) → *NH3 → * NH2 + *H (Figure 6), *NH2 + *H → *NH + 2*H, and *NH → *N + *H (Figures S13 and S14, respectively). We draw some general conclusions from this series of PECs next.

Figure 6

Figure 6. Ground- and excited-state MEP energetics as calculated with emb-CASPT2 for NH3 (gas) → *NH3 → * NH2 + *H. Red arrows represent vertical excitation of S0 → SN-1, where SN-1 represents the highest excited state. Green arrows represent reaction barriers for S0 or SN–1. Each panel represents different numbers of states included in the N-state emb-SA-CASSCF calculations: N = (a) 5, (b) 7, (c) 9, (d) 11, (e) 15, and (f) 17.

First, in the case of distinguishable bands (characterized by a large energy gap between two consecutive states), the data highlight that the entire band must be included in the emb-SA-CASSCF calculations to avoid root flipping and abnormal excitation energies. For example, three distinguishable bands exist at the reactant configuration (RC ∼ 4.8 Å, Figure 6f) in the first N–H bond breaking when looking at the emb-CASPT2 PECs from the 17-state (the largest number of states we were able to include) emb-SA-CASSCF calculation. These bands of states are composed of the singlet ground state to the fourth excited state (S0–S4), fifth to 14th singlet excited states (S5–S14), and 15th to 16th singlet excited states (S15–S16). When the second band is only partially included in the emb-SA-CASSCF calculations, as in the seven-, nine-, and 11-state calculations (Figure 6b–d), the excited states S2–S4 are significantly different from the results when either only the first band is included, as in the five-state calculation (Figure 6a), or the entire first and second bands are included, as in the 15-state calculation (Figure 6e). Increasing the number of excited states by groups of bands generally preserves the overall shape of the lower band, although the excitation energies of the lower band will change slightly when higher bands are included. When including the third band, as in the 17-state calculation (Figure 6f), the lower bands of the three images with RC < 2.5 Å shift to much higher energies, in comparison to the results with only the first and second bands (e.g., 15-state, Figure 6e). Therefore, including these high-lying excited states, namely, S15 and S16, deteriorates the quality of lower excited states. With respect to NH3 adsorption (images 0–4), we conclude that inclusion of the entire first and second bands (15-state, Figure 6e) yields the appropriate description, while the effect of the excited states on the NH2–H bond breaking barrier is best described when the additional third band of excited states is also included (17-state, Figure 6f).
Comparing the above PECs for the first N–H bond scission to the subsequent H abstraction steps, it appears that the breaking of the third and final N–H bond (Figure S14) is less sensitive to the choice of the number of states included in the SA-CASSCF calculations than that of the first (Figure 6) and second (Figure S13) N–H bonds. Inspection of the dipole moments of all studied states reveals that they are small and of similar magnitude for *NH → *N + *H, while several states have significantly different or very large dipole moments for the NH3 (gas) → *NH3 → * NH2 + *H and *NH2 + *H → *NH + 2*H reactions (an Excel document containing tables of dipole moments for all states of each structure along each of the bond dissociation reaction paths may be found in the Supporting Information. Each sheet contains tables for the dipole moments of all distinct N-state SA-CASSCF calculations carried out). The dipole moments of the different states therefore can be used as a secondary method to group the excited states into different classes, serving as a convenient alternative indicator to determine if additional states added in the SA procedure will belong to the same already sampled band of states (without the need to calculate even more states to identify bands). As stated above, it is crucial to sample adequately a band to avoid root flipping. We found that both NH3 (gas) → *NH3 → * NH2 + *H and *NH2 + *H → *NH + 2*H (Supporting Information Excel document, tabs labeled as NH3 image# and NH2 image#, respectively) exhibit varying polarization profiles among different excited states, thus, explaining the sensitivity of the energy landscape with the number of states in SA-CASSCF. For certain reactions, one may encounter less distinct groups of excited states as indicated by their dipole moments, e.g., *NH → *N + *H (Supporting Information Excel document, sheets labeled as NH image#), and wherever that happens, the result will be less sensitive to the number of states being averaged in SA-CASSCF.
These results highlight the challenge presented to emb-SA-CASSCF simulations, which rely on a single set of compromise canonical orbitals used to expand the wavefunctions of all states being averaged in the SA-CASSCF procedure, when treating states with different characters. In view of the foregoing, we conclude that two factors prevent the study of higher-lying excited states with emb-SA-CASSCF-based methods. First, adding higher excited states deteriorates the quality of the SA-CASSCF wavefunction, especially when the characters of the excited states are very different (e.g., *NH2 + *H → *NH + 2*H) or a large energy gap exists between the higher and the lower excited states (e.g., NH3 (gas) → *NH3 → * NH2 + *H). This problem may be alleviated using variable weighted SA-CASSCF; however, this solution introduces more variables to optimize and potential additional sources of error. Second, higher-lying excited states become too dense to yield any additional insights into the effect of excitation on the effective barrier and will only make the calculation more expensive (as is seen in the *NH2 + *H → *NH + 2*H and *NH → *N + *H reactions). These issues potentially could be solved by using multistate CASPT2; (80) however, this method is significantly more costly and proved to be intractable for even just two states of the systems we considered here.

4. Summary and Conclusions

ARTICLE SECTIONS
Jump To

In the present study, we used a Pd(111) + NH3 model system to systematically determine the optimal approximations for embedded (SA-)CASSCF/CASPT2 simulations conducted within DFET. Specifically, we considered how PECs are affected by cluster size, basis set, structure of the cluster when Vemb is determined, method utilized to generate initial MO guesses for the CASSCF AS, and the number of states included in the SA-CASSCF calculations of excited states. We conclude that the energetics converge quickly even with a small cluster of 10 atoms embedded in a fairly large periodic slab, as long as all direct metal–adsorbate interactions are included. The relative insensitivity to the cluster size is due to the DFET approach that damps out edge/finite-size effects rapidly and the DFET energy expression which limits the size-dependence due to the regional nature of the energy correction. Second, we found that the aug-cc-pVDZ basis set provides the best compromise between accuracy and computational cost, with the larger aug-cc-pVTZ basis set yielding only minor improvements while significantly reducing computational tractability. That said, we recommend complete basis set extrapolation rather than CP corrections, which overestimate the finite basis set errors. Third, we found that it is best to determine Vemb without the adsorbates but using a cluster structure reflecting the structure of the surface as if adsorbates are present, although a fixed-clean-surface approximation holds well even when the surface undergoes significant structural changes. Fourth, by comparing merging and creeping approaches to generate initial AS guesses for the three consecutive dehydrogenation steps of NH3 on Pd(111), we found significant shortcomings in the ability of a creeping approach to generate consistent ASs along reaction pathways when simulating surfaces consisting of highly localized and near-degenerate, partly filled d-orbital manifolds, such as in the case of Pd (which adopts an s1d9 valence electronic structure in its bulk phase), leading to discontinuities in the CASSCF/CASPT2 PECs. In such cases, we recommend utilizing a merging approach, generating ASs from the molecular fragments at each surveyed configuration. Finally, when aiming to study excited-state reaction pathways, e.g., in the study of photocatalytic or plasmonic processes, we demonstrated that it is essential to include bands of excited states in their entirety to avoid discontinuities in the PECs. This generally requires performing benchmarks that identify how excited states group into bands in SA-CASSCF calculations.
The results presented in this study provide important insights for future studies aiming to elucidate mechanisms of heterogeneously catalyzed reactions utilizing (SA-)CASSCF-based ECW theory simulations. Importantly, the guidelines derived here will significantly reduce the need to perform initial benchmarking calculations when commencing an ECW theory study of a new chemical system, which are generally necessary to choose the correct level of theory and often consume significant computational resources.

Supporting Information

ARTICLE SECTIONS
Jump To

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.jctc.4c00558.

  • Plots of all geometries, clean-surface approximation energetics, cluster-size effects, PW to GTO basis set effects, Pd 4s4p dynamic correlation effects, overview of BSSE and CP, plots and energetics of geometry effects on Vemb, NO plots of all structures surveyed, and SA-CASPT2 ground- and excited-state MEPs utilizing different state averaging (PDF)

  • Tables of SA-CASSCF dipole moments (XLSX)

Terms & Conditions

Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.

Author Information

ARTICLE SECTIONS
Jump To

  • Corresponding Author
    • Emily A. Carter - Department of Mechanical and Aerospace Engineering, Princeton University, Princeton, New Jersey 08544-5263, United StatesPrinceton Plasma Physics Laboratory, Princeton, New Jersey 08540-6655, United StatesAndlinger Center for Energy and the Environment, Princeton University, Princeton, New Jersey 08544-5263, United StatesOrcidhttps://orcid.org/0000-0001-7330-7554 Email: [email protected]
  • Authors
    • Xuelan Wen - Department of Mechanical and Aerospace Engineering, Princeton University, Princeton, New Jersey 08544-5263, United StatesOrcidhttps://orcid.org/0000-0002-7151-6157
    • Jan-Niklas Boyn - Department of Mechanical and Aerospace Engineering, Princeton University, Princeton, New Jersey 08544-5263, United StatesOrcidhttps://orcid.org/0000-0002-6240-3759
    • John Mark P. Martirez - Princeton Plasma Physics Laboratory, Princeton, New Jersey 08540-6655, United StatesOrcidhttps://orcid.org/0000-0003-0566-6605
    • Qing Zhao - Department of Mechanical and Aerospace Engineering, Princeton University, Princeton, New Jersey 08544-5263, United StatesPresent Address: Department of Chemical Engineering, Northeastern University, Boston, Massachusetts 02115–5005, United StatesOrcidhttps://orcid.org/0000-0002-5003-9355
  • Author Contributions

    The manuscript was written through contributions of all authors. All authors have given approval to the final version of the manuscript.

  • Notes
    The authors declare no competing financial interest.

Acknowledgments

ARTICLE SECTIONS
Jump To

The Air Force Office of Scientific Research via the Department of Defense Multidisciplinary University Research Initiative, under award no. FA9550-15-1-0022, provided the financial support for the majority of this work. E.A.C. and J.M.P.M. were also supported by the U.S. Department of Energy, Office of Science, Basic Energy Sciences (BES), under award no. DE-SC0023357. Princeton University’s Terascale Infrastructure for Groundbreaking Research in Engineering and Science (TIGRESS) and the High Performance Computing Modernization Program (HPCMP) of the US Department of Defense provided the computational resources.

References

ARTICLE SECTIONS
Jump To

This article references 80 other publications.

  1. 1
    Roos, B. O. The Complete Active Space Self-Consistent Field Method and its Applications in Electronic Structure Calculations. Adv. Chem. Phys. 1987, 69, 399445,  DOI: 10.1002/9780470142943.ch7
  2. 2
    Bofill, J. M.; Pulay, P. The unrestricted natural orbital–complete active space (UNO–CAS) method: An inexpensive alternative to the complete active space–self-consistent-field (CAS–SCF) method. J. Chem. Phys. 1989, 90 (7), 36373646,  DOI: 10.1063/1.455822
  3. 3
    Veryazov, V.; Malmqvist, P. Å.; Roos, B. O. How to select active space for multiconfigurational quantum chemistry?. Int. J. Quantum Chem. 2011, 111 (13), 33293338,  DOI: 10.1002/qua.23068
  4. 4
    Jensen, H. J. A.; Jørgensen, P.; Ågren, H.; Olsen, J. Second-order Møller–Plesset perturbation theory as a configuration and orbital generator in multiconfiguration self-consistent field calculations. J. Chem. Phys. 1988, 88 (6), 38343839,  DOI: 10.1063/1.453884
  5. 5
    Wouters, S.; Bogaerts, T.; Van Der Voort, P.; Van Speybroeck, V.; Van Neck, D. Communication: DMRG-SCF study of the singlet, triplet, and quintet states of oxo-Mn(Salen). J. Chem. Phys. 2014, 140 (24), 241103,  DOI: 10.1063/1.4885815
  6. 6
    Khedkar, A.; Roemelt, M. Active Space Selection Based on Natural Orbital Occupation Numbers from n-Electron Valence Perturbation Theory. J. Chem. Theory Comput. 2019, 15 (6), 35223536,  DOI: 10.1021/acs.jctc.8b01293
  7. 7
    Sayfutyarova, E. R.; Sun, Q.; Chan, G. K.-L.; Knizia, G. Automated Construction of Molecular Active Spaces from Atomic Valence Orbitals. J. Chem. Theory Comput. 2017, 13 (9), 40634078,  DOI: 10.1021/acs.jctc.7b00128
  8. 8
    Stein, C. J.; Reiher, M. autoCAS: A Program for Fully Automated Multiconfigurational Calculations. J. Comput. Chem. 2019, 40, 22162226,  DOI: 10.1002/jcc.25869
  9. 9
    Jeong, W.; Stoneburner, S. J.; King, D.; Li, R.; Walker, A.; Lindh, R.; Gagliardi, L. Automation of Active Space Selection for Multireference Methods via Machine Learning on Chemical Bond Dissociation. J. Chem. Theory Comput. 2020, 16 (4), 23892399,  DOI: 10.1021/acs.jctc.9b01297
  10. 10
    Golub, P.; Antalik, A.; Veis, L.; Brabec, J. Machine Learning-Assisted Selection of Active Spaces for Strongly Correlated Transition Metal Systems. J. Chem. Theory Comput. 2021, 17 (10), 60536072,  DOI: 10.1021/acs.jctc.1c00235
  11. 11
    Han, R.; Luber, S. Complete active space analysis of a reaction pathway: Investigation of the oxygen–oxygen bond formation. J. Comput. Chem. 2020, 41 (17), 15861597,  DOI: 10.1002/jcc.26201
  12. 12
    Stein, C. J.; Reiher, M. Automated Identification of Relevant Frontier Orbitals for Chemical Compounds and Processes. Chimia 2017, 71 (4), 170,  DOI: 10.2533/chimia.2017.170
  13. 13
    Libisch, F.; Cheng, J.; Carter, E. A. Electron-Transfer-Induced Dissociation of H2 on Gold Nanoparticles: Excited-State Potential Energy Surfaces via Embedded Correlated Wavefunction Theory. Z. Phys. Chem. 2013, 227 (11), 14551466,  DOI: 10.1524/zpch.2013.0406
  14. 14
    Libisch, F.; Krauter, C. M.; Carter, E. A. Corrigendum to: Plasmon-Driven Dissociation of H2 on Gold Nanoclusters. Z. Phys. Chem. 2016, 230 (1), 131132,  DOI: 10.1515/zpch-2015-5001
  15. 15
    Mukherjee, S.; Libisch, F.; Large, N.; Neumann, O.; Brown, L. V.; Cheng, J.; Lassiter, J. B.; Carter, E. A.; Nordlander, P.; Halas, N. J. Hot Electrons Do the Impossible: Plasmon-Induced Dissociation of H2 on Au. Nano Lett. 2013, 13 (1), 240247,  DOI: 10.1021/nl303940z
  16. 16
    Martirez, J. M. P.; Carter, E. A. Excited-State N2 Dissociation Pathway on Fe-Functionalized Au. J. Am. Chem. Soc. 2017, 139 (12), 43904398,  DOI: 10.1021/jacs.6b12301
  17. 17
    Martirez, J. M. P.; Carter, E. A. Prediction of a low-temperature N2 dissociation catalyst exploiting near-IR–to–visible light nanoplasmonics. Sci. Adv. 2017, 3 (12), eaao4710  DOI: 10.1126/sciadv.aao4710
  18. 18
    Yuan, Y.; Zhou, L.; Robatjazi, H.; Bao, J. L.; Zhou, J.; Bayles, A.; Yuan, L.; Lou, M.; Lou, M.; Khatiwada, S. Earth-abundant photocatalyst for H2 generation from NH3 with light-emitting diode illumination. Science 2022, 378 (6622), 889893,  DOI: 10.1126/science.abn5636
  19. 19
    Bao, J. L.; Carter, E. A. Surface-Plasmon-Induced Ammonia Decomposition on Copper: Excited-State Reaction Pathways Revealed by Embedded Correlated Wavefunction Theory. ACS Nano 2019, 13 (9), 99449957,  DOI: 10.1021/acsnano.9b05030
  20. 20
    Bao, J. L.; Carter, E. A. Rationalizing the Hot-Carrier-Mediated Reaction Mechanisms and Kinetics for Ammonia Decomposition on Ruthenium-Doped Copper Nanoparticles. J. Am. Chem. Soc. 2019, 141 (34), 1332013323,  DOI: 10.1021/jacs.9b06804
  21. 21
    Zhou, L.; Martirez, J. M. P.; Finzel, J.; Zhang, C.; Swearer, D. F.; Tian, S.; Robatjazi, H.; Lou, M.; Dong, L.; Henderson, L. Light-driven methane dry reforming with single atomic site antenna-reactor plasmonic photocatalysts. Nat. Energy 2020, 5 (1), 6170,  DOI: 10.1038/s41560-019-0517-9
  22. 22
    Spata, V. A.; Carter, E. A. Mechanistic Insights into Photocatalyzed Hydrogen Desorption from Palladium Surfaces Assisted by Localized Surface Plasmon Resonances. ACS Nano 2018, 12 (4), 35123522,  DOI: 10.1021/acsnano.8b00352
  23. 23
    Huang, C.; Pavone, M.; Carter, E. A. Quantum mechanical embedding theory based on a unique embedding potential. J. Chem. Phys. 2011, 134 (15), 154110,  DOI: 10.1063/1.3577516
  24. 24
    Yu, K.; Libisch, F.; Carter, E. A. Implementation of density functional embedding theory within the projector-augmented-wave method and applications to semiconductor defect states. J. Chem. Phys. 2015, 143 (10), 102806,  DOI: 10.1063/1.4922260
  25. 25
    Libisch, F.; Huang, C.; Carter, E. A. Embedded Correlated Wavefunction Schemes: Theory and Applications. Acc. Chem. Res. 2014, 47 (9), 27682775,  DOI: 10.1021/ar500086h
  26. 26
    Tóth, Z.; Pulay, P. Comparison of Methods for Active Orbital Selection in Multiconfigurational Calculations. J. Chem. Theory Comput. 2020, 16 (12), 73287341,  DOI: 10.1021/acs.jctc.0c00123
  27. 27
    Monkhorst, H. J. Hartree-Fock density of states for extended systems. Phys. Rev. B: Condens. Matter Mater. Phys. 1979, 20 (4), 15041513,  DOI: 10.1103/PhysRevB.20.1504
  28. 28
    McAdon, M. H.; Goddard, W. A. Charge density waves, spin density waves, and Peierls distortions in one-dimensional metals. I. Hartree–Fock studies of Cu, Ag, Au, Li, and Na. J. Chem. Phys. 1988, 88 (1), 277302,  DOI: 10.1063/1.454654
  29. 29
    Klüner, T.; Govind, N.; Wang, Y. A.; Carter, E. A. Prediction of Electronic Excited States of Adsorbates on Metal Surfaces from First Principles. Phys. Rev. Lett. 2001, 86 (26), 59545957,  DOI: 10.1103/PhysRevLett.86.5954
  30. 30
    Klüner, T.; Govind, N.; Wang, Y. A.; Carter, E. A. Periodic density functional embedding theory for complete active space self-consistent field and configuration interaction calculations: Ground and excited states. J. Chem. Phys. 2002, 116 (1), 4254,  DOI: 10.1063/1.1420748
  31. 31
    Hirano, T.; Nagashima, U. Ro-vibrational properties of FeCO in the X̃ 3Σ- and ã 5Σ- electronic states: A computational molecular spectroscopy study. J. Mol. Spectrosc. 2015, 314, 3547,  DOI: 10.1016/j.jms.2015.05.007
  32. 32
    Cimpoesu, F.; Dahan, F.; Ladeira, S.; Ferbinteanu, M.; Costes, J.-P. Chiral Crystallization of a Heterodinuclear Ni-Ln Series: Comprehensive Analysis of the Magnetic Properties. Inorg. Chem. 2012, 51 (21), 1127911293,  DOI: 10.1021/ic3001784
  33. 33
    Andersson, K.; Malmqvist, P. Å.; Roos, B. O. Second-order perturbation theory with a complete active space self-consistent field reference function. J. Chem. Phys. 1992, 96 (2), 12181226,  DOI: 10.1063/1.462209
  34. 34
    Martirez, J. M. P.; Bao, J. L.; Carter, E. A. First-Principles Insights into Plasmon-Induced Catalysis. Annu. Rev. Phys. Chem. 2021, 72 (1), 99119,  DOI: 10.1146/annurev-physchem-061020-053501
  35. 35
    Kresse, G.; Hafner, J. Ab initio molecular dynamics for liquid metals. Phys. Rev. B: Condens. Matter Mater. Phys. 1993, 47 (1), 558561,  DOI: 10.1103/PhysRevB.47.558
  36. 36
    Kresse, G.; Furthmuller, J. Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set. Phys. Rev. B: Condens. Matter Mater. Phys. 1996, 54 (16), 1116911186,  DOI: 10.1103/PhysRevB.54.11169
  37. 37
    Kresse, G.; Furthmüller, J. Efficiency of ab-initio total energy calculations for metals and semiconductors using a plane-wave basis set. Comput. Mater. Sci. 1996, 6 (1), 1550,  DOI: 10.1016/0927-0256(96)00008-0
  38. 38
    Perdew, J. P.; Burke, K.; Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 1996, 77 (18), 38653868,  DOI: 10.1103/PhysRevLett.77.3865
  39. 39
    Singh-Miller, N. E.; Marzari, N. Surface energies, work functions, and surface relaxations of low-index metallic surfaces from first principles. Phys. Rev. B: Condens. Matter Mater. Phys. 2009, 80 (23), 235407,  DOI: 10.1103/PhysRevB.80.235407
  40. 40
    Makov, G.; Payne, M. C. Periodic boundary conditions in ab initio calculations. Phys. Rev. B: Condens. Matter Mater. Phys. 1995, 51 (7), 40144022,  DOI: 10.1103/PhysRevB.51.4014
  41. 41
    Monkhorst, H. J.; Pack, J. D. Special points for Brillouin-zone integrations. Phys. Rev. B: Solid State 1976, 13 (12), 51885192,  DOI: 10.1103/PhysRevB.13.5188
  42. 42
    Methfessel, M.; Paxton, A. T. High-precision sampling for Brillouin-zone integration in metals. Phys. Rev. B: Condens. Matter Mater. Phys. 1989, 40 (6), 36163621,  DOI: 10.1103/PhysRevB.40.3616
  43. 43
    Dutta, B. N.; Dayal, B. Lattice Constants and Thermal Expansion of Palladium and Tungsten up to 878 °C by X-Ray Method. Phys. Status Solidi B 1963, 3 (12), 22532259,  DOI: 10.1002/pssb.19630031207
  44. 44
    Jonsson, H.; Mills, G.; Jacobsen, K. W. Nudged elastic band method for finding minimum energy paths of transitions. In Classical and Quantum Dynamics in Condensed Phase Simulations; World Scientific, 1998; pp 385404.
  45. 45
    Henkelman, G.; Uberuaga, B. P.; Jónsson, H. A climbing image nudged elastic band method for finding saddle points and minimum energy paths. J. Chem. Phys. 2000, 113 (22), 99019904,  DOI: 10.1063/1.1329672
  46. 46
    Yu, K.; Krauter, C. M.; Dieterich, J. M.; Carter, E. A. Density and Potential Functional Embedding: Theory and Practice. In Fragmentation: Toward Accurate Calculations on Complex Molecular Systems; Gordon, M., Ed.; John Wiley & Sons, 2017; pp 81117.
  47. 47
    Yu, K.; Carter, E. A. VASP density functional embedding theory. https://github.com/EACcodes/VASPEmbedding (accessed May 16, 2022).
  48. 48
    Wu, Q.; Yang, W. A direct optimization method for calculating density functionals and exchange–correlation potentials from electron densities. J. Chem. Phys. 2003, 118 (6), 24982509,  DOI: 10.1063/1.1535422
  49. 49
    Krauter, C. M.; Carter, E. A. Embedding Integral Generator. https://github.com/EACcodes/EmbeddingIntegralGenerator (accessed May 16, 2022).
  50. 50
    Werner, H.-J.; Knowles, P. J.; Knizia, G.; Manby, F. R.; Schütz, M. Molpro: a general-purpose quantum chemistry program package. Wiley Interdiscip. Rev. Comput. Mol. Sci. 2012, 2 (2), 242253,  DOI: 10.1002/wcms.82
  51. 51
    Werner, H.-J.; Knowles, P. J.; Manby, F. R.; Black, J. A.; Doll, K.; Heßelmann, A.; Kats, D.; Köhn, A.; Korona, T.; Kreplin, D. A. The Molpro quantum chemistry package. J. Chem. Phys. 2020, 152 (14), 144107,  DOI: 10.1063/5.0005081
  52. 52
    Werner, H.-J.; Knowles, P. J.; Knizia, G.; Manby, F. R.; Schütz, M.; Celani, P.; Györffy, W.; Kats, D.; Korona, T.; Lindh, R.; MOLPRO , version 2021.2, a package of ab initio programs, 2021. https://www.molpro.net.
  53. 53
    Ghigo, G.; Roos, B. O.; Malmqvist, P.-Å. A modified definition of the zeroth-order Hamiltonian in multiconfigurational perturbation theory (CASPT2). Chem. Phys. Lett. 2004, 396 (1–3), 142149,  DOI: 10.1016/j.cplett.2004.08.032
  54. 54
    Roos, B. O.; Andersson, K. Multiconfigurational perturbation theory with level shift ─ the Cr2 potential revisited. Chem. Phys. Lett. 1995, 245 (2–3), 215223,  DOI: 10.1016/0009-2614(95)01010-7
  55. 55
    Wen, X.; Martirez, J. M. P.; Carter, E. A. Plasmon-driven ammonia decomposition on Pd(111): Hole transfer’s role in changing rate-limiting steps. ACS Catal. 2024, 14, 95399553,  DOI: 10.1021/acscatal.4c01869
  56. 56
    Stradella, L. Heats of adsorption of different gases on polycrystalline transition metals. Adsorpt. Sci. Technol. 1992, 9 (3), 190198,  DOI: 10.1177/026361749200900304
  57. 57
    Peterson, K. A.; Figgen, D.; Dolg, M.; Stoll, H. Energy-consistent relativistic pseudopotentials and correlation consistent basis sets for the 4d elements Y–Pd. J. Chem. Phys. 2007, 126 (12), 124101,  DOI: 10.1063/1.2647019
  58. 58
    Dunning, T. H., Jr. Gaussian basis sets for use in correlated molecular calculations. I. The atoms boron through neon and hydrogen. J. Chem. Phys. 1989, 90 (2), 10071023,  DOI: 10.1063/1.456153
  59. 59
    Kendall, R. A.; Dunning, T. H., Jr.; Harrison, R. J. Electron affinities of the first-row atoms revisited. Systematic basis sets and wave functions. J. Chem. Phys. 1992, 96 (9), 67966806,  DOI: 10.1063/1.462569
  60. 60
    Peterson, K. A.; Dunning, T. H., Jr. Accurate correlation consistent basis sets for molecular core–valence correlation effects: The second row atoms Al–Ar, and the first row atoms B–Ne revisited. J. Chem. Phys. 2002, 117 (23), 1054810560,  DOI: 10.1063/1.1520138
  61. 61
    Pritchard, B. P.; Altarawy, D.; Didier, B.; Gibson, T. D.; Windus, T. L. New Basis Set Exchange: An Open, Up-to-Date Resource for the Molecular Sciences Community. J. Chem. Inf. Model. 2019, 59 (11), 48144820,  DOI: 10.1021/acs.jcim.9b00725
  62. 62
    Hill, J. G.; Peterson, K. A. Gaussian basis sets for use in correlated molecular calculations. XI. Pseudopotential-based and all-electron relativistic basis sets for alkali metal (K–Fr) and alkaline earth (Ca–Ra) elements. J. Chem. Phys. 2017, 147 (24), 244106,  DOI: 10.1063/1.5010587
  63. 63
    Balabanov, N. B.; Peterson, K. A. Systematically convergent basis sets for transition metals. I. All-electron correlation consistent basis sets for the 3d elements Sc–Zn. J. Chem. Phys. 2005, 123 (6), 064107,  DOI: 10.1063/1.1998907
  64. 64
    Jansen, G.; Hess, B. A. Revision of the Douglas-Kroll transformation. Phys. Rev. A 1989, 39 (11), 60166017,  DOI: 10.1103/PhysRevA.39.6016
  65. 65
    Douglas, M.; Kroll, N. M. Quantum electrodynamical corrections to the fine structure of helium. Ann. Phys. 1974, 82 (1), 89155,  DOI: 10.1016/0003-4916(74)90333-9
  66. 66
    Silva-Junior, M. R.; Schreiber, M.; Sauer, S. P. A.; Thiel, W. Benchmarks of electronically excited states: Basis set effects on CASPT2 results. J. Chem. Phys. 2010, 133 (17), 174318,  DOI: 10.1063/1.3499598
  67. 67
    Kánnár, D.; Tajti, A.; Szalay, P. G. Accuracy of Coupled Cluster Excitation Energies in Diffuse Basis Sets. J. Chem. Theory Comput. 2017, 13 (1), 202209,  DOI: 10.1021/acs.jctc.6b00875
  68. 68
    Jansen, H. B.; Ros, P. Non-empirical molecular orbital calculations on the protonation of carbon monoxide. Chem. Phys. Lett. 1969, 3 (3), 140143,  DOI: 10.1016/0009-2614(69)80118-1
  69. 69
    Liu, B.; McLean, A. D. Accurate calculation of the attractive interaction of two ground state helium atoms. J. Chem. Phys. 1973, 59 (8), 45574558,  DOI: 10.1063/1.1680654
  70. 70
    Gutowski, M.; Van Lenthe, J. H.; Verbeek, J.; Van Duijneveldt, F. B.; Chałasinski, G. The basis set superposition error in correlated electronic structure calculations. Chem. Phys. Lett. 1986, 124 (4), 370375,  DOI: 10.1016/0009-2614(86)85036-9
  71. 71
    Simon, S.; Duran, M.; Dannenberg, J. J. How does basis set superposition error change the potential surfaces for hydrogen-bonded dimers?. J. Chem. Phys. 1996, 105 (24), 1102411031,  DOI: 10.1063/1.472902
  72. 72
    Boys, S. F.; Bernardi, F. The calculation of small molecular interactions by the differences of separate total energies. Some procedures with reduced errors. Mol. Phys. 1970, 19 (4), 553566,  DOI: 10.1080/00268977000101561
  73. 73
    Gray, M.; Bowling, P. E.; Herbert, J. M. Systematic Evaluation of Counterpoise Correction in Density Functional Theory. J. Chem. Theory Comput. 2022, 18 (11), 67426756,  DOI: 10.1021/acs.jctc.2c00883
  74. 74
    Mentel, Ł. M.; Baerends, E. J. Can the Counterpoise Correction for Basis Set Superposition Effect Be Justified?. J. Chem. Theory Comput. 2014, 10 (1), 252267,  DOI: 10.1021/ct400990u
  75. 75
    van Duijneveldt, F. B.; van Duijneveldt-van de Rijdt, J. G. C. M.; van Lenthe, J. H. State of the Art in Counterpoise Theory. Chem. Rev. 1994, 94 (7), 18731885,  DOI: 10.1021/cr00031a007
  76. 76
    Burns, L. A.; Marshall, M. S.; Sherrill, C. D. Comparing Counterpoise-Corrected, Uncorrected, and Averaged Binding Energies for Benchmarking Noncovalent Interactions. J. Chem. Theory Comput. 2014, 10 (1), 4957,  DOI: 10.1021/ct400149j
  77. 77
    Helgaker, T.; Klopper, W.; Koch, H.; Noga, J. Basis-set convergence of correlated calculations on water. J. Chem. Phys. 1997, 106 (23), 96399646,  DOI: 10.1063/1.473863
  78. 78
    Wei, Z.; Martirez, J. M. P.; Carter, E. A. Introducing the embedded random phase approximation: H2 dissociative adsorption on Cu(111) as an exemplar. J. Chem. Phys. 2023, 159 (19), 194108,  DOI: 10.1063/5.0181229
  79. 79
    Zhou, L.; Swearer, D. F.; Zhang, C.; Robatjazi, H.; Zhao, H.; Henderson, L.; Dong, L.; Christopher, P.; Carter, E. A.; Nordlander, P. Quantifying hot carrier and thermal contributions in plasmonic photocatalysis. Science 2018, 362 (6410), 6972,  DOI: 10.1126/science.aat6967
  80. 80
    Finley, J.; Malmqvist, P.-Å.; Roos, B. O.; Serrano-Andrés, L. The multi-state CASPT2 method. Chem. Phys. Lett. 1998, 288 (2–4), 299306,  DOI: 10.1016/S0009-2614(98)00252-8

Cited By

ARTICLE SECTIONS
Jump To

This article has not yet been cited by other publications.

  • Abstract

    Figure 1

    Figure 1. (a) Four-layer 5 × 5 Pd(111) (Pd100) periodic slab model. Top-down view of the (b) Pd10 cluster, (c) Pd12 cluster, and (d) Pd14 cluster carved out from the Pd100 periodic slab. Subsurface atoms are faded out. Isosurface plots (yellow: +1.2 V, cyan: −1.2 V) of the optimized embedding potentials generated for the (e) Pd10 cluster in its Pd90 environment, (f) Pd12 cluster in its Pd88 environment, and (g) Pd14 cluster in its Pd86 environment.

    Figure 2

    Figure 2. Effect of cluster size for NH3(gas) → *NH3 → *NH2 + *H. We calculate the emb-LDA energies (defined in eq 4) for Pd10, Pd12, and Pd14 clusters. Numerical indices correspond to structures shown in Figure S1.

    Figure 3

    Figure 3. Basis set effect for NH3(gas) → *NH3 → *NH2 + *H at the emb-CASPT2 level using an embedded Pd10 cluster. Basis sets for Pd and adsorbates are listed as Pd basis/adsorbate basis in the legend. DFT-corrected Einit,expt is the DFT-corrected electronic energy based on the experimental enthalpy of NH3 adsorption on Pd powder. See detailed explanation in the text.

    Figure 4

    Figure 4. CP corrections for NH3(gas) → *NH3 → *NH2 + *H on embedded Pd10. Combined CP corrections of the adsorbates and the Pd10 cluster are plotted for emb-CASPT2 with various basis sets (denoted as Pd basis/NH3 basis in the legends) and DFT-PBE with the AVDZ basis set.

    Figure 5

    Figure 5. Emb-CASPT2 results using creeping to obtain CASSCF orbitals for (a) NH3(gas) → *NH3 → *NH2 + *H, (b) *NH2 + *H → *NH+ 2*H, and (c) *NH → *N+ *H. The corresponding NOs in the ASs using creeping are given in Figures S7–S9, while the ones from merging are shown in Figures S10–S12 in the Supporting Information. All calculations utilize the Pd10 cluster model.

    Figure 6

    Figure 6. Ground- and excited-state MEP energetics as calculated with emb-CASPT2 for NH3 (gas) → *NH3 → * NH2 + *H. Red arrows represent vertical excitation of S0 → SN-1, where SN-1 represents the highest excited state. Green arrows represent reaction barriers for S0 or SN–1. Each panel represents different numbers of states included in the N-state emb-SA-CASSCF calculations: N = (a) 5, (b) 7, (c) 9, (d) 11, (e) 15, and (f) 17.

  • References

    ARTICLE SECTIONS
    Jump To

    This article references 80 other publications.

    1. 1
      Roos, B. O. The Complete Active Space Self-Consistent Field Method and its Applications in Electronic Structure Calculations. Adv. Chem. Phys. 1987, 69, 399445,  DOI: 10.1002/9780470142943.ch7
    2. 2
      Bofill, J. M.; Pulay, P. The unrestricted natural orbital–complete active space (UNO–CAS) method: An inexpensive alternative to the complete active space–self-consistent-field (CAS–SCF) method. J. Chem. Phys. 1989, 90 (7), 36373646,  DOI: 10.1063/1.455822
    3. 3
      Veryazov, V.; Malmqvist, P. Å.; Roos, B. O. How to select active space for multiconfigurational quantum chemistry?. Int. J. Quantum Chem. 2011, 111 (13), 33293338,  DOI: 10.1002/qua.23068
    4. 4
      Jensen, H. J. A.; Jørgensen, P.; Ågren, H.; Olsen, J. Second-order Møller–Plesset perturbation theory as a configuration and orbital generator in multiconfiguration self-consistent field calculations. J. Chem. Phys. 1988, 88 (6), 38343839,  DOI: 10.1063/1.453884
    5. 5
      Wouters, S.; Bogaerts, T.; Van Der Voort, P.; Van Speybroeck, V.; Van Neck, D. Communication: DMRG-SCF study of the singlet, triplet, and quintet states of oxo-Mn(Salen). J. Chem. Phys. 2014, 140 (24), 241103,  DOI: 10.1063/1.4885815
    6. 6
      Khedkar, A.; Roemelt, M. Active Space Selection Based on Natural Orbital Occupation Numbers from n-Electron Valence Perturbation Theory. J. Chem. Theory Comput. 2019, 15 (6), 35223536,  DOI: 10.1021/acs.jctc.8b01293
    7. 7
      Sayfutyarova, E. R.; Sun, Q.; Chan, G. K.-L.; Knizia, G. Automated Construction of Molecular Active Spaces from Atomic Valence Orbitals. J. Chem. Theory Comput. 2017, 13 (9), 40634078,  DOI: 10.1021/acs.jctc.7b00128
    8. 8
      Stein, C. J.; Reiher, M. autoCAS: A Program for Fully Automated Multiconfigurational Calculations. J. Comput. Chem. 2019, 40, 22162226,  DOI: 10.1002/jcc.25869
    9. 9
      Jeong, W.; Stoneburner, S. J.; King, D.; Li, R.; Walker, A.; Lindh, R.; Gagliardi, L. Automation of Active Space Selection for Multireference Methods via Machine Learning on Chemical Bond Dissociation. J. Chem. Theory Comput. 2020, 16 (4), 23892399,  DOI: 10.1021/acs.jctc.9b01297
    10. 10
      Golub, P.; Antalik, A.; Veis, L.; Brabec, J. Machine Learning-Assisted Selection of Active Spaces for Strongly Correlated Transition Metal Systems. J. Chem. Theory Comput. 2021, 17 (10), 60536072,  DOI: 10.1021/acs.jctc.1c00235
    11. 11
      Han, R.; Luber, S. Complete active space analysis of a reaction pathway: Investigation of the oxygen–oxygen bond formation. J. Comput. Chem. 2020, 41 (17), 15861597,  DOI: 10.1002/jcc.26201
    12. 12
      Stein, C. J.; Reiher, M. Automated Identification of Relevant Frontier Orbitals for Chemical Compounds and Processes. Chimia 2017, 71 (4), 170,  DOI: 10.2533/chimia.2017.170
    13. 13
      Libisch, F.; Cheng, J.; Carter, E. A. Electron-Transfer-Induced Dissociation of H2 on Gold Nanoparticles: Excited-State Potential Energy Surfaces via Embedded Correlated Wavefunction Theory. Z. Phys. Chem. 2013, 227 (11), 14551466,  DOI: 10.1524/zpch.2013.0406
    14. 14
      Libisch, F.; Krauter, C. M.; Carter, E. A. Corrigendum to: Plasmon-Driven Dissociation of H2 on Gold Nanoclusters. Z. Phys. Chem. 2016, 230 (1), 131132,  DOI: 10.1515/zpch-2015-5001
    15. 15
      Mukherjee, S.; Libisch, F.; Large, N.; Neumann, O.; Brown, L. V.; Cheng, J.; Lassiter, J. B.; Carter, E. A.; Nordlander, P.; Halas, N. J. Hot Electrons Do the Impossible: Plasmon-Induced Dissociation of H2 on Au. Nano Lett. 2013, 13 (1), 240247,  DOI: 10.1021/nl303940z
    16. 16
      Martirez, J. M. P.; Carter, E. A. Excited-State N2 Dissociation Pathway on Fe-Functionalized Au. J. Am. Chem. Soc. 2017, 139 (12), 43904398,  DOI: 10.1021/jacs.6b12301
    17. 17
      Martirez, J. M. P.; Carter, E. A. Prediction of a low-temperature N2 dissociation catalyst exploiting near-IR–to–visible light nanoplasmonics. Sci. Adv. 2017, 3 (12), eaao4710  DOI: 10.1126/sciadv.aao4710
    18. 18
      Yuan, Y.; Zhou, L.; Robatjazi, H.; Bao, J. L.; Zhou, J.; Bayles, A.; Yuan, L.; Lou, M.; Lou, M.; Khatiwada, S. Earth-abundant photocatalyst for H2 generation from NH3 with light-emitting diode illumination. Science 2022, 378 (6622), 889893,  DOI: 10.1126/science.abn5636
    19. 19
      Bao, J. L.; Carter, E. A. Surface-Plasmon-Induced Ammonia Decomposition on Copper: Excited-State Reaction Pathways Revealed by Embedded Correlated Wavefunction Theory. ACS Nano 2019, 13 (9), 99449957,  DOI: 10.1021/acsnano.9b05030
    20. 20
      Bao, J. L.; Carter, E. A. Rationalizing the Hot-Carrier-Mediated Reaction Mechanisms and Kinetics for Ammonia Decomposition on Ruthenium-Doped Copper Nanoparticles. J. Am. Chem. Soc. 2019, 141 (34), 1332013323,  DOI: 10.1021/jacs.9b06804
    21. 21
      Zhou, L.; Martirez, J. M. P.; Finzel, J.; Zhang, C.; Swearer, D. F.; Tian, S.; Robatjazi, H.; Lou, M.; Dong, L.; Henderson, L. Light-driven methane dry reforming with single atomic site antenna-reactor plasmonic photocatalysts. Nat. Energy 2020, 5 (1), 6170,  DOI: 10.1038/s41560-019-0517-9
    22. 22
      Spata, V. A.; Carter, E. A. Mechanistic Insights into Photocatalyzed Hydrogen Desorption from Palladium Surfaces Assisted by Localized Surface Plasmon Resonances. ACS Nano 2018, 12 (4), 35123522,  DOI: 10.1021/acsnano.8b00352
    23. 23
      Huang, C.; Pavone, M.; Carter, E. A. Quantum mechanical embedding theory based on a unique embedding potential. J. Chem. Phys. 2011, 134 (15), 154110,  DOI: 10.1063/1.3577516
    24. 24
      Yu, K.; Libisch, F.; Carter, E. A. Implementation of density functional embedding theory within the projector-augmented-wave method and applications to semiconductor defect states. J. Chem. Phys. 2015, 143 (10), 102806,  DOI: 10.1063/1.4922260
    25. 25
      Libisch, F.; Huang, C.; Carter, E. A. Embedded Correlated Wavefunction Schemes: Theory and Applications. Acc. Chem. Res. 2014, 47 (9), 27682775,  DOI: 10.1021/ar500086h
    26. 26
      Tóth, Z.; Pulay, P. Comparison of Methods for Active Orbital Selection in Multiconfigurational Calculations. J. Chem. Theory Comput. 2020, 16 (12), 73287341,  DOI: 10.1021/acs.jctc.0c00123
    27. 27
      Monkhorst, H. J. Hartree-Fock density of states for extended systems. Phys. Rev. B: Condens. Matter Mater. Phys. 1979, 20 (4), 15041513,  DOI: 10.1103/PhysRevB.20.1504
    28. 28
      McAdon, M. H.; Goddard, W. A. Charge density waves, spin density waves, and Peierls distortions in one-dimensional metals. I. Hartree–Fock studies of Cu, Ag, Au, Li, and Na. J. Chem. Phys. 1988, 88 (1), 277302,  DOI: 10.1063/1.454654
    29. 29
      Klüner, T.; Govind, N.; Wang, Y. A.; Carter, E. A. Prediction of Electronic Excited States of Adsorbates on Metal Surfaces from First Principles. Phys. Rev. Lett. 2001, 86 (26), 59545957,  DOI: 10.1103/PhysRevLett.86.5954
    30. 30
      Klüner, T.; Govind, N.; Wang, Y. A.; Carter, E. A. Periodic density functional embedding theory for complete active space self-consistent field and configuration interaction calculations: Ground and excited states. J. Chem. Phys. 2002, 116 (1), 4254,  DOI: 10.1063/1.1420748
    31. 31
      Hirano, T.; Nagashima, U. Ro-vibrational properties of FeCO in the X̃ 3Σ- and ã 5Σ- electronic states: A computational molecular spectroscopy study. J. Mol. Spectrosc. 2015, 314, 3547,  DOI: 10.1016/j.jms.2015.05.007
    32. 32
      Cimpoesu, F.; Dahan, F.; Ladeira, S.; Ferbinteanu, M.; Costes, J.-P. Chiral Crystallization of a Heterodinuclear Ni-Ln Series: Comprehensive Analysis of the Magnetic Properties. Inorg. Chem. 2012, 51 (21), 1127911293,  DOI: 10.1021/ic3001784
    33. 33
      Andersson, K.; Malmqvist, P. Å.; Roos, B. O. Second-order perturbation theory with a complete active space self-consistent field reference function. J. Chem. Phys. 1992, 96 (2), 12181226,  DOI: 10.1063/1.462209
    34. 34
      Martirez, J. M. P.; Bao, J. L.; Carter, E. A. First-Principles Insights into Plasmon-Induced Catalysis. Annu. Rev. Phys. Chem. 2021, 72 (1), 99119,  DOI: 10.1146/annurev-physchem-061020-053501
    35. 35
      Kresse, G.; Hafner, J. Ab initio molecular dynamics for liquid metals. Phys. Rev. B: Condens. Matter Mater. Phys. 1993, 47 (1), 558561,  DOI: 10.1103/PhysRevB.47.558
    36. 36
      Kresse, G.; Furthmuller, J. Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set. Phys. Rev. B: Condens. Matter Mater. Phys. 1996, 54 (16), 1116911186,  DOI: 10.1103/PhysRevB.54.11169
    37. 37
      Kresse, G.; Furthmüller, J. Efficiency of ab-initio total energy calculations for metals and semiconductors using a plane-wave basis set. Comput. Mater. Sci. 1996, 6 (1), 1550,  DOI: 10.1016/0927-0256(96)00008-0
    38. 38
      Perdew, J. P.; Burke, K.; Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 1996, 77 (18), 38653868,  DOI: 10.1103/PhysRevLett.77.3865
    39. 39
      Singh-Miller, N. E.; Marzari, N. Surface energies, work functions, and surface relaxations of low-index metallic surfaces from first principles. Phys. Rev. B: Condens. Matter Mater. Phys. 2009, 80 (23), 235407,  DOI: 10.1103/PhysRevB.80.235407
    40. 40
      Makov, G.; Payne, M. C. Periodic boundary conditions in ab initio calculations. Phys. Rev. B: Condens. Matter Mater. Phys. 1995, 51 (7), 40144022,  DOI: 10.1103/PhysRevB.51.4014
    41. 41
      Monkhorst, H. J.; Pack, J. D. Special points for Brillouin-zone integrations. Phys. Rev. B: Solid State 1976, 13 (12), 51885192,  DOI: 10.1103/PhysRevB.13.5188
    42. 42
      Methfessel, M.; Paxton, A. T. High-precision sampling for Brillouin-zone integration in metals. Phys. Rev. B: Condens. Matter Mater. Phys. 1989, 40 (6), 36163621,  DOI: 10.1103/PhysRevB.40.3616
    43. 43
      Dutta, B. N.; Dayal, B. Lattice Constants and Thermal Expansion of Palladium and Tungsten up to 878 °C by X-Ray Method. Phys. Status Solidi B 1963, 3 (12), 22532259,  DOI: 10.1002/pssb.19630031207
    44. 44
      Jonsson, H.; Mills, G.; Jacobsen, K. W. Nudged elastic band method for finding minimum energy paths of transitions. In Classical and Quantum Dynamics in Condensed Phase Simulations; World Scientific, 1998; pp 385404.
    45. 45
      Henkelman, G.; Uberuaga, B. P.; Jónsson, H. A climbing image nudged elastic band method for finding saddle points and minimum energy paths. J. Chem. Phys. 2000, 113 (22), 99019904,  DOI: 10.1063/1.1329672
    46. 46
      Yu, K.; Krauter, C. M.; Dieterich, J. M.; Carter, E. A. Density and Potential Functional Embedding: Theory and Practice. In Fragmentation: Toward Accurate Calculations on Complex Molecular Systems; Gordon, M., Ed.; John Wiley & Sons, 2017; pp 81117.
    47. 47
      Yu, K.; Carter, E. A. VASP density functional embedding theory. https://github.com/EACcodes/VASPEmbedding (accessed May 16, 2022).
    48. 48
      Wu, Q.; Yang, W. A direct optimization method for calculating density functionals and exchange–correlation potentials from electron densities. J. Chem. Phys. 2003, 118 (6), 24982509,  DOI: 10.1063/1.1535422
    49. 49
      Krauter, C. M.; Carter, E. A. Embedding Integral Generator. https://github.com/EACcodes/EmbeddingIntegralGenerator (accessed May 16, 2022).
    50. 50
      Werner, H.-J.; Knowles, P. J.; Knizia, G.; Manby, F. R.; Schütz, M. Molpro: a general-purpose quantum chemistry program package. Wiley Interdiscip. Rev. Comput. Mol. Sci. 2012, 2 (2), 242253,  DOI: 10.1002/wcms.82
    51. 51
      Werner, H.-J.; Knowles, P. J.; Manby, F. R.; Black, J. A.; Doll, K.; Heßelmann, A.; Kats, D.; Köhn, A.; Korona, T.; Kreplin, D. A. The Molpro quantum chemistry package. J. Chem. Phys. 2020, 152 (14), 144107,  DOI: 10.1063/5.0005081
    52. 52
      Werner, H.-J.; Knowles, P. J.; Knizia, G.; Manby, F. R.; Schütz, M.; Celani, P.; Györffy, W.; Kats, D.; Korona, T.; Lindh, R.; MOLPRO , version 2021.2, a package of ab initio programs, 2021. https://www.molpro.net.
    53. 53
      Ghigo, G.; Roos, B. O.; Malmqvist, P.-Å. A modified definition of the zeroth-order Hamiltonian in multiconfigurational perturbation theory (CASPT2). Chem. Phys. Lett. 2004, 396 (1–3), 142149,  DOI: 10.1016/j.cplett.2004.08.032
    54. 54
      Roos, B. O.; Andersson, K. Multiconfigurational perturbation theory with level shift ─ the Cr2 potential revisited. Chem. Phys. Lett. 1995, 245 (2–3), 215223,  DOI: 10.1016/0009-2614(95)01010-7
    55. 55
      Wen, X.; Martirez, J. M. P.; Carter, E. A. Plasmon-driven ammonia decomposition on Pd(111): Hole transfer’s role in changing rate-limiting steps. ACS Catal. 2024, 14, 95399553,  DOI: 10.1021/acscatal.4c01869
    56. 56
      Stradella, L. Heats of adsorption of different gases on polycrystalline transition metals. Adsorpt. Sci. Technol. 1992, 9 (3), 190198,  DOI: 10.1177/026361749200900304
    57. 57
      Peterson, K. A.; Figgen, D.; Dolg, M.; Stoll, H. Energy-consistent relativistic pseudopotentials and correlation consistent basis sets for the 4d elements Y–Pd. J. Chem. Phys. 2007, 126 (12), 124101,  DOI: 10.1063/1.2647019
    58. 58
      Dunning, T. H., Jr. Gaussian basis sets for use in correlated molecular calculations. I. The atoms boron through neon and hydrogen. J. Chem. Phys. 1989, 90 (2), 10071023,  DOI: 10.1063/1.456153
    59. 59
      Kendall, R. A.; Dunning, T. H., Jr.; Harrison, R. J. Electron affinities of the first-row atoms revisited. Systematic basis sets and wave functions. J. Chem. Phys. 1992, 96 (9), 67966806,  DOI: 10.1063/1.462569
    60. 60
      Peterson, K. A.; Dunning, T. H., Jr. Accurate correlation consistent basis sets for molecular core–valence correlation effects: The second row atoms Al–Ar, and the first row atoms B–Ne revisited. J. Chem. Phys. 2002, 117 (23), 1054810560,  DOI: 10.1063/1.1520138
    61. 61
      Pritchard, B. P.; Altarawy, D.; Didier, B.; Gibson, T. D.; Windus, T. L. New Basis Set Exchange: An Open, Up-to-Date Resource for the Molecular Sciences Community. J. Chem. Inf. Model. 2019, 59 (11), 48144820,  DOI: 10.1021/acs.jcim.9b00725
    62. 62
      Hill, J. G.; Peterson, K. A. Gaussian basis sets for use in correlated molecular calculations. XI. Pseudopotential-based and all-electron relativistic basis sets for alkali metal (K–Fr) and alkaline earth (Ca–Ra) elements. J. Chem. Phys. 2017, 147 (24), 244106,  DOI: 10.1063/1.5010587
    63. 63
      Balabanov, N. B.; Peterson, K. A. Systematically convergent basis sets for transition metals. I. All-electron correlation consistent basis sets for the 3d elements Sc–Zn. J. Chem. Phys. 2005, 123 (6), 064107,  DOI: 10.1063/1.1998907
    64. 64
      Jansen, G.; Hess, B. A. Revision of the Douglas-Kroll transformation. Phys. Rev. A 1989, 39 (11), 60166017,  DOI: 10.1103/PhysRevA.39.6016
    65. 65
      Douglas, M.; Kroll, N. M. Quantum electrodynamical corrections to the fine structure of helium. Ann. Phys. 1974, 82 (1), 89155,  DOI: 10.1016/0003-4916(74)90333-9
    66. 66
      Silva-Junior, M. R.; Schreiber, M.; Sauer, S. P. A.; Thiel, W. Benchmarks of electronically excited states: Basis set effects on CASPT2 results. J. Chem. Phys. 2010, 133 (17), 174318,  DOI: 10.1063/1.3499598
    67. 67
      Kánnár, D.; Tajti, A.; Szalay, P. G. Accuracy of Coupled Cluster Excitation Energies in Diffuse Basis Sets. J. Chem. Theory Comput. 2017, 13 (1), 202209,  DOI: 10.1021/acs.jctc.6b00875
    68. 68
      Jansen, H. B.; Ros, P. Non-empirical molecular orbital calculations on the protonation of carbon monoxide. Chem. Phys. Lett. 1969, 3 (3), 140143,  DOI: 10.1016/0009-2614(69)80118-1
    69. 69
      Liu, B.; McLean, A. D. Accurate calculation of the attractive interaction of two ground state helium atoms. J. Chem. Phys. 1973, 59 (8), 45574558,  DOI: 10.1063/1.1680654
    70. 70
      Gutowski, M.; Van Lenthe, J. H.; Verbeek, J.; Van Duijneveldt, F. B.; Chałasinski, G. The basis set superposition error in correlated electronic structure calculations. Chem. Phys. Lett. 1986, 124 (4), 370375,  DOI: 10.1016/0009-2614(86)85036-9
    71. 71
      Simon, S.; Duran, M.; Dannenberg, J. J. How does basis set superposition error change the potential surfaces for hydrogen-bonded dimers?. J. Chem. Phys. 1996, 105 (24), 1102411031,  DOI: 10.1063/1.472902
    72. 72
      Boys, S. F.; Bernardi, F. The calculation of small molecular interactions by the differences of separate total energies. Some procedures with reduced errors. Mol. Phys. 1970, 19 (4), 553566,  DOI: 10.1080/00268977000101561
    73. 73
      Gray, M.; Bowling, P. E.; Herbert, J. M. Systematic Evaluation of Counterpoise Correction in Density Functional Theory. J. Chem. Theory Comput. 2022, 18 (11), 67426756,  DOI: 10.1021/acs.jctc.2c00883
    74. 74
      Mentel, Ł. M.; Baerends, E. J. Can the Counterpoise Correction for Basis Set Superposition Effect Be Justified?. J. Chem. Theory Comput. 2014, 10 (1), 252267,  DOI: 10.1021/ct400990u
    75. 75
      van Duijneveldt, F. B.; van Duijneveldt-van de Rijdt, J. G. C. M.; van Lenthe, J. H. State of the Art in Counterpoise Theory. Chem. Rev. 1994, 94 (7), 18731885,  DOI: 10.1021/cr00031a007
    76. 76
      Burns, L. A.; Marshall, M. S.; Sherrill, C. D. Comparing Counterpoise-Corrected, Uncorrected, and Averaged Binding Energies for Benchmarking Noncovalent Interactions. J. Chem. Theory Comput. 2014, 10 (1), 4957,  DOI: 10.1021/ct400149j
    77. 77
      Helgaker, T.; Klopper, W.; Koch, H.; Noga, J. Basis-set convergence of correlated calculations on water. J. Chem. Phys. 1997, 106 (23), 96399646,  DOI: 10.1063/1.473863
    78. 78
      Wei, Z.; Martirez, J. M. P.; Carter, E. A. Introducing the embedded random phase approximation: H2 dissociative adsorption on Cu(111) as an exemplar. J. Chem. Phys. 2023, 159 (19), 194108,  DOI: 10.1063/5.0181229
    79. 79
      Zhou, L.; Swearer, D. F.; Zhang, C.; Robatjazi, H.; Zhao, H.; Henderson, L.; Dong, L.; Christopher, P.; Carter, E. A.; Nordlander, P. Quantifying hot carrier and thermal contributions in plasmonic photocatalysis. Science 2018, 362 (6410), 6972,  DOI: 10.1126/science.aat6967
    80. 80
      Finley, J.; Malmqvist, P.-Å.; Roos, B. O.; Serrano-Andrés, L. The multi-state CASPT2 method. Chem. Phys. Lett. 1998, 288 (2–4), 299306,  DOI: 10.1016/S0009-2614(98)00252-8
  • Supporting Information

    Supporting Information

    ARTICLE SECTIONS
    Jump To

    The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.jctc.4c00558.

    • Plots of all geometries, clean-surface approximation energetics, cluster-size effects, PW to GTO basis set effects, Pd 4s4p dynamic correlation effects, overview of BSSE and CP, plots and energetics of geometry effects on Vemb, NO plots of all structures surveyed, and SA-CASPT2 ground- and excited-state MEPs utilizing different state averaging (PDF)

    • Tables of SA-CASSCF dipole moments (XLSX)


    Terms & Conditions

    Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.