A production-level implementation of equation-of-motion coupled-cluster singles and doubles (EOM-CCSD) for electron attachment and excitation energies augmented by a complex absorbing potential (CAP) is presented. The new method enables the treatment of metastable states within the EOM-CC formalism in a similar manner as bound states. The numeric performance of the method and the sensitivity of resonance positions and lifetimes to the CAP parameters and the choice of one-electron basis set are investigated. A protocol for studying molecular shape resonances based on the use of standard basis sets and a universal criterion for choosing the CAP parameters are presented. Our results for a variety of π* shape resonances of small to medium-size molecules demonstrate that CAP-augmented EOM-CCSD is competitive relative to other theoretical approaches for the treatment of resonances and is often able to reproduce experimental results.
I. INTRODUCTION
Metastable electronic states are important in diverse areas of science and technology ranging from high-energy applications (plasmas, attosecond and X-ray spectroscopies) to electron-molecule collisions (interstellar chemistry, radiolysis, DNA damage by slow electrons). These states (called resonances) can be accessed when molecules are excited above their ionization threshold, via electron attachment to closed-shell species, or by core ionization.
A concise and pedagogical introduction to the topic as well as references to earlier reviews can be found in Ref. 1. From the quantum mechanical point of view, resonance states belong to the continuum part of the spectrum and, therefore, their wave functions are not L2-integrable. Yet, their wave functions bear certain resemblance to the bound states within the interaction region (i.e., close to the nuclei). Using boundary conditions for the outgoing wave (Siegert or Gamow formalism, see Ref. 1), one arrives at the following form of the resonance wave function:
where the phase-isolated part ϕR(x) resembles a bound-state wave function in the interaction region and ER and Γ (real and imaginary parts of the complex energy E = ER − iΓ/2) determine resonance position and width. The latter is inversely proportional to the resonance lifetime. Thus, the resonances appear as solutions of the Schrödinger equation with complex energy.1–4 One can arrive at the same concept of complex energy via a completely different formalism (Feshbach approach) based on a separation of the Hamiltonian into coupled bound and continuum parts; in this approach, the resonance is described as a bound state coupled to the continuum, and the complex energy emerges from solving a non-Hermitian eigenproblem with an effective Hamiltonian.5
One can avoid the inconveniences of working with continuum functions or fiddling with boundary conditions by reformulating the problem using complex variables.2–4 The most rigorous approach is the complex-scaling formalism2,6 in which all coordinates are scaled by a complex number e−iθ; however, practical applications of this method are limited by its extreme sensitivity to the one-electron basis set7–11 as well as conceptual difficulties regarding the separation of nuclear and electronic motions of the scaled Hamiltonian.12–15
These problems are avoided in an alternative approach in which the original (non-scaled) Hamiltonian is augmented by a complex potential
The use of CAPs in practical calculations is complicated by possible reflections leading to false resonances, sensitivity of the results to the form of the CAP W, and a strong basis set dependence.19,22 Furthermore, one has to determine an optimal value for η, which is usually achieved by calculating trajectories E(η) and requiring |η dE/dη| = min.
In terms more familiar to electronic structure practitioners, reflections can be described as perturbations of the resonance wave functions and, consequently, energies caused by a finite-strength CAP. Finite basis sets give rise to additional reflections. Several reflection-free CAPs have been proposed;19,23 such CAPs are either energy-dependent or non-local.
In our previous paper,24 we introduced a simple density-matrix based correction to the energy that removes the perturbation due to the CAP. The correction was derived based on energy decomposition analysis and response theory. Our starting point was the observation that the energy's dependence on η becomes roughly linear beyond some critical value of η. By analyzing the response equations, we also proposed an alternative criterion for finding an optimal value for η. Physically, our approach is grounded in the behavior of the resonance wave function and, ultimately, the one-particle density matrix. It was shown24 that when the CAP is sufficiently strong, both real and imaginary parts of the density become near-stationary indicating that the resonance is stabilized. Then the perturbation to the resonance position by the CAP can be eliminated by subtracting the term η Tr[γW] from the energy. The optimal η is found by considering the de-perturbed resonance energies; moreover, we argued that ηopt is not the same for real and imaginary parts.24 Preliminary benchmarks illustrated that this approach results in a computationally more robust scheme in which the dependence on the onset of the CAP is significantly reduced compared to the straightforward application of a CAP along with the original energy-based criterion for finding the optimal η, as was done in most CAP applications.25–27 We note that Moiseyev et al.28,29 also observed a linear energy dependence on η beyond some critical value of η and proposed the use of Padé approximants to extrapolate the energy of the stabilized resonance to the zero η limit. If the energy depends strictly linearly on η, their approach yields results identical to our first-order correction.
One of the difficulties of understanding the capabilities and limitations of different approaches is that a method that has shown excellent performance for a small model problem may fail when applied to a realistic system. In the context of electronic structure, the results of calculations of resonances will also be affected by the quality of standard approximations such as the incompleteness of one- and many-electron basis sets.30 Thus, it is important to test different methods for meta-stable states within robust and accurate ab initio approaches. For bound states, the coupled-cluster (CC) and equation-of-motion (EOM) hierarchies of methods31–35 provide a reliable and predictive set of theoretical model chemistries.36 These methods can be systematically improved to approach the exact solution, are size-extensive (size-intensive for excitation energies), describe dynamical and non-dynamical correlation in one computational step, and do not involve system-dependent parameterization. The CC hierarchy of methods works best for wave functions dominated by a single Slater determinant, however, the EOM-CC approach extends this single-reference formalism to tackle various open-shell and multi-configurational cases.33,37
In EOM-CC, the target-state wave function is described by an excitation operator
where
Recently, we presented an implementation of complex-scaled EOM-CC singles and doubles (EOM-CCSD) methods and illustrated their performance by considering several atomic systems (He, H−, Be). Here, we present an implementation of CAPs within the EOM-CC family of methods. Our main focus is on the EOM-EA-CC variant; however, our implementation also includes EOM-EE-CC. While limited implementations of CAPs within EOM-CC have been reported before (e.g., Refs. 25 and 26), this work presents the first formally complete and production-level implementation of the method.
The main focus of the paper is on investigating the numeric performance of the method and the sensitivity of the results towards the CAP parameters and the choice of basis set. Our goal is to develop a black-box approach that could be calibrated and then applied to the calculation of resonances without any prior knowledge of the system, as advocated by John Pople.36 In particular, we want to avoid the system-dependent optimization of basis sets and the CAP's shape and onset. Thus, rather than aiming at results converged with respect to all computational parameters individually for each system, we wish to establish a uniform protocol that can be applied to any system and can be characterized by error bars estimated from prior calibration studies, as routinely performed in electronic structure calculations.30
We note that the validation of the accuracy of computed resonance lifetimes in molecular systems is difficult.38 A complete theoretical description should involve coupled electronic and nuclear dynamics; this is beyond the scope of the present paper, where we only compute the lifetime of the resonance state at a fixed molecular geometry. This is appropriate for resonances whose lifetimes are short relative to nuclear motions, or when nuclear motions do not strongly affect the computed Γ values (Condon-like approximation). Thus, our focus is on the comparison with other theoretical studies and the robustness of the results with respect to the one-electron basis set as well as variations of the CAP parameters.
In this context, we add that the sensitivity of the results to the one-electron basis set is of a fundamentally different origin in CAP calculations as compared to complex scaling. In the latter case, the basis should be sufficiently flexible to describe the resonance wave function at different values of the scaling angle, whereas in the former case, one simply needs to supply a basis set of a sufficient spatial extent to represent a given CAP and a stabilized resonance wave function. This implies that the diffuseness of the basis must be coordinated with the CAP onset, e.g., in a compact basis, the CAP onset should be smaller, otherwise, the calculation will be blind to the CAP. Thus, although the basis-set dependence is a nuisance, its simpler nature in CAP calculations suggests that a solution can be found.
Originally, CAP methods were introduced to study shape resonances. Since the decay of Feshbach resonances is a two-electron process (and the CAP is a one-electron operator), one may expect difficulties in describing Feshbach resonances within the CAP formalism. Moiseyev et al.39–41 showed that additional steps need to be taken for the construction of reflection-free CAPs in order to reliably calculate Feshbach-type resonances. The present paper focuses solely on understanding CAPs in the context of molecular shape resonances.
The article is structured as follows: Secs. II and III present the formalism of the CAP-augmented EOM-CC method and our implementation. In Sec. IV, we put forward a protocol for determining the resonance positions and lifetimes and investigate its robustness towards the choice of the one-electron basis set and the CAP's onset. In Sec. V, we subsequently apply our new scheme to a variety of molecular resonance states and compare the results to those obtained from experiment as well as using other theoretical approaches. Sec. VI provides concluding remarks.
II. THEORY
The basic idea of the CAP method16–19 is the addition of an artificial complex potential to the original Hamiltonian
where W aims to absorb the outgoing electron and η controls its strength. As in complex scaling,2–4,11,42 the addition of the CAP results in a non-Hermitian complex symmetric operator H(η)16 converting resonances into square integrable (L2) wave functions. In our calculations, we choose the CAP as a quadratic potential with an unaffected region of cuboid (i.e., box) shape
with rα denoting the three Cartesian coordinates (α = x, y, z). Thus, the CAP is controlled by 4 parameters: 3 parameters for the onset in each direction (
In the complete one-electron basis set, the exact position of the resonance in the complex plane can be obtained as limη → 0E(η).16 That is, an infinitesimally weak CAP, which is represented exactly (and, therefore, goes to infinity at large r), is sufficient to stabilize the resonance without perturbing it. Working with finite Gaussian basis sets requires one to perform series of calculations for different η in order to find an optimal value of the strength parameter ηopt along the η-trajectory and the corresponding value of the resonance energy E(ηopt). A commonly used criterion for determining the optimal value of the strength parameter η is finding the minimum of the logarithmic velocity16,19
Unfortunately, the position and width of the resonance computed using this criterion are very sensitive to the CAP onset17,24,27; thus, this approach does not provide a black-box scheme. In our recent paper,24 the first-order deperturbative correction to the raw zeroth-order resonance energies ER and EI was introduced as
with γ(η) as the one-particle density matrix. The correction is based on perturbation theory; it removes the explicit dependence on the CAP from the computed resonance energies. As was illustrated in Ref. 24, the corrected energies, UR(η) and UI(η), exhibit nearly constant behavior for large η (that is, past the stabilization point, when the resonance wave function does not change much anymore); furthermore, the corrected trajectories computed using different CAP onsets become much more similar in the asymptotic region as opposed to the uncorrected trajectories, ER(η) and EI(η).
By looking separately at the real and imaginary parts of the deperturbed energy (UR and UI) as a function of η, we showed that the energy becomes near stationary at certain values of optimal strength (
In our scheme, the CAP is introduced at the HF level of theory, where we obtain a set of complex molecular orbitals (MOs) as the solution for a given strength η. Hence, a complex Koopmans' theorem holds for the virtual orbitals, i.e., they can be interpreted as zeroth-order approximations of the resonance and rotated continuum states.
As the next step, we solve the CCSD equations for the reference state31,32,43–45 using H(η)
with Φμ denoting the excited determinants. The resulting amplitudes tη are also complex.
To compute electronically excited and electron-attached resonance states, we use EOM-EE-CCSD and EOM-EA-CCSD33,46–50 methods that provide accurate and predictive descriptions for such target states. The wave function of the resonance state is found by solving a non-Hermitian eigenvalue problem for the right eigenvectors
which yields a set of complex amplitudes Rη and complex excitation energies Ωη. The latter are the raw, η-dependent resonance energies (which are equal to the difference between the total energy of the excited/attached EOM-CCSD state and the reference CCSD energy for a given η).
We note that for moderate CAP strengths and CAP onsets comparable with the spatial extent of the electron density of the reference state, the perturbation to the reference CCSD energy is small (10−5 a.u.) in contrast to complex-scaled calculations.11 To compute the first-order correction to the raw resonance energies, the one-electron density matrix needs to be calculated. We employ an unrelaxed one-electron EOM-CCSD density matrix containing no amplitude- or orbital-response terms51
where Lη and Rη are the left and right EOM-CCSD eigenvectors, respectively.
Because of the non-Hermitian nature of
where i and j denote different electronic states. Once the density matrix is computed, the energy of the resonance state is corrected according to Eqs. (7) and (8).
The equations for CAP-CCSD and CAP-EOM-EE/EA-CCSD are identical to the original CCSD and EOM-EE/EA-CCSD equations except that all quantities such as the Fock matrix, the two-electron integrals, the MO matrix C, the T and R/L amplitudes are now complex and η-dependent. There is no need to add the CAP explicitly to the CCSD or EOM-CCSD equations since it is already included at the HF level.
In our paper on complex-scaled EOM-CC,11 we considered several variants of implementation, including one in which the HF and CCSD equations for the reference state were solved for the unscaled Hamiltonian and the scaling was introduced only at the EOM-CC level. This required significant reformulation of the EOM-CC equations. By analogy, one may also consider an implementation of CAP-EOM-CC in which the CAP is introduced only at the EOM-CC level; this will be the subject of future work.
Due to the CAP, the Hamiltonian becomes non-Hermitian and complex symmetric, which necessitates using a different metric, the so-called complex symmetric scalar product (c-product),16,42,52,53 such that the variational principle is maintained
The difference to the regular scalar product is that the bra-vector is not complex conjugated. Mathematically, the c-product is a pseudoscalar product which does not induce a valid metric norm.42,52 However, one can still define the c-norm (f|f) which is, contrary to the regular norm, complex in general and might become zero for a non-zero function f (“self-orthogonality”)
where ⟨|⟩ is a regular scalar product which is equivalent to the c-product for real functions. Thus, the normalization of all vectors (for example, left and right EOM-CC eigenvectors) is done by multiplying by a complex number42
resulting in
III. IMPLEMENTATION
The suite of CAP-EOM-CC methods has been implemented in the Q-Chem electronic structure package54,55 and was released in version 4.2. All complex CCSD and EOM-EE/EA-CCSD equations were implemented using the libtensor library56 for high-performance tensor operations.
The calculations begin by solving the CAP-augmented restricted HF equations (CAP-RHF). CAP-RHF has been implemented as an extension of regular RHF using the object-oriented SCF library SCFman57 in Q-Chem that employs the Armadillo linear algebra library58 for matrix computations. We add that an adaptation of our implementation for CAP-UHF will be straightforward. The CAP is introduced as an additional term in the regular Fock matrix
The molecular orbitals must satisfy the following orthonormalization condition in the c-product metric:
where S is the overlap matrix in the atomic orbital (AO) basis and (Cη)T is transposed but not conjugated. Since the augmented Fock matrix Fη is non-Hermitian, the orbitals obtained using standard linear algebra routines for the diagonalization of general matrices are not normalized. In order to satisfy the orthonormality condition [Eq. (18)], the MOs are orthogonalized by using a modified Gram-Schmidt procedure with projections calculated using the c-product.
The Fock matrix
The CAP is evaluated in the AO basis through numerical quadrature using a Becke-type grid61 of (99, 590) points (99 radial points and 590 angular points per radial point). Currently, we have implemented a shifted quadratic potential (r − r0)2 for a rectangular cuboid [Eqs. (4) and (5)], but our implementation allows for an easy extension of the shape of the unaffected region as well as the type of potential, e.g., higher order monomials (r − r0)4, (r − r0)6, etc. Our implementation also includes the optional addition of a real potential to the CAP, as was advocated in Ref. 41.
Relative to the conventional CCSD and EOM-CCSD methods, the addition of the CAP does not change the scaling of the computational cost [O(N6) for CCSD and EOM-EE-CCSD, O(N5) for EOM-EA-CCSD] or the memory requirements [O(N4)]. However, because we need to work with complex numbers, the computational cost increases roughly by a factor of 4 and the storage requirements increase by a factor of 2. Furthermore, computation of the first-order energy correction requires the left eigenvectors, which increases the computation time relative to EOM-CCSD energy calculations. Another possible issue arises when the resonance state is lying high in energy, so that the standard Davidson procedure has to find all lower roots that might require a lot of iterations to converge. To solve this issue, we have implemented iterative solvers for interior eigenstates for conventional EOM-CCSD methods. Implementation of the interior eigenvalue solvers for CAP-EOM-CCSD methods is a subject of future work. Finally, the necessity to compute η-trajectories requires to run calculations for different values of η, but these calculations can be performed independently and can therefore be run in parallel.
IV. BENCHMARK CALCULATIONS
The necessity to find optimal values for the strength and the onset of the CAP as well as a pronounced basis set dependence of the results have precluded routine applications of CAP-based methods so far. Hence, investigating the numeric performance of CAP-EOM-CCSD, in particular with respect to the two aforementioned issues, is crucial for it to become a useful tool for studying resonances.
As for the one-electron basis set, it has been established that the straightforward application of standard basis sets yields poor results. Additional diffuse functions need to be incorporated for two reasons: (i) to obtain a sufficiently good basis-set representation of the CAP and (ii) to describe the outgoing electron correctly.16,19,62 Owing to these requirements, CAP-based computations were often carried out using non-standard basis sets.26,27,62–67 While such approaches are able to provide results that agree with experiment for some resonance states, a treatment based on standard basis sets not involving any optimization procedure would be superior as it is of black-box type, has predictive power, and is also computationally less demanding. Since the shape of the resonance wave function is similar to a bound-state wave function in the interaction region, it should be possible to lessen the basis-set dependence to the degree observed in regular EOM-CCSD calculations.
Concerning the choice of the CAP onset, one has to realize that the artificial nature of the CAP implies that it is impossible to deduce from basic physical laws a universally applicable procedure for finding optimal parameters for the CAP strength, onset, and shape. While this is unsatisfying from a formal point of view, a pragmatic approach is to mitigate the dependence of the physically meaningful results on the artificial parameters as much as possible. Along these lines, we introduced a first-order correction24 that was shown to desensitize resonance positions and widths to the choice of onset parameters by removing the perturbation due to the CAP. However, as this dependence cannot be removed completely, one has to develop a protocol for the unique and system-independent choice of the CAP onset to make CAP-EOM-CCSD applicable in a routine manner.
A. Computational details
In the following, we examine the π* shape resonances of CO− and
The basis sets used in our calculations were derived from the aug-cc-pVXZ (X = D, T, Q, 5) series68 through augmentation by additional even-tempered basis functions. In all cases, the exponents for the first additional basis functions were obtained as one half of the exponent of the most diffuse basis function with the same angular momentum in the parent aug-cc-pVXZ basis set. The exponents for the remaining additional basis functions were calculated as one half of the exponent of the preceding function.
We explored two different series of basis sets, namely, one where we augmented the basis sets for all atoms except hydrogen (denoted as (A) below) and one where we placed only one set of diffuse functions with averaged exponents in the center of the molecule (denoted as (C) below). Since the second approach is computationally less demanding, it is especially preferable when targeting larger systems. To ensure that this strategy does not give rise to artifacts, we computed EOM-EE-CCSD excitation energies for a number of bound excited states of CO and C2H4 using the aug-cc-pVTZ+3s3p3d(C) and aug-cc-pVTZ+3s3p3d(A) bases. The results are reported in Table I together with the corresponding values for ⟨R2⟩, which are helpful in distinguishing valence states from Rydberg states. As apparent from Table I excitation energies computed with the two basis sets differ by not more than 0.04 eV for the very diffuse 4 1A1 state of CO and by just 0.001 eV for all other states. This shows that placing the additional diffuse functions in the center does not lead to inferior results for the bound excited states and thus suggests the application of this scheme to resonances. We note that a similar scheme has been employed before in the context of stabilization techniques.69,70
EOM-EE-CCSD excitation energies and expectation values ⟨R2⟩ for several excited states of CO and C2H4 computed using the aug-cc-pVTZ basis set with additional diffuse basis functions placed at all heavy atoms (A) or at the center of the molecule (C).
. | . | aug-cc-pVTZ . | aug-cc-pVTZ . | |
---|---|---|---|---|
. | . | +3s3p3d(A) . | +3s3p3d(C) . | |
Molecule . | State . | ⟨R2⟩ (a.u.)a . | E (eV) . | E (eV) . |
CO | 2 1A1 | 77.8 | 10.961 | 10.961 |
4 1A1 | 153.7 | 12.559 | 12.597 | |
1 1B2 | 41.4 | 8.625 | 8.626 | |
C2H4 | 2 1Ag | 152.0 | 8.445 | 8.446 |
1 1B1g | 116.7 | 9.791 | 9.791 | |
1 1B1u | 118.7 | 7.392 | 7.392 |
. | . | aug-cc-pVTZ . | aug-cc-pVTZ . | |
---|---|---|---|---|
. | . | +3s3p3d(A) . | +3s3p3d(C) . | |
Molecule . | State . | ⟨R2⟩ (a.u.)a . | E (eV) . | E (eV) . |
CO | 2 1A1 | 77.8 | 10.961 | 10.961 |
4 1A1 | 153.7 | 12.559 | 12.597 | |
1 1B2 | 41.4 | 8.625 | 8.626 | |
C2H4 | 2 1Ag | 152.0 | 8.445 | 8.446 |
1 1B1g | 116.7 | 9.791 | 9.791 | |
1 1B1u | 118.7 | 7.392 | 7.392 |
The corresponding values for the ground states are 40.0 a.u. for CO and 83.3 a.u. for C2H4.
As for the choice of the CAP onset, we employed the square roots of the expectation values ⟨α2⟩ (α = x, y, z) for the ground states calculated at the CCSD level of theory as a starting point and considered the impact of small variations. The values used are
B. The impact of the CAP onset
Table II compiles resonance positions ER and widths Γ for the 2Π resonance of CO− and the 2B2g resonance of
Dependence of resonance positions ER and widths Γ of the 2Π resonance of CO− and the 2B2g resonance of
Relative CAP onseta . | Zeroth-order values . | First-order values . | . | ||||||
---|---|---|---|---|---|---|---|---|---|
|$\Delta r^0_x/\Delta r^0_y/\Delta r^0_z$| (a.u.) . | ER (eV) . | Γ (eV) . | ηopt (a.u.) . | |$\eta \frac{dE}{d\eta }$| (a.u.) . | ER (eV)b . | Γ (eV)b . | |$\eta ^R_{\rm opt}$| (a.u.) . | |$\eta ^I_{\rm opt}$| (a.u.) . | ‖W‖ (a.u.)c . |
CO−, 2Π resonance | |||||||||
0.0/0.0/0.0 | 2.088 | 0.650 | 0.0028 | 0.0017 | 1.981 | 0.585 | 0.0054 | 0.0048 | 241.9 |
0.5/0.5/0.0 | 2.061 | 0.612 | 0.0036 | 0.0021 | 1.956 | 0.573 | 0.0066 | 0.0060 | 214.0 |
−0.5/−0.5/0.0 | 2.113 | 0.691 | 0.0022 | 0.0011 | 1.999 | 0.591 | 0.0044 | 0.0040 | 274.2 |
0.0/0.0/0.5 | 2.087 | 0.644 | 0.0030 | 0.0017 | 1.981 | 0.582 | 0.0056 | 0.0050 | 235.4 |
0.0/0.0/−0.5 | 2.091 | 0.654 | 0.0028 | 0.0015 | 1.980 | 0.591 | 0.0052 | 0.0046 | 249.6 |
|${\rm C}_2{\rm H}_4^-$|, 2B2g resonance | |||||||||
0.0/0.0/0.0 | 2.091 | 0.430 | 0.0046 | 0.0023 | 2.032 | 0.328 | 0.0060 | 0.0085 | 272.0 |
0.5/0.0/0.0 | 2.089 | 0.434 | 0.0045 | 0.0023 | 2.032 | 0.330 | 0.0054 | 0.0089 | 266.7 |
−0.5/0.0/0.0 | 2.093 | 0.427 | 0.0046 | 0.0019 | 2.031 | 0.328 | 0.0054 | 0.0083 | 278.3 |
0.0/0.5/0.0 | 2.095 | 0.429 | 0.0050 | 0.0022 | 2.033 | 0.330 | 0.0057 | 0.0087 | 260.5 |
0.0/−0.5/0.0 | 2.093 | 0.431 | 0.0046 | 0.0019 | 2.032 | 0.326 | 0.0054 | 0.0083 | 285.4 |
0.0/0.0/0.5 | 2.088 | 0.388 | 0.0071 | 0.0020 | 2.023 | 0.301 | 0.0070 | 0.0109 | 255.3 |
0.0/0.0/−0.5 | 2.106 | 0.478 | 0.0037 | 0.0025 | 2.039 | 0.353 | 0.0042 | 0.0068 | 291.2 |
Relative CAP onseta . | Zeroth-order values . | First-order values . | . | ||||||
---|---|---|---|---|---|---|---|---|---|
|$\Delta r^0_x/\Delta r^0_y/\Delta r^0_z$| (a.u.) . | ER (eV) . | Γ (eV) . | ηopt (a.u.) . | |$\eta \frac{dE}{d\eta }$| (a.u.) . | ER (eV)b . | Γ (eV)b . | |$\eta ^R_{\rm opt}$| (a.u.) . | |$\eta ^I_{\rm opt}$| (a.u.) . | ‖W‖ (a.u.)c . |
CO−, 2Π resonance | |||||||||
0.0/0.0/0.0 | 2.088 | 0.650 | 0.0028 | 0.0017 | 1.981 | 0.585 | 0.0054 | 0.0048 | 241.9 |
0.5/0.5/0.0 | 2.061 | 0.612 | 0.0036 | 0.0021 | 1.956 | 0.573 | 0.0066 | 0.0060 | 214.0 |
−0.5/−0.5/0.0 | 2.113 | 0.691 | 0.0022 | 0.0011 | 1.999 | 0.591 | 0.0044 | 0.0040 | 274.2 |
0.0/0.0/0.5 | 2.087 | 0.644 | 0.0030 | 0.0017 | 1.981 | 0.582 | 0.0056 | 0.0050 | 235.4 |
0.0/0.0/−0.5 | 2.091 | 0.654 | 0.0028 | 0.0015 | 1.980 | 0.591 | 0.0052 | 0.0046 | 249.6 |
|${\rm C}_2{\rm H}_4^-$|, 2B2g resonance | |||||||||
0.0/0.0/0.0 | 2.091 | 0.430 | 0.0046 | 0.0023 | 2.032 | 0.328 | 0.0060 | 0.0085 | 272.0 |
0.5/0.0/0.0 | 2.089 | 0.434 | 0.0045 | 0.0023 | 2.032 | 0.330 | 0.0054 | 0.0089 | 266.7 |
−0.5/0.0/0.0 | 2.093 | 0.427 | 0.0046 | 0.0019 | 2.031 | 0.328 | 0.0054 | 0.0083 | 278.3 |
0.0/0.5/0.0 | 2.095 | 0.429 | 0.0050 | 0.0022 | 2.033 | 0.330 | 0.0057 | 0.0087 | 260.5 |
0.0/−0.5/0.0 | 2.093 | 0.431 | 0.0046 | 0.0019 | 2.032 | 0.326 | 0.0054 | 0.0083 | 285.4 |
0.0/0.0/0.5 | 2.088 | 0.388 | 0.0071 | 0.0020 | 2.023 | 0.301 | 0.0070 | 0.0109 | 255.3 |
0.0/0.0/−0.5 | 2.106 | 0.478 | 0.0037 | 0.0025 | 2.039 | 0.353 | 0.0042 | 0.0068 | 291.2 |
A value of 0.0 refers to |$r^0_\alpha =\sqrt{\langle \alpha ^2 \rangle }$|, α = x, y, z. 0.5/0.5/0.0 means, for example, that |$r^0_x=\sqrt{\langle x^2 \rangle } + 0.5$|, |$r^0_y=\sqrt{\langle y^2 \rangle } +0.5$|, |$r^0_z=\sqrt{ \langle z^2 \rangle }$|.
Identical to UR and −1/2 UI from Eqs. (7) and (8).
Computed as Frobenius norm.
Table II shows that both zeroth-order and first-order resonance positions and widths become smaller when increasing
C. The role of diffuse basis functions
To analyze the convergence of resonance positions and widths with respect to the addition of diffuse basis functions, we studied the π* resonances of CO− and
Resonance positions ER and widths Γ as well as values for ηopt and ‖W‖ for the 2Π resonance of CO− and the 2B2g resonance of
. | Zeroth-order values . | First-order values . | . | |||||
---|---|---|---|---|---|---|---|---|
Basis set . | ER (eV) . | Γ (eV) . | ηopt (a.u.) . | ER (eV)a . | Γ (eV)a . | |$\eta ^R_{\rm opt}$| (a.u.) . | |$\eta ^I_{\rm opt}$| (a.u.) . | ‖W‖ (a.u.)b . |
CO−, 2Π resonance | ||||||||
aug-cc-pVTZ | 2.517 ± 0.013 | 0.482 ± 0.120 | 0.0180 | …c | 0.252 ± 0.020 | …c | 0.0660 | 12.1 |
aug-cc-pVTZ + 3s(C) | 2.517 ± 0.013 | 0.469 ± 0.120 | 0.0190 | …c | 0.215 ± 0.021 | …c | 0.0585 | 49.3 |
aug-cc-pVTZ + 6s(C) | 2.518 ± 0.014 | 0.468 ± 0.012 | 0.0190 | …c | …c | …c | …c | 412.0 |
aug-cc-pVTZ + 3s3p(C) | 2.025 ± 0.015 | 0.549 ± 0.024 | 0.0024 | 1.825 ± 0.004 | 0.357 ± 0.009 | 0.0082 | 0.0048 | 229.5 |
aug-cc-pVTZ + 6s6p(C) | 2.023 ± 0.017 | 0.551 ± 0.023 | 0.0024 | 1.830 ± 0.004 | 0.364 ± 0.008 | 0.0082 | 0.0048 | 1078.3 |
aug-cc-pVTZ + 3s3p3d(C) | 2.088 ± 0.026 | 0.650 ± 0.040 | 0.0028 | 1.981 ± 0.022 | 0.585 ± 0.012 | 0.0054 | 0.0048 | 241.9 |
aug-cc-pVTZ + 6s6p6d(C) | 2.112 ± 0.027 | 0.665 ± 0.055 | 0.0034 | 2.015 ± 0.025 | 0.555 ± 0.057 | 0.0048 | 0.0340 | 1441.8 |
aug-cc-pVTZ + 3s3p3d3f(C) | 2.081 ± 0.028 | 0.654 ± 0.043 | 0.0030 | 1.969 ± 0.022 | 0.588 ± 0.015 | 0.0056 | 0.0050 | 243.7 |
aug-cc-pVTZ + 3s3p(A) | 2.102 ± 0.024 | 0.604 ± 0.050 | 0.0034 | 1.961 ± 0.016 | 0.446 ± 0.0026 | 0.0098 | 0.0066 | 505.9 |
aug-cc-pVTZ + 3s3p3d(A) | 2.060 ± 0.022 | 0.885 ± 0.022 | 0.0016 | 1.969 ± 0.025 | 0.474 ± 0.032 | 0.0026 | 0.0305 | 541.0 |
|${\rm C}_2{\rm H}_4^-$|, 2B2g resonance | ||||||||
aug-cc-pVTZ | 2.433d | 0.479d | 0.0380 | 2.411d | 0.140d | 0.0880 | 0.1610 | 11.8 |
aug-cc-pVTZ + 3s(C) | 2.433d | 0.479d | 0.0383 | 2.411d | …c | 0.0878 | …c | 51.9 |
aug-cc-pVTZ + 3s3p(C) | 2.422 ± 0.024 | 0.495 ± 0.012 | 0.0370 | 2.410 ± 0.007 | 0.208 ± 0.030 | 0.1200 | 0.1700 | 253.5 |
aug-cc-pVTZ + 6s6p(C) | 2.419d | 0.494d | 0.0360 | 2.409d | 0.222d | 0.1250 | 0.1690 | 1429.0 |
aug-cc-pVTZ + 3s3p3d(C) | 2.091 ± 0.015 | 0.430 ± 0.046 | 0.0046 | 2.032 ± 0.009 | 0.328 ± 0.027 | 0.0060 | 0.0085 | 272.0 |
aug-cc-pVTZ + 6s6p6d(C) | 1.976d | 0.563d | 0.0013 | 2.043d | 0.334d | 0.0076 | 0.0100 | 2014.7 |
aug-cc-pVTZ + 3s3p3d3f(C) | 2.088d | 0.441d | 0.0046 | 2.028d | 0.334d | 0.0051 | 0.0085 | 274.0 |
aug-cc-pVTZ + 3s3p(A) | 2.108 ± 0.018 | 0.420 ± 0.045 | 0.0049 | 2.017 ± 0.007 | 0.305 ± 0.030 | 0.0110 | 0.0085 | 502.8 |
aug-cc-pVTZ + 3s3p3d(A) | 2.302 ± 0.043 | 0.536 ± 0.096 | 0.0245 | 2.180 ± 0.008 | 0.373 ± 0.050 | 0.0370 | 0.0290 | 539.2 |
. | Zeroth-order values . | First-order values . | . | |||||
---|---|---|---|---|---|---|---|---|
Basis set . | ER (eV) . | Γ (eV) . | ηopt (a.u.) . | ER (eV)a . | Γ (eV)a . | |$\eta ^R_{\rm opt}$| (a.u.) . | |$\eta ^I_{\rm opt}$| (a.u.) . | ‖W‖ (a.u.)b . |
CO−, 2Π resonance | ||||||||
aug-cc-pVTZ | 2.517 ± 0.013 | 0.482 ± 0.120 | 0.0180 | …c | 0.252 ± 0.020 | …c | 0.0660 | 12.1 |
aug-cc-pVTZ + 3s(C) | 2.517 ± 0.013 | 0.469 ± 0.120 | 0.0190 | …c | 0.215 ± 0.021 | …c | 0.0585 | 49.3 |
aug-cc-pVTZ + 6s(C) | 2.518 ± 0.014 | 0.468 ± 0.012 | 0.0190 | …c | …c | …c | …c | 412.0 |
aug-cc-pVTZ + 3s3p(C) | 2.025 ± 0.015 | 0.549 ± 0.024 | 0.0024 | 1.825 ± 0.004 | 0.357 ± 0.009 | 0.0082 | 0.0048 | 229.5 |
aug-cc-pVTZ + 6s6p(C) | 2.023 ± 0.017 | 0.551 ± 0.023 | 0.0024 | 1.830 ± 0.004 | 0.364 ± 0.008 | 0.0082 | 0.0048 | 1078.3 |
aug-cc-pVTZ + 3s3p3d(C) | 2.088 ± 0.026 | 0.650 ± 0.040 | 0.0028 | 1.981 ± 0.022 | 0.585 ± 0.012 | 0.0054 | 0.0048 | 241.9 |
aug-cc-pVTZ + 6s6p6d(C) | 2.112 ± 0.027 | 0.665 ± 0.055 | 0.0034 | 2.015 ± 0.025 | 0.555 ± 0.057 | 0.0048 | 0.0340 | 1441.8 |
aug-cc-pVTZ + 3s3p3d3f(C) | 2.081 ± 0.028 | 0.654 ± 0.043 | 0.0030 | 1.969 ± 0.022 | 0.588 ± 0.015 | 0.0056 | 0.0050 | 243.7 |
aug-cc-pVTZ + 3s3p(A) | 2.102 ± 0.024 | 0.604 ± 0.050 | 0.0034 | 1.961 ± 0.016 | 0.446 ± 0.0026 | 0.0098 | 0.0066 | 505.9 |
aug-cc-pVTZ + 3s3p3d(A) | 2.060 ± 0.022 | 0.885 ± 0.022 | 0.0016 | 1.969 ± 0.025 | 0.474 ± 0.032 | 0.0026 | 0.0305 | 541.0 |
|${\rm C}_2{\rm H}_4^-$|, 2B2g resonance | ||||||||
aug-cc-pVTZ | 2.433d | 0.479d | 0.0380 | 2.411d | 0.140d | 0.0880 | 0.1610 | 11.8 |
aug-cc-pVTZ + 3s(C) | 2.433d | 0.479d | 0.0383 | 2.411d | …c | 0.0878 | …c | 51.9 |
aug-cc-pVTZ + 3s3p(C) | 2.422 ± 0.024 | 0.495 ± 0.012 | 0.0370 | 2.410 ± 0.007 | 0.208 ± 0.030 | 0.1200 | 0.1700 | 253.5 |
aug-cc-pVTZ + 6s6p(C) | 2.419d | 0.494d | 0.0360 | 2.409d | 0.222d | 0.1250 | 0.1690 | 1429.0 |
aug-cc-pVTZ + 3s3p3d(C) | 2.091 ± 0.015 | 0.430 ± 0.046 | 0.0046 | 2.032 ± 0.009 | 0.328 ± 0.027 | 0.0060 | 0.0085 | 272.0 |
aug-cc-pVTZ + 6s6p6d(C) | 1.976d | 0.563d | 0.0013 | 2.043d | 0.334d | 0.0076 | 0.0100 | 2014.7 |
aug-cc-pVTZ + 3s3p3d3f(C) | 2.088d | 0.441d | 0.0046 | 2.028d | 0.334d | 0.0051 | 0.0085 | 274.0 |
aug-cc-pVTZ + 3s3p(A) | 2.108 ± 0.018 | 0.420 ± 0.045 | 0.0049 | 2.017 ± 0.007 | 0.305 ± 0.030 | 0.0110 | 0.0085 | 502.8 |
aug-cc-pVTZ + 3s3p3d(A) | 2.302 ± 0.043 | 0.536 ± 0.096 | 0.0245 | 2.180 ± 0.008 | 0.373 ± 0.050 | 0.0370 | 0.0290 | 539.2 |
Identical to UR and −1/2 UI from Eqs. (7) and (8).
Computed as Frobenius norm.
No stationary point could be located in the range 0 <η < 0.2.
We did not vary the CAP onset in calculations using these basis sets.
For
Real (right) and imaginary (left) parts of E and U as a function of the CAP strength parameter η for the 2Πg resonance of C2H4. All values computed by CAP-EOM-EA-CCSD/aug-cc-pVTZ with different additional diffuse functions. • refers to zeroth-order values, × to first-order values.
Real (right) and imaginary (left) parts of E and U as a function of the CAP strength parameter η for the 2Πg resonance of C2H4. All values computed by CAP-EOM-EA-CCSD/aug-cc-pVTZ with different additional diffuse functions. • refers to zeroth-order values, × to first-order values.
We also investigated the effect of adding more than three additional diffuse s, p, and d-functions. The corresponding results in Table III show that, while values for ‖W‖ become considerably larger indicating a more complete basis-set representation of W, the impact on resonance positions and widths does not exceed 0.035 eV except for one case: For
Table III also reports the results from calculations with the aug-cc-pVTZ+3s3p(A) and aug-cc-pVTZ+3s3p3d(A) bases. Contrary to what we observed for bound states (cf. Table I), the differences between values for ER and Γ obtained with the two augmentation schemes “C” and “A” are not negligible. Zeroth-order values differ by up to 0.23 eV and first-order values still by up to 0.15 eV. In one case, namely,
Real (right) and imaginary (left) parts of E and U as a function of the CAP strength parameter η for the 2Πg resonance of C2H4. All values computed by CAP-EOM-EA-CCSD/aug-cc-pVTZ+3s3p3d(C) and aug-cc-pVTZ+3s3p3d(A), respectively. • refers to zeroth-order values, × to first-order values.
Real (right) and imaginary (left) parts of E and U as a function of the CAP strength parameter η for the 2Πg resonance of C2H4. All values computed by CAP-EOM-EA-CCSD/aug-cc-pVTZ+3s3p3d(C) and aug-cc-pVTZ+3s3p3d(A), respectively. • refers to zeroth-order values, × to first-order values.
D. The role of the valence basis set
Besides the impact of additional diffuse functions, variations in the valence basis set also influence ER and Γ. This is illustrated by Table IV, which reports values for resonance positions and widths of the 2Π resonance of CO− and the 2B2g resonance of
Resonance positions ER and widths Γ as well as values for ηopt for the 2Π resonance of CO− and the 2B2g resonance of
. | aug-cc-pVDZ . | aug-cc-pVTZ . | aug-cc-pVQZ . | aug-cc-pV5Z . |
---|---|---|---|---|
. | +3s3p3d(C) . | +3s3p3d(C) . | +3s3p3d(C) . | +3s3p3d(C) . |
2Π resonance of CO− | ||||
ER (0th order) (eV) | 2.303 | 2.088 | 1.987 | 1.926 |
Γ (0th order) (eV) | 0.727 | 0.650 | 0.696 | 0.804 |
ηopt (a.u.) | 0.0046 | 0.0028 | 0.0020 | 0.0015 |
ER (1st order) (eV)a | 2.182 | 1.981 | 1.851 | 1.762 |
Γ (1st order) (eV)a | 0.667 | 0.585 | 0.673 | 0.604 |
$\eta ^R_{\rm opt}$ (a.u.) | 0.0175 | 0.0054 | 0.0062 | 0.0034 |
$\eta ^I_{\rm opt}$ (a.u.) | 0.0100 | 0.0048 | 0.0046 | 0.0028 |
Bound states of CO | ||||
E (2 1A1) (eV) | 10.777 | 10.961 | 11.021 | 11.046 |
E (4 1A1) (eV) | 12.446 | 12.597 | 12.642 | 12.663 |
E (1 1B2) (eV) | 8.703 | 8.626 | 8.612 | 8.608 |
2B2g resonance of ${\rm C}_2{\rm H}_4^-$ | ||||
ER (0th order) (eV) | 2.191 | 2.091 | 1.988 | … |
Γ (0th order) (eV) | 0.436 | 0.430 | 0.447 | … |
ηopt (a.u.) | 0.0032 | 0.0046 | 0.0025 | … |
ER (1st order) (eV)a | 2.230 | 2.032 | 1.903 | … |
Γ (1st order) (eV)a | 0.302 | 0.328 | 0.373 | … |
$\eta ^R_{\rm opt}$ (a.u.) | 0.0210 | 0.0060 | 0.0054 | … |
$\eta ^I_{\rm opt}$ (a.u.) | 0.0248 | 0.0085 | 0.0043 | … |
Bound states of C2H4 | ||||
E (2 1Ag) (eV) | 8.315 | 8.446 | 8.493 | … |
E (1 1B1g) (eV) | 9.681 | 9.791 | 9.830 | … |
E (1 1B1u) (eV) | 7.279 | 7.392 | 7.436 | … |
. | aug-cc-pVDZ . | aug-cc-pVTZ . | aug-cc-pVQZ . | aug-cc-pV5Z . |
---|---|---|---|---|
. | +3s3p3d(C) . | +3s3p3d(C) . | +3s3p3d(C) . | +3s3p3d(C) . |
2Π resonance of CO− | ||||
ER (0th order) (eV) | 2.303 | 2.088 | 1.987 | 1.926 |
Γ (0th order) (eV) | 0.727 | 0.650 | 0.696 | 0.804 |
ηopt (a.u.) | 0.0046 | 0.0028 | 0.0020 | 0.0015 |
ER (1st order) (eV)a | 2.182 | 1.981 | 1.851 | 1.762 |
Γ (1st order) (eV)a | 0.667 | 0.585 | 0.673 | 0.604 |
$\eta ^R_{\rm opt}$ (a.u.) | 0.0175 | 0.0054 | 0.0062 | 0.0034 |
$\eta ^I_{\rm opt}$ (a.u.) | 0.0100 | 0.0048 | 0.0046 | 0.0028 |
Bound states of CO | ||||
E (2 1A1) (eV) | 10.777 | 10.961 | 11.021 | 11.046 |
E (4 1A1) (eV) | 12.446 | 12.597 | 12.642 | 12.663 |
E (1 1B2) (eV) | 8.703 | 8.626 | 8.612 | 8.608 |
2B2g resonance of ${\rm C}_2{\rm H}_4^-$ | ||||
ER (0th order) (eV) | 2.191 | 2.091 | 1.988 | … |
Γ (0th order) (eV) | 0.436 | 0.430 | 0.447 | … |
ηopt (a.u.) | 0.0032 | 0.0046 | 0.0025 | … |
ER (1st order) (eV)a | 2.230 | 2.032 | 1.903 | … |
Γ (1st order) (eV)a | 0.302 | 0.328 | 0.373 | … |
$\eta ^R_{\rm opt}$ (a.u.) | 0.0210 | 0.0060 | 0.0054 | … |
$\eta ^I_{\rm opt}$ (a.u.) | 0.0248 | 0.0085 | 0.0043 | … |
Bound states of C2H4 | ||||
E (2 1Ag) (eV) | 8.315 | 8.446 | 8.493 | … |
E (1 1B1g) (eV) | 9.681 | 9.791 | 9.830 | … |
E (1 1B1u) (eV) | 7.279 | 7.392 | 7.436 | … |
Identical to UR and −1/2 UI from Eqs. (7) and (8).
Concerning the resonance width, trends are less clear. For CO−, both zeroth-order and first-order values show a non-monotonic behavior with respect to the basis-set size and no convergence is observed. The magnitude of the variations in Γ is however comparable to that in ER. For
To gain further insight into the dependence of ER and Γ on the size of the valence basis set, we performed an energy decomposition analysis for the 2Π resonance of CO− and the 2B2g resonance of
where the CAP is considered as a part of Fpq and
Energy decomposition analysis for the real and imaginary parts of the energiesa of the 2Π resonance of CO− and the 2B2g resonance of
. | aug-cc-pVDZ . | aug-cc-pVTZ . | aug-cc-pVQZ . | aug-cc-pV5Z . |
---|---|---|---|---|
. | +3s3p3d(C) . | +3s3p3d(C) . | +3s3p3d(C) . | +3s3p3d(C) . |
2Π resonance of CO− | ||||
Total energy (real) | −112.9835 | −113.0979 | −113.1573 | −113.1805 |
One-electron part (real)b | −112.5652 | −112.5920 | −112.6078 | −112.6167 |
Two-electron part (real) | −0.4183 | −0.5059 | −0.5495 | −0.5638 |
Total energy (imaginary) | −0.0136 | −0.0121 | −0.0129 | −0.0148 |
One-electron part (imaginary) | −0.0479 | −0.0485 | −0.0559 | −0.0725 |
Two-electron part (imaginary) | 0.0343 | 0.0364 | 0.0430 | 0.0577 |
2B2g resonance of |${\rm C}_2{\rm H}_4^-$| | ||||
Total energy (real) | −78.2836 | −78.3881 | −78.4333 | … |
One-electron part (real)b | −77.8614 | −77.8781 | −77.8935 | … |
Two-electron part (real) | −0.4223 | −0.5100 | −0.5399 | … |
Total energy (imaginary) | −0.0080 | −0.0080 | −0.0082 | … |
One-electron part (imaginary) | −0.0289 | −0.0318 | −0.0389 | … |
Two-electron part (imaginary) | 0.0209 | 0.0238 | 0.0307 | … |
. | aug-cc-pVDZ . | aug-cc-pVTZ . | aug-cc-pVQZ . | aug-cc-pV5Z . |
---|---|---|---|---|
. | +3s3p3d(C) . | +3s3p3d(C) . | +3s3p3d(C) . | +3s3p3d(C) . |
2Π resonance of CO− | ||||
Total energy (real) | −112.9835 | −113.0979 | −113.1573 | −113.1805 |
One-electron part (real)b | −112.5652 | −112.5920 | −112.6078 | −112.6167 |
Two-electron part (real) | −0.4183 | −0.5059 | −0.5495 | −0.5638 |
Total energy (imaginary) | −0.0136 | −0.0121 | −0.0129 | −0.0148 |
One-electron part (imaginary) | −0.0479 | −0.0485 | −0.0559 | −0.0725 |
Two-electron part (imaginary) | 0.0343 | 0.0364 | 0.0430 | 0.0577 |
2B2g resonance of |${\rm C}_2{\rm H}_4^-$| | ||||
Total energy (real) | −78.2836 | −78.3881 | −78.4333 | … |
One-electron part (real)b | −77.8614 | −77.8781 | −77.8935 | … |
Two-electron part (real) | −0.4223 | −0.5100 | −0.5399 | … |
Total energy (imaginary) | −0.0080 | −0.0080 | −0.0082 | … |
One-electron part (imaginary) | −0.0289 | −0.0318 | −0.0389 | … |
Two-electron part (imaginary) | 0.0209 | 0.0238 | 0.0307 | … |
Evaluated at the respective ηopt, see values in Table IV.
Including nuclear repulsion energy.
One might be tempted to relate the basis set dependence of Γ to an insufficient description of the interaction of the resonance state with the continuum, but we point out that the addition of further diffuse functions has only little impact on ER and Γ (cf. Sec. IV C), which suggests the opposite. In sum, we feel that the behavior of Γ requires further investigation in order to develop a scheme for the extrapolation to the complete basis set limit, but such an extension is beyond the scope of the present article. We point out, however, that the variations in Γ observed for CO− do not exceed 0.15 eV so that reliable computations are still possible based on our current approach.
V. APPLICATIONS
In this section, we will report resonance positions and widths for a number of shape resonances of small to medium-size molecules and compare the results from our CAP-EOM-EA-CCSD scheme to those obtained using other theoretical approaches or through experiment. Systems included in this study are
We add that only the real part of the resonance wave function has a well defined single-attachment character, while the imaginary part has a considerably different form. Most often, it exhibits multiconfigurational character represented by several single attachments to very diffuse orbitals. Also, its dependence on the CAP strength is more pronounced than that of the real part. A detailed investigation of this phenomenon will be conducted in future work.
In all calculations, we employed the scheme developed in Sec. IV, i.e., we chose the CAP onset based on the spatial extent of the ground-state wave function and used the aug-cc-pVXZ+3s3p3d(C) bases. Computational details are compiled in Table VI.
Computational details of CAP-EOM-EA-CCSD calculations on
Molecule . | Leading configuration in the . | Bond lengths (in bohr) . | . | CAP onset (in bohr) . | ||
---|---|---|---|---|---|---|
(state) . | real part of the wave function . | and angles (in degrees) . | Orientation . | |$r^0_x$| . | |$r^0_y$| . | |$r^0_z$| . |
|${\rm N}_2^-$| | (1σg)2(1σu)2(2σg)2(2σu)2 | R(NN) = 2.0740 | z-axis = molecular axis | 2.76 | 2.76 | 4.88 |
(1 2Πg) | (3σg)2(1πu)4(1πg)1 | |||||
CO− | (1σ)2(2σ)2(3σ)2(4σ)2 | R(CO) = 2.1316 | z-axis = molecular axis | 2.76 | 2.76 | 4.97 |
(1 2Π) | (5σ)2(1π)4(2π)1 | |||||
|${\rm C}_2{\rm H}_2^-$| | (1σg)2(1σu)2(2σg)2(2σu)2 | R(CC) = 2.2733 | z-axis = molecular axis | 3.20 | 3.20 | 6.35 |
(1 2Πg) | (3σg)2(1πu)4(1πg)1 | R(CH) = 2.0088 | ||||
|${\rm C}_2{\rm H}_4^-$| | (1ag)2(1b1u)2(2ag)2(2b1u)2 | R(CC) = 2.5303 | z-axis ⊥ molecular plane | 7.10 | 4.65 | 3.40 |
(1 2B2g) | (1b2u)2(3ag)2(1b3g)2(1b3u)2(1b2g)1 | R(CH) = 2.0522 | x-axis = CC bond | |||
∠(CCH) = 121.2 | ||||||
CH2O− | (1a1)2(2a1)2(3a1)2(4a1)2 | R(CO) = 2.2771 | z-axis = CO bond | 3.85 | 2.95 | 6.10 |
(1 2B1) | (1b2)2(5a1)2(1b1)2(2b2)2(2b1)1 | R(CH) = 2.0995 | y-axis ⊥ molecular plane | |||
∠(HCO) = 121.9 | ||||||
|${\rm CO}_2^-$| | (1σg)2(1σu)2(2σg)2(2σu)2(3σg)2 | R(CO) = 2.1978 | z-axis = molecular axis | 3.33 | 3.33 | 9.57 |
(1 2Πu) | 1(πu)4(3σu)2(4σg)2(1πg)4(2πu)1 | |||||
C4|${\rm H}_6^-$| | 16.20 | 7.25 | 4.65 | |||
(1 2Au) | (core)2(6bu)2(7ag)2(1au)2(1bg)2(2au)1a | b | ||||
(1 2Bg) | (core)2(6bu)2(7ag)2(1au)2(1bg)2(2bg)1a |
Molecule . | Leading configuration in the . | Bond lengths (in bohr) . | . | CAP onset (in bohr) . | ||
---|---|---|---|---|---|---|
(state) . | real part of the wave function . | and angles (in degrees) . | Orientation . | |$r^0_x$| . | |$r^0_y$| . | |$r^0_z$| . |
|${\rm N}_2^-$| | (1σg)2(1σu)2(2σg)2(2σu)2 | R(NN) = 2.0740 | z-axis = molecular axis | 2.76 | 2.76 | 4.88 |
(1 2Πg) | (3σg)2(1πu)4(1πg)1 | |||||
CO− | (1σ)2(2σ)2(3σ)2(4σ)2 | R(CO) = 2.1316 | z-axis = molecular axis | 2.76 | 2.76 | 4.97 |
(1 2Π) | (5σ)2(1π)4(2π)1 | |||||
|${\rm C}_2{\rm H}_2^-$| | (1σg)2(1σu)2(2σg)2(2σu)2 | R(CC) = 2.2733 | z-axis = molecular axis | 3.20 | 3.20 | 6.35 |
(1 2Πg) | (3σg)2(1πu)4(1πg)1 | R(CH) = 2.0088 | ||||
|${\rm C}_2{\rm H}_4^-$| | (1ag)2(1b1u)2(2ag)2(2b1u)2 | R(CC) = 2.5303 | z-axis ⊥ molecular plane | 7.10 | 4.65 | 3.40 |
(1 2B2g) | (1b2u)2(3ag)2(1b3g)2(1b3u)2(1b2g)1 | R(CH) = 2.0522 | x-axis = CC bond | |||
∠(CCH) = 121.2 | ||||||
CH2O− | (1a1)2(2a1)2(3a1)2(4a1)2 | R(CO) = 2.2771 | z-axis = CO bond | 3.85 | 2.95 | 6.10 |
(1 2B1) | (1b2)2(5a1)2(1b1)2(2b2)2(2b1)1 | R(CH) = 2.0995 | y-axis ⊥ molecular plane | |||
∠(HCO) = 121.9 | ||||||
|${\rm CO}_2^-$| | (1σg)2(1σu)2(2σg)2(2σu)2(3σg)2 | R(CO) = 2.1978 | z-axis = molecular axis | 3.33 | 3.33 | 9.57 |
(1 2Πu) | 1(πu)4(3σu)2(4σg)2(1πg)4(2πu)1 | |||||
C4|${\rm H}_6^-$| | 16.20 | 7.25 | 4.65 | |||
(1 2Au) | (core)2(6bu)2(7ag)2(1au)2(1bg)2(2au)1a | b | ||||
(1 2Bg) | (core)2(6bu)2(7ag)2(1au)2(1bg)2(2bg)1a |
(core)2 = (1bu)2(1ag)2(2bu)2(2ag)2(3ag)2(3bu)2(4ag)2(4bu)2(5bu)2(5ag)2(6ag)2.
Cartesian coordinates of symmetry-unique nuclei in bohr: C(1.15735915, 0.76277974, 0.0), C(3.48113356, −0.22898634, 0.0), H(0.92483778, 2.80751355, 0.0), H(3.78431564, −2.26062161, 0.0), H(5.15268958, 0.95838430, 0.0).
A. 2Πg resonance in |${\rm N}_2^-$|
The scattering of slow electrons by the N2 molecule has been studied experimentally several times71–78 so that the 2Πg resonance of
CAP-EOM-EA-CCSD results obtained for the resonance position and width are compiled in Table VII together with several values available from the literature. The optimal CAP strengths found in our calculations are 0.0015 (0.0037) a.u. for the zeroth-order result and 0.0119 (0.0025) a.u. and 0.0148 (0.0071) a.u. for the real and imaginary part of the first-order result obtained with the aug-cc-pVTZ+3s3p3d(C) (aug-cc-pVQZ+3s3p3d(C)) basis set. We refrained from including experimental results in Table VII except for the fixed-nuclei estimate by Berman et al.89 (ER = 2.32 eV, Γ = 0.41 eV), which is usually considered as the reference value in previous theoretical studies. We add that this value was derived through a fit to the experimental data using Feshbach's projection operator formalism.
Resonance positions ER and widths Γ for the 2Πg resonance state of |${\rm N}_2^-$| obtained using different methods.
Method . | ER (eV) . | Γ (eV) . |
---|---|---|
Stieltjes imaging technique/special basis seta | 2.23 | 0.40 |
Schwinger variational method/ADC(3)/(11s8p3d)/[5s7p3d]b | 2.534 | 0.536 |
Complex scaling/HF-SCF/Dunning's (9s5p)/[5s3p]+2d+(5p2d)/[3p2d]+4p6dc | 3.19 | 0.44 |
Complex scaling/MR-CI/Dunning's (9s5p)/[5s3p]+1d+10pd | 1.38 | 0.414 |
Complex scaling/Σ3 decouplings of the e−-propagator/[4s9p]e | 2.11 | 0.18 |
Stabilization method/MR-CI/ Dunning's (9s5p)/[5s3p]+3p2d+4s1p1d(C)f | 2.62 | 0.45 |
Stabilization method/MR-CI/6-31+G*+3pg | 2.34 | 0.51 |
Stabilization method/MP-PT2/ANO(14s9p4d3f)/[4s3p2d1f]+2s2p7d4g(C)h | 2.36 | 0.42 |
Stabilization method/CIS/aug-cc-pVTZ+3pi | 3.77 | 1.14 |
Stabilization method/TDDFT(HFE_PBE)/aug-cc-pVTZ+3pi | 3.078 | 0.54 |
Stabilization method/EOM-EA-CCSD/aug-cc-pVTZ+3pi | 2.58 | 0.570 |
Stabilization method/EOM-EA-CCSD/aug-cc-pVQZ+3pi | 2.49 | 0.502 |
Stabilization method/EOM-EA-CCSD/aug-cc-pV5Z+3pi | 2.49 | 0.496 |
CAP/static exchange/[5s10p13d] (0th order)j | 3.888 | 1.363 |
CAP/static exchange/[5s10p13d] (1st order)j | 3.776 | 1.199 |
CAP-HF-SCF/(11s7p2d)/[5s4p2d]k | 3.28 | 0.395 |
CAP-DFT(LSD/XC)/(11s7p2d)/[5s4p2d]k | 3.39 | 0.506 |
CAP-MRCI/Dunning's (9s5p)/[5s3p]+(12p)/[9p]+2dl | 2.97 | 0.65 |
TCAP-MRCI/Dunning's (11s6p)/[5s3p]+7p3d2fm | 2.42 | 0.45 |
CAP-Σ(ADC(2))/TZP+9p2d2fn | 2.58 | 0.55 |
CAP-FSMRCC/TZ(7p2d)o | 2.52 | 0.39 |
CAP-CIP/TZ(7p2d)p | 2.28 | 0.482 |
CAP-EOM-EA-CCSD/(11s8p3d)/[5s7p3d]+3pq | 2.07 | 0.42 |
CAP-EOM-EA-CCSD/aug-cc-pVTZ+3s3p3d(C) (0th order)r | 2.487 | 0.417 |
CAP-EOM-EA-CCSD/aug-cc-pVTZ+3s3p3d(C) (1st order)r | 2.571 | 0.255 |
CAP-EOM-EA-CCSD/aug-cc-pVQZ+3s3p3d(C) (0th order)r | 2.508 | 0.364 |
CAP-EOM-EA-CCSD/aug-cc-pVQZ+3s3p3d(C) (1st order)r | 2.478 | 0.286 |
Estimate via Feshbach projection formalism based on experimental datas | 2.32 | 0.41 |
Method . | ER (eV) . | Γ (eV) . |
---|---|---|
Stieltjes imaging technique/special basis seta | 2.23 | 0.40 |
Schwinger variational method/ADC(3)/(11s8p3d)/[5s7p3d]b | 2.534 | 0.536 |
Complex scaling/HF-SCF/Dunning's (9s5p)/[5s3p]+2d+(5p2d)/[3p2d]+4p6dc | 3.19 | 0.44 |
Complex scaling/MR-CI/Dunning's (9s5p)/[5s3p]+1d+10pd | 1.38 | 0.414 |
Complex scaling/Σ3 decouplings of the e−-propagator/[4s9p]e | 2.11 | 0.18 |
Stabilization method/MR-CI/ Dunning's (9s5p)/[5s3p]+3p2d+4s1p1d(C)f | 2.62 | 0.45 |
Stabilization method/MR-CI/6-31+G*+3pg | 2.34 | 0.51 |
Stabilization method/MP-PT2/ANO(14s9p4d3f)/[4s3p2d1f]+2s2p7d4g(C)h | 2.36 | 0.42 |
Stabilization method/CIS/aug-cc-pVTZ+3pi | 3.77 | 1.14 |
Stabilization method/TDDFT(HFE_PBE)/aug-cc-pVTZ+3pi | 3.078 | 0.54 |
Stabilization method/EOM-EA-CCSD/aug-cc-pVTZ+3pi | 2.58 | 0.570 |
Stabilization method/EOM-EA-CCSD/aug-cc-pVQZ+3pi | 2.49 | 0.502 |
Stabilization method/EOM-EA-CCSD/aug-cc-pV5Z+3pi | 2.49 | 0.496 |
CAP/static exchange/[5s10p13d] (0th order)j | 3.888 | 1.363 |
CAP/static exchange/[5s10p13d] (1st order)j | 3.776 | 1.199 |
CAP-HF-SCF/(11s7p2d)/[5s4p2d]k | 3.28 | 0.395 |
CAP-DFT(LSD/XC)/(11s7p2d)/[5s4p2d]k | 3.39 | 0.506 |
CAP-MRCI/Dunning's (9s5p)/[5s3p]+(12p)/[9p]+2dl | 2.97 | 0.65 |
TCAP-MRCI/Dunning's (11s6p)/[5s3p]+7p3d2fm | 2.42 | 0.45 |
CAP-Σ(ADC(2))/TZP+9p2d2fn | 2.58 | 0.55 |
CAP-FSMRCC/TZ(7p2d)o | 2.52 | 0.39 |
CAP-CIP/TZ(7p2d)p | 2.28 | 0.482 |
CAP-EOM-EA-CCSD/(11s8p3d)/[5s7p3d]+3pq | 2.07 | 0.42 |
CAP-EOM-EA-CCSD/aug-cc-pVTZ+3s3p3d(C) (0th order)r | 2.487 | 0.417 |
CAP-EOM-EA-CCSD/aug-cc-pVTZ+3s3p3d(C) (1st order)r | 2.571 | 0.255 |
CAP-EOM-EA-CCSD/aug-cc-pVQZ+3s3p3d(C) (0th order)r | 2.508 | 0.364 |
CAP-EOM-EA-CCSD/aug-cc-pVQZ+3s3p3d(C) (1st order)r | 2.478 | 0.286 |
Estimate via Feshbach projection formalism based on experimental datas | 2.32 | 0.41 |
Table VII illustrates that a confusing plethora of values for the resonance position and width of the 2Πg state of
Table VII also shows that most approaches yield too high values for the resonance position, whereas the reference value for the resonance width from Ref. 89 was often surprisingly well reproduced. The impact of electron correlation is illustrated by the comparison of high-level correlated methods to lower levels of theory. The position of the resonance state is consistently calculated to be above 3 eV with HF, density functional theory (DFT), and CIS (configuration interaction singles) based methods, regardless of whether stabilization techniques, complex scaling, or CAPs are employed, while the use of correlated methods leads to a significantly better agreement with the reference value. The only notable exception is the complex-scaled MRCI (multireference configuration interaction) result (1.38 eV) from Ref. 10, which is almost 1 eV below the reference value. Interestingly, CAP-augmented MRCI calculations63 with a rather similar basis set yielded a quite different resonance position (2.97 eV).
Compared to its effect on the resonance position, the role of electron correlation for the width Γ is less clear. For example, the complex-scaled HF and CAP-HF calculations from Refs. 81 and 27 agree with the reference value within 0.03 eV and 0.015 eV, respectively, while some correlated calculations led to deviations of more than 0.2 eV.63 In fact, it was stated explicitly62 that several values reported in the literature might have benefitted from error cancellation. We note that most authors reported values for Γ that were higher than the reference value with some low-level approaches overestimating the width by a factor of more than two, whereas our calculations underestimated Γ.
Regarding the basis-set dependence, a comparison between different schemes is hampered by the fact that a variety of different basis sets has been used in previous studies. Bearing in mind our findings from Sec. IV D, it seems justified to conclude for CAP-based methods that basis-set effects may account at least partly for the differences between values for Γ reported by different authors. In addition, the results from Ref. 80 suggest that, as compared to our CAP-augmented EOM-EA-CCSD scheme, the stabilization method combined with EOM-EA-CCSD leads to a somewhat faster convergence with respect to basis-set size. However, the highest-level (aug-cc-pV5Z+3p) results obtained with the latter method (ER = 2.49 eV, Γ = 0.496 eV) still show a sizable deviation from the reference values.
B. 2Π resonance in CO−
While we have employed the 2Π resonance of CO− as a test system in Sec. IV, we will consider it in this section with a different focus, i.e., we will compare our results to those obtained using other methods. In contrast to
Resonance positions ER and widths Γ for the 2Π resonance state of CO− obtained using different methods.
Method . | ER (eV) . | Γ (eV) . |
---|---|---|
Boomerang modela | 1.52 | 0.80 |
Close coupling methodb | 1.75 | 0.28 |
T-matrix/static exchange/(9s5p1d)/[4s3p1d]c | 3.4 | 1.65 |
Complex scaling/Σ2 decouplings of the e− propagator (real SCF)/4s5pd | 1.71 | 0.08 |
Complex scaling/Σ3 decouplings of the e− propagator/4s5pe | 1.65 | 0.14 |
Stabilization method/MP-PT2/ANO(14s9p4d3f)/[4s3p2d1f]+2s4p7d5f(C)f | 2.02 | 0.35 |
CAP-EOM-EA-CCSD/4s5p(C)+4s5p1d(O)g | 1.32 | 0.12 |
CAP-EOM-EA-CCSD/maug-cc-pV(D+d)Z+3pg | 1.42 | 0.44 |
CAP-EOM-EA-CCSD/aug-cc-pVTZ+3s3p(A) (1st order)h | 1.954 | 0.433 |
CAP-EOM-EA-CCSD/aug-cc-pVTZ+3s3p3d(C) (0th order)i | 2.088 | 0.650 |
CAP-EOM-EA-CCSD/aug-cc-pVTZ+3s3p3d(C) (1st order)i | 1.981 | 0.585 |
CAP-EOM-EA-CCSD/aug-cc-pV5Z+3s3p3d(C) (0th order)i | 1.926 | 0.804 |
CAP-EOM-EA-CCSD/aug-cc-pV5Z+3s3p3d(C) (1st order)i | 1.762 | 0.604 |
Experimentj | 1.50 | 0.40 |
Method . | ER (eV) . | Γ (eV) . |
---|---|---|
Boomerang modela | 1.52 | 0.80 |
Close coupling methodb | 1.75 | 0.28 |
T-matrix/static exchange/(9s5p1d)/[4s3p1d]c | 3.4 | 1.65 |
Complex scaling/Σ2 decouplings of the e− propagator (real SCF)/4s5pd | 1.71 | 0.08 |
Complex scaling/Σ3 decouplings of the e− propagator/4s5pe | 1.65 | 0.14 |
Stabilization method/MP-PT2/ANO(14s9p4d3f)/[4s3p2d1f]+2s4p7d5f(C)f | 2.02 | 0.35 |
CAP-EOM-EA-CCSD/4s5p(C)+4s5p1d(O)g | 1.32 | 0.12 |
CAP-EOM-EA-CCSD/maug-cc-pV(D+d)Z+3pg | 1.42 | 0.44 |
CAP-EOM-EA-CCSD/aug-cc-pVTZ+3s3p(A) (1st order)h | 1.954 | 0.433 |
CAP-EOM-EA-CCSD/aug-cc-pVTZ+3s3p3d(C) (0th order)i | 2.088 | 0.650 |
CAP-EOM-EA-CCSD/aug-cc-pVTZ+3s3p3d(C) (1st order)i | 1.981 | 0.585 |
CAP-EOM-EA-CCSD/aug-cc-pV5Z+3s3p3d(C) (0th order)i | 1.926 | 0.804 |
CAP-EOM-EA-CCSD/aug-cc-pV5Z+3s3p3d(C) (1st order)i | 1.762 | 0.604 |
Experimentj | 1.50 | 0.40 |
Regarding the resonance position, Table VIII shows that most theoretical values are higher than the experimental value (1.50 eV)78,94 as in the case of
The available values for the resonance width of CO− differ by more than an order of magnitude as again illustrated by Table VIII. The largest value reported (1.65 eV) was obtained in the static exchange approximation, while the narrowest width (0.08 eV) was computed from the Σ2 decouplings of the electron propagator, a pattern that is again similar to
C. 2Πg resonance in |${\rm C}_2{\rm H}_2^-$|
Resonance positions ER and widths Γ for the 2Πg resonance state of
Method . | ER (eV) . | Γ (eV) . |
---|---|---|
Theory | ||
Multiple scattering Xαa | 2.6 | 1.0 |
Feshbach projection/MR-CI/Dunning's (9s5p)/[5s3p]+1p1d+3pb | 2.96 | 1.11 |
Complex scaling/Σ3 decouplings of the e−-propagator/5s9p1d, 3s3p(H)c | 2.50 | 0.21 |
Stabilization method/EOM-EA-CCSD/aug-cc-pVTZ+3pd | 2.77 | 1.50 |
Stabilization method/TDDFT(HFE_PBE)/aug-cc-pVTZ+3pd | 2.4 | 0.6 |
CAP-EOM-EA-CCSD/Dunning's (9s5p)/[5s3p]+4p1d(C), 2s1p(H)e | 1.79 | 0.80 |
CAP-EOM-EA-CCSD/aug-cc-pVTZ+3s3p3d(C) (0th order)f | 2.655 | 0.979 |
CAP-EOM-EA-CCSD/aug-cc-pVTZ+3s3p3d(C) (1st order)f | 2.450 | 0.831 |
Experiment | ||
Trapped electrong | 1.80/1.85 | … |
Vibrational excitationh | 2.6 | >1.0 |
Electron impacti | 2.5 | … |
Dissociative attachment/electron transmissionj | 2.6 | … |
Electron transmissionk | 2.6 | ∼0.8 |
Method . | ER (eV) . | Γ (eV) . |
---|---|---|
Theory | ||
Multiple scattering Xαa | 2.6 | 1.0 |
Feshbach projection/MR-CI/Dunning's (9s5p)/[5s3p]+1p1d+3pb | 2.96 | 1.11 |
Complex scaling/Σ3 decouplings of the e−-propagator/5s9p1d, 3s3p(H)c | 2.50 | 0.21 |
Stabilization method/EOM-EA-CCSD/aug-cc-pVTZ+3pd | 2.77 | 1.50 |
Stabilization method/TDDFT(HFE_PBE)/aug-cc-pVTZ+3pd | 2.4 | 0.6 |
CAP-EOM-EA-CCSD/Dunning's (9s5p)/[5s3p]+4p1d(C), 2s1p(H)e | 1.79 | 0.80 |
CAP-EOM-EA-CCSD/aug-cc-pVTZ+3s3p3d(C) (0th order)f | 2.655 | 0.979 |
CAP-EOM-EA-CCSD/aug-cc-pVTZ+3s3p3d(C) (1st order)f | 2.450 | 0.831 |
Experiment | ||
Trapped electrong | 1.80/1.85 | … |
Vibrational excitationh | 2.6 | >1.0 |
Electron impacti | 2.5 | … |
Dissociative attachment/electron transmissionj | 2.6 | … |
Electron transmissionk | 2.6 | ∼0.8 |
Table IX shows that our results for the resonance position are in qualitative agreement with those obtained from most experiments as well as from other theoretical approaches. Only when using the trapped electron method98,99,103 considerably lower (∼0.7 eV) resonance positions were found. Our zeroth-order CAP-EOM-EA-CCSD result (2.655 eV) agrees within 0.05 eV with the experimental values from Refs. 99 and 100, and 102. We note that the first-order correction lowers the CAP-EOM-EA-CCSD result for ER by about 0.2 eV bringing it closer to the experimental value from Ref. 101, but basis-set effects may have an impact of similar magnitude.
Theoretical values for the resonance width vary between 0.19 eV and 1.11 eV and only two rough estimates of 0.8 eV99 and >1.0 eV100 are available from experiment. Our calculations qualitatively confirm these two estimates with the zeroth-order result (0.979 eV) being closer to one value and the first-order result (0.831 eV) being closer to the other value. As mentioned for the resonance position, basis-set effects may have a sizable impact so that an ultimate decision between the two experimental values cannot be made.
D. 2B2g resonance in |${\rm C}_2{\rm H}_4^-$|
Similar to CO−, the 2B2g resonance in
Resonance positions ER and widths Γ for the 2B2g resonance state of
Method . | ER (eV) . | Γ (eV) . |
---|---|---|
Theory | ||
Complex scaling/second order rotated propagator/5s7pa | 1.94 | 0.110 |
Complex scaling/second order rotated propagator/5s8pa | 2.49 | 0.234 |
Complex scaling/second order rotated propagator/5s9pa | 1.88 | 0.442 |
Complex Kohn methodb | 1.83 | 0.460 |
Complex scaling/bi-variational SCF/5s7pc | 1.93 | 0.2 |
Complex scaling/ Second order biorthogonal e− propagator/5s7pd | 1.86 | 0.18 |
Complex scaling/ Diagonal 2ph-TDA biorthogonal e− propagator/5s7pd | 1.89 | 0.18 |
Stabilization method/EOM-CCSD/aug-cc-pVTZ+3pe | 2.06 | 0.64 |
Stabilization method/EOM-MP2/aug-cc-pVTZ+3pe | 1.91 | 0.60 |
Stabilization method/ADC(2)/aug-cc-pVTZ+3pe | 1.78 | 0.49 |
Stabilization method/Koopmans theorem (HFE_BLYP)/aug-cc-pVTZ+3pe | 2.58 | 1.32 |
Stabilization method/Koopmans theorem (HFE_BPE)/aug-cc-pVTZ+3pe | 2.62 | 1.08 |
Stabilization method/TDDFT (HFE_BPE)/aug-cc-pVTZ+3pe | 2.49 | 0.31 |
CAP-CIP- $V^{(1,0)}_c (\eta ^2)$ /5s9pf | 1.778 | 0.9076 |
CAP-FSMRCC- $V^{(1,0)}_c (\eta ^2)$ /aug-cc-pVDZg | 1.802 | 0.3662 |
CAP-EOM-EA-CCSD/aug-cc-pVTZ+3s3p3d(C) (0th order)h | 2.091 | 0.430 |
CAP-EOM-EA-CCSD/aug-cc-pVTZ+3s3p3d(C) (1st order)h | 2.032 | 0.328 |
CAP-EOM-EA-CCSD/aug-cc-pVQZ+3s3p3d(C) (0th order)h | 1.988 | 0.447 |
CAP-EOM-EA-CCSD/aug-cc-pVQZ+3s3p3d(C) (1st order)h | 1.903 | 0.373 |
Experiment | ||
Electron scatteringi | 1.78 | … |
Electron impactj | 1.8 | 0.7 |
Elastic scatteringk | 1.8 | 0.7 |
Method . | ER (eV) . | Γ (eV) . |
---|---|---|
Theory | ||
Complex scaling/second order rotated propagator/5s7pa | 1.94 | 0.110 |
Complex scaling/second order rotated propagator/5s8pa | 2.49 | 0.234 |
Complex scaling/second order rotated propagator/5s9pa | 1.88 | 0.442 |
Complex Kohn methodb | 1.83 | 0.460 |
Complex scaling/bi-variational SCF/5s7pc | 1.93 | 0.2 |
Complex scaling/ Second order biorthogonal e− propagator/5s7pd | 1.86 | 0.18 |
Complex scaling/ Diagonal 2ph-TDA biorthogonal e− propagator/5s7pd | 1.89 | 0.18 |
Stabilization method/EOM-CCSD/aug-cc-pVTZ+3pe | 2.06 | 0.64 |
Stabilization method/EOM-MP2/aug-cc-pVTZ+3pe | 1.91 | 0.60 |
Stabilization method/ADC(2)/aug-cc-pVTZ+3pe | 1.78 | 0.49 |
Stabilization method/Koopmans theorem (HFE_BLYP)/aug-cc-pVTZ+3pe | 2.58 | 1.32 |
Stabilization method/Koopmans theorem (HFE_BPE)/aug-cc-pVTZ+3pe | 2.62 | 1.08 |
Stabilization method/TDDFT (HFE_BPE)/aug-cc-pVTZ+3pe | 2.49 | 0.31 |
CAP-CIP- $V^{(1,0)}_c (\eta ^2)$ /5s9pf | 1.778 | 0.9076 |
CAP-FSMRCC- $V^{(1,0)}_c (\eta ^2)$ /aug-cc-pVDZg | 1.802 | 0.3662 |
CAP-EOM-EA-CCSD/aug-cc-pVTZ+3s3p3d(C) (0th order)h | 2.091 | 0.430 |
CAP-EOM-EA-CCSD/aug-cc-pVTZ+3s3p3d(C) (1st order)h | 2.032 | 0.328 |
CAP-EOM-EA-CCSD/aug-cc-pVQZ+3s3p3d(C) (0th order)h | 1.988 | 0.447 |
CAP-EOM-EA-CCSD/aug-cc-pVQZ+3s3p3d(C) (1st order)h | 1.903 | 0.373 |
Experiment | ||
Electron scatteringi | 1.78 | … |
Electron impactj | 1.8 | 0.7 |
Elastic scatteringk | 1.8 | 0.7 |
Similar to the trends observed for the previously discussed systems, the largest values for both position and width are found by DFT in combination with the stabilization technique,80 whereas the narrowest widths are obtained when using electron propagator methods.110 The resonance position obtained by CAP-EOM-EA-CCSD lies reasonably close (within 0.3 eV) to the experimental value (1.8 eV) for all basis sets used. Similar to results from EOM-CCSD calculations using stabilization techniques,80 the value of the resonance position is overestimated by CAP-EOM-EA-CCSD, but we observe a positive trend when enlarging the valence part of the basis set from triple zeta to quadruple zeta. Our best estimate for the position of the resonance (first-order CAP-EOM-EA-CCSD/aug-cc-pVQZ+3s3p3d(C)) is 1.903 eV, which differs from the experimental value by only 0.1 eV and is thus not worse than what is usually found in EOM-EA-CCSD calculations for bound states.
The resonance width calculated with CAP-EOM-EA-CCSD is underestimated in comparison to experiment by roughly a factor of two. However, similar to the position, the increase of the valence basis set size brings the theoretical value of the width closer to the experimental one. But in contrast to the position, the first-order correction worsens the value for the width as compared to experiment so that the best estimate from our calculations (0.373 eV, first-order CAP-EOM-EA-CCSD/aug-cc-pVQZ+3s3p3d(C)) still deviates by more than 0.3 eV. We add that we observed a similar underestimation of the resonance width in Sec. V A for
E. 2B1 resonance in CH2O−
Formaldehyde is of particular interest for our study since it is the smallest molecule containing the highly polar carbonyl group, which means that an accurate description of polarization and correlation effects is especially important. Zeroth-order and first-order estimates of the position and the width of the 2B1 resonance state of CH2O− along with previous experimental and theoretical data are compiled in Table XI. The optimal CAP strength for the zeroth-order CAP-EOM-EA-CCSD values is 0.01 a.u. and for the corresponding first-order values 0.024 a.u. and 0.021 a.u. for the real and imaginary part, respectively.
Resonance positions ER and widths Γ for the 2B1 resonance state of CH2O− obtained using different methods.
Method . | ER (eV) . | Γ (eV) . |
---|---|---|
Theory | ||
Complex Kohn methoda | 1.0 | 0.5 |
Static exchangea | 3.0 | … |
Complex scaling/zeroth-order e− propagator/4s6p1d(C)/2s1p(H)b | 1.0 | 0.1 |
Complex scaling/quasiparticle second-order e− propagator/4s6p1d(C)/2s1p(H)b | 0.99 | 0.1 |
Complex scaling/diagonal 2ph-TDA e− propagator/4s6p1d(C)/2s1p(H)b | 0.89 | 0.12 |
R-matrix method/augmented DZPc | 1.32 | 0.546 |
R-matrix method/DZPd | 1.46 | 0.794 |
Finite-element-discrete-model methode | 0.682 | 0.429 |
CAP-SAC-CI/cc-pVDZ+[2s5p2d/2s2p] f | 1.219 | 0.488 |
CAP-SAC-CI/cc-pVTZ+[2s5p2d/2s2p] f | 1.119 | 0.462 |
CAP-SAC-CI/cc-pVQZ+[2s5p2d/2s2p] f | 1.094 | 0.418 |
CAP-EOM-EA-CCSD/aug-cc-pVTZ+3s3p3d(C) (0th order)g | 1.352 | 0.376 |
CAP-EOM-EA-CCSD/aug-cc-pVTZ+3s3p3d(C) (1st order)g | 1.314 | 0.277 |
Experiment | ||
Electron transmission spectroscopyh | 0.86 | … |
Vibrational excitationi | 0.87 | … |
Method . | ER (eV) . | Γ (eV) . |
---|---|---|
Theory | ||
Complex Kohn methoda | 1.0 | 0.5 |
Static exchangea | 3.0 | … |
Complex scaling/zeroth-order e− propagator/4s6p1d(C)/2s1p(H)b | 1.0 | 0.1 |
Complex scaling/quasiparticle second-order e− propagator/4s6p1d(C)/2s1p(H)b | 0.99 | 0.1 |
Complex scaling/diagonal 2ph-TDA e− propagator/4s6p1d(C)/2s1p(H)b | 0.89 | 0.12 |
R-matrix method/augmented DZPc | 1.32 | 0.546 |
R-matrix method/DZPd | 1.46 | 0.794 |
Finite-element-discrete-model methode | 0.682 | 0.429 |
CAP-SAC-CI/cc-pVDZ+[2s5p2d/2s2p] f | 1.219 | 0.488 |
CAP-SAC-CI/cc-pVTZ+[2s5p2d/2s2p] f | 1.119 | 0.462 |
CAP-SAC-CI/cc-pVQZ+[2s5p2d/2s2p] f | 1.094 | 0.418 |
CAP-EOM-EA-CCSD/aug-cc-pVTZ+3s3p3d(C) (0th order)g | 1.352 | 0.376 |
CAP-EOM-EA-CCSD/aug-cc-pVTZ+3s3p3d(C) (1st order)g | 1.314 | 0.277 |
Experiment | ||
Electron transmission spectroscopyh | 0.86 | … |
Vibrational excitationi | 0.87 | … |
See Refs. 115 and 116.
See Ref. 117.
See Ref. 118.
See Ref. 119.
See Ref. 120.
See Ref. 25.
This work.
See Refs. 111 and 112.
See Ref. 113.
Experiments by electron transmission spectros-copy111,112 and vibrational excitation113 report values of 0.86-0.87 eV for the resonance position. For the resonance width, no experimental value is available from the literature, but only an estimate based on electron collision experiments near 1 eV,113 which shows the lifetime to be of the same order of magnitude as the period of the ν2 vibrational mode (0.216 eV for neutral formaldehyde114). This vibrational excitation (ν2) corresponds to the CO stretch mode, which is mainly excited after autodetachment of the electron from the 2B1 resonance state.113 We deduce that the width of the resonance should be of the order of 0.1 eV as it is the case for the π* resonances of the molecules discussed before.
Previously reported theoretical values vary from 0.682 eV to 3.0 eV for the position and from 0.1 to 0.794 eV for the width of the resonance.25,115–120 Similar to the molecules considered above, the static exchange approximation overestimates the position of the resonance by roughly a factor of three.115,116 This can be explained by the high polarity of the carbonyl group and the reorganization effects when the molecule undergoes electron attachment, which results in strong correlation between the incident electron and the electrons of the neutral formaldehyde and shows the need for an accurate treatment of electron correlation.
We also note that the smallest width (0.1-0.12 eV) is reported for electron propagator methods,117 a pattern similar to
F. 2Πu resonance in |${\rm CO}_2^-$|
The scattering of slow electrons by the CO2 molecule is well characterized experimentally.121–127 Besides higher-lying resonance states, the existence of a 2Πu metastable state in the range of 3.8-4 eV was established. Theoretically, this system has been studied most often with a special emphasis on the changes when going from the linear to a bent structure, where the 2Πu state splits into the 2A1 and 2B2 components.128–131 The role of a virtual state near 2 eV132 and the interplay with other resonance states131 have also been investigated. Somewhat surprisingly, the position and width of the 2Πu resonance of the linear molecule have not yet been studied using high-level quantum-chemical methods but only within the static exchange approximation.133–136
Table XII reports the results from the CAP-EOM-EA-CCSD calculations for the 2Πu resonance of linear
Resonance positions ER and widths Γ for the 2Πu resonance state of
Method . | ER (eV) . | Γ (eV) . |
---|---|---|
Theory | ||
Scattering/static exchange+polarization/DZP basis seta | 3.8 | 0.5 |
Schwinger variational method/static exchange/[5p4d/5p4d1f]b | 5.39 | 0.64 |
Schwinger variational method/static exchange+polarization/[5s3p]+4s3p3dc | 3.78 | 0.23 |
Close coupling/static exchange+polarization/DZP+add. diffuse functionsd | 3.88 | 0.34 |
CAP-EOM-EA-CCSD/aug-cc-pVTZ+3s3p3d(C) (0th order)e | 4.020 | 0.119 |
CAP-EOM-EA-CCSD/aug-cc-pVTZ+3s3p3d(C) (1st order)e | 3.997 | 0.198 |
Experiment | ||
Electron scatteringf | 3.8 | … |
Electron transmission spectroscopyg | 3.14h | 0.20 ± 0.07 |
Electron impacti | 3.8 | … |
Impact of slow electronsj | 3.6 | … |
Electron transmission spectroscopyk | 3.58 | … |
High resolution attachment spectrometryl | 4.4 | … |
Method . | ER (eV) . | Γ (eV) . |
---|---|---|
Theory | ||
Scattering/static exchange+polarization/DZP basis seta | 3.8 | 0.5 |
Schwinger variational method/static exchange/[5p4d/5p4d1f]b | 5.39 | 0.64 |
Schwinger variational method/static exchange+polarization/[5s3p]+4s3p3dc | 3.78 | 0.23 |
Close coupling/static exchange+polarization/DZP+add. diffuse functionsd | 3.88 | 0.34 |
CAP-EOM-EA-CCSD/aug-cc-pVTZ+3s3p3d(C) (0th order)e | 4.020 | 0.119 |
CAP-EOM-EA-CCSD/aug-cc-pVTZ+3s3p3d(C) (1st order)e | 3.997 | 0.198 |
Experiment | ||
Electron scatteringf | 3.8 | … |
Electron transmission spectroscopyg | 3.14h | 0.20 ± 0.07 |
Electron impacti | 3.8 | … |
Impact of slow electronsj | 3.6 | … |
Electron transmission spectroscopyk | 3.58 | … |
High resolution attachment spectrometryl | 4.4 | … |
G. 2Au and 2Bg resonances in C4|${\rm H}_6^-$|
1,3-butadiene is different from all species discussed above in that its π system extends over more than a single double bond. Two low-lying π* resonances of 2Au and 2Bg symmetry result from this electronic structure, both of which have been characterized experimentally.104,137 However, while experimental values for the resonance position (0.62 eV and 2.82 eV) are available, no values for the width of either state have been reported in the literature. It was only concluded that the lower lying 2Au state should be longer lived as its spectrum exhibits vibrational structure. As for previous theoretical treatments of these resonances, only one study on the 2Au state employing conventional DFT/B3LYP, which found a surprisingly good agreement with experiment for the resonance position (0.76 eV), is available from the literature,138 but the resonance widths have apparently never been studied theoretically.
In Table XIII, we report CAP-EOM-EA-CCSD results for the resonance positions and widths of both π* resonances of 1,3-butadiene. For technical reasons, we employed the aug-cc-pVDZ+3s3p3d(C) basis instead of the aug-cc-pVTZ+3s3p3d(C) basis, but in order to test again the validity of the “C” as compared to the “A” scheme, results obtained with the larger aug-cc-pVDZ+3s3p3d(A) basis are also included in Table XIII. Optimal CAP strengths corresponding to the results for the 2Au state in Table XIII are 0.0074 a.u, 0.0115 a.u., and 0.0210 a.u. for the zeroth-order and first-order CAP-EOM-EA-CCSD calculations with the “C” basis set and 0.0135 a.u., 0.0175 a.u., and 0.0310 a.u. for the respective calculations with the “A” basis set. For the 2Bg state, optimal CAP strengths of 0.0098 a.u., 0.0270 a.u., and 0.0190 a.u. were obtained with the “C” basis set and of 0.0165 a.u., 0.0170 a.u., and 0.0350 a.u. with the “A” basis set.
Resonance positions ER and widths Γ for the 2Au and 2Bg resonance states of C4
Method . | ER (eV) . | Γ (eV) . |
---|---|---|
2Au state | ||
Theory | ||
DFT/B3-LYP/6-311+G(2df,p)a | 0.76 | … |
CAP-EOM-EA-CCSD/aug-cc-pVDZ+3s3p3d(C) (0th order)b | 1.336 | 0.110 |
CAP-EOM-EA-CCSD/aug-cc-pVDZ+3s3p3d(C) (1st order)b | 1.327 | 0.059 |
CAP-EOM-EA-CCSD/aug-cc-pVDZ+3s3p3d(A) (0th order)b | 1.348 | 0.145 |
CAP-EOM-EA-CCSD/aug-cc-pVDZ+3s3p3d(A) (1st order)b | 1.332 | 0.103 |
Experimentc | 0.62 | … |
2Bg state | ||
Theory | ||
CAP-EOM-EA-CCSD/aug-cc-pVDZ+3s3p3d(C) (0th order)b | 2.683 | 0.720 |
CAP-EOM-EA-CCSD/aug-cc-pVDZ+3s3p3d(C) (1st order)b | 2.538 | 0.509 |
CAP-EOM-EA-CCSD/aug-cc-pVDZ+3s3p3d(A) (0th order)b | 2.647 | 0.919 |
CAP-EOM-EA-CCSD/aug-cc-pVDZ+3s3p3d(A) (1st order)b | 2.544 | 0.630 |
Experimentc | 2.82 | … |
Method . | ER (eV) . | Γ (eV) . |
---|---|---|
2Au state | ||
Theory | ||
DFT/B3-LYP/6-311+G(2df,p)a | 0.76 | … |
CAP-EOM-EA-CCSD/aug-cc-pVDZ+3s3p3d(C) (0th order)b | 1.336 | 0.110 |
CAP-EOM-EA-CCSD/aug-cc-pVDZ+3s3p3d(C) (1st order)b | 1.327 | 0.059 |
CAP-EOM-EA-CCSD/aug-cc-pVDZ+3s3p3d(A) (0th order)b | 1.348 | 0.145 |
CAP-EOM-EA-CCSD/aug-cc-pVDZ+3s3p3d(A) (1st order)b | 1.332 | 0.103 |
Experimentc | 0.62 | … |
2Bg state | ||
Theory | ||
CAP-EOM-EA-CCSD/aug-cc-pVDZ+3s3p3d(C) (0th order)b | 2.683 | 0.720 |
CAP-EOM-EA-CCSD/aug-cc-pVDZ+3s3p3d(C) (1st order)b | 2.538 | 0.509 |
CAP-EOM-EA-CCSD/aug-cc-pVDZ+3s3p3d(A) (0th order)b | 2.647 | 0.919 |
CAP-EOM-EA-CCSD/aug-cc-pVDZ+3s3p3d(A) (1st order)b | 2.544 | 0.630 |
Experimentc | 2.82 | … |
Table XIII illustrates that aug-cc-pVDZ+3s3p3d(C) and aug-cc-pVDZ+3s3p3d(A) yield similar results for the resonance position. CAP-EOM-EA-CCSD overestimates the position of the 2Au resonance by about 0.7 eV regardless of the basis set used and also independent of whether the first-order correction is applied. For the 2Bg state, an overall better agreement with experiment is found (0.2-0.3 eV), but the first-order correction makes a sizable impact and moves the CAP-EOM-EA-CCSD values away from the experimental value. Note that based on the findings from Sec. IV, one should expect a significant change of all results when increasing the valence basis set.
Concerning the resonance width, our results support the experiment's hypothesis that the 2Au state is considerably longer lived than the 2Bg state. We also note that first-order results for the resonance width are smaller by 0.05 eV for the 2Au state and by 0.2-0.3 eV for the 2Bg state. The change from the “C” to the “A” basis set makes an impact of similar magnitude but in opposite direction. A final judgment of the accuracy of the results, however, cannot be made due to the lack of other theoretical or experimental estimates for the resonance width.
VI. CONCLUSIONS
A complete and robust implementation of CAPs within EOM-EE-CCSD and EOM-EA-CCSD methods has been presented together with a protocol for studying molecular shape resonances without system-dependent optimization of basis set and CAP parameters.
In our approach, we choose the onset of the CAP as the expectation value of the spatial extent of the reference-state wave function, which ensures that the reference state is minimally perturbed by the CAP (∼10−5 a.u.). We showed that resonance positions and lifetimes obtained from energies that are corrected for the CAP perturbation in first order16,24 are less sensitive (∼ 0.03 eV) to variation of the CAP onset than uncorrected zeroth-order energies. To determine the optimal CAP strength, we used the criterion from Ref. 24 for the separate stabilization of the real and imaginary part of the first-order corrected energy instead of the most widely used criterion |η dE/dη| = min based on the zeroth-order energy.
Benchmark studies for the π* resonances of CO− and
While the current paper shows the potential of CAP-EOM-CCSD approaches, it is also clear that the application to larger systems is hampered by the need to calculate η-trajectories, i.e., to recalculate the energy for different values of the CAP strength, which increases the computational cost considerably as compared to conventional EOM-CCSD calculations. To make our current implementation of CAP-EOM-CCSD faster and to increase its black-box character, a number of improvements will be pursued in future work. As the wave function changes smoothly with the CAP strength, one can expect that the use of the wave function parameters from the previous step as a guess will accelerate the calculation of η-trajectories significantly provided that sufficiently small step sizes are used. A further automatization will be possible by the implementation of analytic derivatives dE/dη as this will enable the determination of optimal CAP strengths without that the user has to specify a step size and a range, where the search is performed. Put together, these developments will enable the application of CAP-EOM-CCSD to resonance states of larger molecules, for example, biochromophores,139,140 where standard EOM-CCSD is routinely used for the characterization of bound states.
ACKNOWLEDGMENTS
This work is supported by the (U.S.) Army Research Office (USARO) through the W911NF-12-1-0543 grant (A.I.K.). We also acknowledge support from the Humboldt Research Foundation (Bessel Award to A.I.K. and Feodor Lynen fellowship to T.C.J.). M.H.G., E.E., and E.S. were supported by the Scientific Discovery through Advanced Computing (SciDAC) program funded by the (U.S.) Department of Energy (DOE), Office of Science, Advanced Scientific Computing Research and Basic Energy Sciences. Additional support for EJS and MHG was provided by the U.S. Department of Energy, Office of Basic Energy Sciences, Chemical Sciences, Geosciences and Biosciences Division, through Contract No. DEAC02-05CH11231. We thank Dr. Y. Sajeev and Professor Nimrod Moiseyev for helpful comments.