Nicola Bogo*a,
Zeyi Zhang
bc,
Martin Head-Gordonbc and
Christopher J. Stein
ad
aDepartment of Chemistry and Catalysis Research Center, TUM School of Natural Sciences, Technische Universität München, Lichtenbergstr. 4, 85748 Garching, Germany. E-mail: christopher.stein@tum.de
bPitzer Center for Theoretical Chemistry, Department of Chemistry, University of California, Berkeley, California 94720, USA
cChemical Sciences Division, Lawrence Berkeley National Laboratory, Berkeley, CA 94720, USA
dAtomistic Modeling Center, Munich Data Science Institute, Technical University of Munich, Garching 85748, Germany
First published on 25th July 2025
Ab initio quantum-chemical methods that perform well for computing the electronic ground state are not straightforwardly transferable to electronically excited states, particularly in large molecular systems. Wave function theory offers high accuracy, but is often prohibitively expensive. Methods based on time-dependent density functional theory (TD-DFT) are crucially sensitive to the chosen exchange–correlation functional (XCF) parameterization, and system-specific tuning protocols were therefore proposed to address the method's robustness. Methods based on the variational relaxation of the excited-state electron density showcased promising results for the calculation of charge-transfer excitations, but the complex shape of the electronic hypersurface makes convergence to a specific excited state much more difficult than for the ground state when standard variational techniques are applied. We address the latter aspect by providing suitable initial guesses, which we obtain by two separate constrained algorithms. Combined with the squared-gradient minimization algorithm for all-electrons relaxation in a freeze-and-release scheme (FRZ-SGM), we demonstrate that orbital-optimized density functional theory (OO-DFT) calculations can reliably converge to the charge-transfer states of interest even for large molecular systems. We test the FRZ-SGM method on a phenothiazine-anthraquinone CT excitation in a supramolecular Pd(II) coordination cage complex as a function of the cage conformation. This compound has been studied experimentally prior to our work. We compare this freeze-and-release scheme to two XCF reparameterizations, which were recently proposed as low-cost TD-DFT-based alternatives to variational methods. Two dye–semiconductor complexes, which were previously investigated in the context of photovoltaic applications, serve as a second example to investigate the convergence and stability of the FRZ-SGM approach. Our results demonstrate that FRZ-SGM provides reliable convergence for charge-transfer excited states and avoids variational collapse to lower-lying electronic states, whereas time-dependent DFT calculations with an adequate tuning procedure for the range-separation parameter provide a computationally efficient initial estimate of the corresponding energies, with a computational cost comparable to that of configuration-interaction singles (CIS) calculations.
A rational strategy for the development of carbon-based semiconductor technology implies the study of charge transfer (CT) in biological systems, its reproduction in bio-inspired energy materials, and their technological application. Charge transfer (CT) refers to the process in which an electron moves from an electron-rich donor (D) moiety to an electron-deficient acceptor (A). The donor and acceptor groups can belong to the same molecule (intramolecular CT), or be located on different molecular units (intermolecular CT). On top of that, light absorption can trigger the formation of CT excitations (photo-induced CT). The distance RDA of the nuclear center of mass of the donor and acceptor moieties is a common measure for the extent of charge separation.12 The fundamental understanding of charge-transfer processes enabled by theoretical work is fundamental, forging guiding principles for molecular design and testing them with computer simulations. Given this premise, excited-state electronic structure methods play a crucial role in the study and discovery of charge transport in molecular systems.13–15 Since their introduction in the 1990s, tools based on density functional theory (DFT) have achieved predictive power in the simulation of ground-state (GS) chemistry16 and are routinely used to develop new chemicals.17,18 Likewise, excited-state electronic-structure methods based on DFT, like time-dependent density functional theory19 (TD-DFT), gained popularity due to the straightforward simulation of light absorption features in small molecules. In a TD-DFT calculation, an approximation of the response function of the system under investigation is computed. The success of TD-DFT resides in the application of efficient approximations that lower its computational demand at the cost of introducing small errors. With a focus on economic simulations, the most popular TD-DFT methods expand the response function in a Taylor series and truncate it to the linear term, applying the Tamm–Dancoff approximation20 (TDA) to recast the resulting electronic structure problem to the well-known configuration-interaction singles21,22 (CIS) method. This method, which we name TDA from here on, is plagued by systematic errors23,24 for common (hybrid) exchange–correlation functionals (XCFs) with fixed parameters. This error can be reduced by system-specific tuning protocols or with functionals that are higher up in the Jacob's ladder classification, such as double-hybrid functionals. When a range-separated hybrid XCF is employed, the range separation parameter ω can be iteratively adjusted such that the HOMO and LUMO energies match the ionization potential (IP) and electron affinity (EA), respectively.25 In double-hybrid XCFs, such as e.g. B2PLYP,26 parametrization schemes for the range-separation parameter also exist to tune the functional to match the IP and EA. Yet another alternative is the local variation of the fraction of exact exchange according to a local mixing function in so-called local hybrid XCFs (e.g. LH20t27), which can be implemented efficiently,28 but leads to a non-trivial parametrization problem. All of these options can significantly reduce the error in TDA calculations with the tuning of range-separated hybrid functionals standing out as an efficient and straightforward procedure.
In our previous publication,24 we demonstrated how orbital-optimized DFT (OO-DFT) methods constitute the most accurate low-scaling electronic structure approximation for the calculation of inter-molecular CT excitations and do not require special adjustments of the range-separation parameter. OO-DFT methods compute the ground state and an excited state variationally in two separate self-consistent field (SCF) calculations. For this reason, this approach is also referred to as ΔSCF for excited states, and its theoretical foundations were proven under several formulations.29–31 In contrast to the ground state, excited-state energies belong to saddle points of the electronic hypersurface, making numerical convergence more challenging. Indeed, we relied on outlier detection rules in our previous publication24 to exclude data points flawed by poor convergence in our benchmark study, where we subsequently tested the accuracy of OO-DFT methods. We introduced a dataset of accurate intermolecular charge-transfer (ICT) excitation energies24 in donor–acceptor molecular dimers at various separation distances (RDA) and found that OO-DFT methods perform well, but struggle to converge the electronic structure at short RDA (3.5–5 Å) because of strong changes in the polarization of the excited state MOs. Following our study, Schmerwitz et al.32 analyzed the performance their freeze-and-release optimization algorithm on our RDA-dataset. The freeze-and-release algorithm is a 2-step optimization algorithm, which combines an iterative constrained (“frozen”) optimization step with a subsequent relaxation of all degrees of freedom. The authors were able to obtain a smooth potential energy surface for the ICT excitation also at short donor–acceptor distances, demonstrating the importance of a well-behaved guess for converging OO-DFT calculations. In this contribution, we introduce a new strategy for the refinement of the MO guess and combine it to the OO-DFT algorithms implemented in the Q-Chem software package,33 testing it on the ICT excitations in the tetrafluoroethylene-ethylene dimer system from the RDA-dataset, that we consider a good benchmark for convergence of OO-DFT methods. We implement and analyze a constrained optimization algorithm to refine the MO guess used in OO-DFT calculations using the efficient geometric direct minimization (GDM) algorithm,34 akin to the approach introduced by Schmerwitz et al.32 Moreover, we investigate an alternative guess for ICT excitations obtained from the absolutely-localized molecular orbitals35,36 (ALMO) method and compare the maximum overlap method to an initial guess37 (IMOM) to the squared-gradient minimization38,39 (SGM) algorithm on these refined guesses.
TDA is certainly of fundamental importance for the calculation of light absorption properties, and is by far the most popular method for the calculation of electronic excitations in molecular systems. The OO-DFT methods we apply in this work also require an initial TDA calculation to set up the OO-DFT calculation. Nevertheless, results obtained with the TDA method depend significantly on the chosen XCF approximation and should therefore be validated for a given compound class by comparing them to reference data. On top of that, CT excitations in organic systems tend to have a very small transition dipole moment (TDM) and therefore a small absorption coefficient compared to the bright transitions that dominate the spectrum. Theory-to-theory validation is therefore the only suitable strategy to validate computational methods for CT excitations,24,40 but a direct assessment is often prohibited by the steep computational scaling of accurate wavefunction methods. For this reason, we demonstrated the accuracy that can be achieved with OO-DFT methods in an extensive benchmark for small systems in our previous work,24 and we confront technical challenges regarding the stable convergence towards the desired CT target states for large molecular systems in the present study. That the accuracy that we demonstrated for small systems does not deteriorate for the prototypical large systems studied in this work is a conjecture we make based on experience with ground-state DFT.
In Section 2, we review the current state-of-the-art methods for variational calculations of electronically excited states. Our implementation of an efficient partitioning of the MO matrix C is discussed in Section 3.2. Convergence and accuracy of the SGM and IMOM methods using the ground-state MO guess, ALMO guess, and refined MO guess through constrained optimization with frozen excited electron and hole MOs (Frozen guess or “FRZ guess” in the following) are tested on the tetrafluoroethylene-ethylene dimer in Section 4.1. The guess refinement algorithm is combined with the SGM algorithm for the calculation of CT excitations in large systems. We test our method on CT excitations on two classes of supramolecular systems: a Pd(II) coordination cage and two dye–semiconductor model materials. The supramolecular cage we studied was previously investigated by Frank et al., who combined spectro-electrochemistry and time-resolved spectroscopy experiments to investigate the photoinduced relaxation upon excitation.10 Upon dissolution of the cage components in a reaction medium, intercalated double-cage complexes were formed by self-assembly. Two banana-shaped ligands, one bearing a phenothiazine (PTZ) dye and one an anthraquinone (ANQ) dye, were employed for the preparation of several intercalated cage dimers. Each cage dimer is composed of four Pd(II) metal centers and eight organic ligands that coordinate the metal with a pyridyl group. The cage dimers observed were either homoleptic, solely composed of the PTZ and ANQ ligands, respectively [Pd4(PTZ)8]8+ and [Pd4(ANQ)8]8+ or heteroleptic with a statistical mixture of PTZ and ANQ ligands ([Pd4(ANQ)n(PTZ)8−n]8+). The homoleptic intercalated cage dimers proved stable towards repetitive oxidation/reduction cycles, and the changes in the visible light absorption spectrum after reduction and oxidation steps were investigated. Transient absorption (TA) of heteroleptic [Pd4(ANQ)n(PTZ)8−n]8+ cage upon 385 nm excitation was also recorded. 100 ps after the excitation pulse, the TA shows similar features to the spectral changes obtained by oxidizing and reducing the homoleptic donor and acceptor complexes, respectively, giving experimental evidence of a charge-transfer excitation formed in the supramolecular structure.
The results obtained with our OO-DFT methods on the PTZ-ANQ CT excitation in a Pd(II) coordination cage are compared to the TDA prediction obtained with recently proposed reparametrizations of the LRC-ωPBE41 XCF for the TDA method and experimental measurements in Section 4.2. Our calculations on the stacked conformation of a heteroleptic cage monomer confirm a PTZ-ANQ CT excitation, and we compared its energy to the PTZ-Pd(II) ligand-to-metal CT (LMCT) state. However, a direct comparison of the computed energy to the experimental recordings is not possible, since the formation of a CT state 100 ps after light excitation is proven experimentally, but the energy of such a state cannot be measured in TA experiments. Excited-state dynamic simulations and excited-state absorption would be required to unambiguously prove the formation of the CT excitation, but no methodology is available to date for systems of this size. We consider OO-DFT methods excellent candidates for dynamic simulations of large systems, and ongoing work in our lab is aimed at making such calculations feasible. Finally, CT excitations of two dye–TiO2 complexes were investigated in Section 4.3. These systems are chemically rather different than the coordination cages previously investigated, because the electron acceptor is a semiconductor material. We were therefore interested in testing the performance of our method also for such a system. The dyes we investigate—which we refer to as JK2 and D102 for simplicity—are highly relevant from an application point of view as a potential material in solar cells, and have been studied both experimentally42–45 and in computer simulations.11,46,47 However, CT excitations typically form weak bands compared to local (valence) states, which dominate the UV-Vis spectrum. Therefore, we are unable to assign contributions to the experimental absorption spectrum to dye–TiO2 CT excitations and we do not expect direct formation of the CT state by light absorption. Although we cannot directly compare our calculations to experimental results, we compared to other theoretical studies and demonstrate the stable convergence of our method also for this chemically distinct class of large heterogeneous systems.
C(i+1) = C(i)U. | (1) |
U = eΔ. | (2) |
Conversely, extrapolation methods, such as the direct inversion of the iterative subspace (DIIS) method from Pulay,49,50 leverage the mathematical properties of the Fock F(i) and density P(i) matrices at convergence to construct an error vector. The commutator between density and Fock matrix vanishes at convergence, such that
e(i) = SP(i)F(i) − F(i)P(i)S | (3) |
To mitigate the complexity of direct saddle-point optimization of the electronic hypersurface, a constraint is applied to the density in constrained DFT51 (CDFT) methods. Constraints can impose site occupation52 conditions or can be applied directly to the density difference between ground and excited state.53 These methods hold great potential when enough information is available about the target excited state (e.g. information about site occupancy), but in this work, we use constraints to compute an improved guess density. In our calculations, the electronic energy is hence computed by optimization of all degrees of freedom of the electronic energy without applying constraints.
To compute specific excited states, Gilbert et al. proposed an adaptation of direct optimization or extrapolation algorithms,54 where the Fock matrix is constructed from the orbitals that overlap maximally with a reference configuration rather than following the Aufbau principle. In this maximum overlap method (MOM), the overlap of the occupied MO set belonging to the i-th iteration to a reference MO set Cref
O = (Cref)†SC(i) | (4) |
![]() | (5) |
Hait and Head-Gordon proposed an alternative approach exploiting DO algorithms,38,39 taking the square of the energy derivative with respect to occupied-virtual orbital rotations
![]() | (6) |
Finally, Selenius et al. proposed an efficient diagonal preconditioner for DO algorithms40
![]() | (7) |
In the end, ΔSCF methods are based on local optimization algorithms, and the underlying problem with variational calculations of excited-state energies resides in the complexity of the energy hypersurface. Rapid convergence to the desired stationary point hence requires a suitable initial guess. Therefore, guess-refinement algorithms are desirable to reduce the number of iterations and ensure convergence to the correct electronic configuration in OO-DFT calculations. Schmerwitz et al. made use of constrained optimization in their variational excited-state methods32,57,58 implemented in the GPAW software package59 with broad success and proved the reliability of their approach by recomputing the outlier data points from our RDA-dataset. This motivated us to implement our variant of a constrained optimization algorithm and test the performance of the IMOM and SGM algorithms on the refined guess. Details about the constrained optimization algorithm are discussed in Section 3.2.
An alternative method for the calculation of inter-molecular excitations is the absolutely-localized molecular orbitals (ALMO) method. ALMO enables the variational calculation of molecular fragments by partitioning the Gaussian basis {φμ} spanning the rows of the C matrix into localized subsets. In the simple case of two fragments, named A and B, the Gaussian basis is partitioned into the localized bases {φν}A and {φκ}B belonging to fragments A and B respectively. The MO matrix C is therefore enforced to maintain the block structure
![]() | (8) |
Finally, in Section 4.2 we compare OO-DFT methods to reparametrizations of the LRC-ωPBE41 XCF in TD-DFT. A system-specific tuning of the range-separation parameter ω proved effective in combination with TD-DFT by recovering properties of the exact XCF with semi-empirical RSH XCFs.60 Two alternative schemes to determine a suitable ω in LRC-ωPBE have recently been proposed. The global density-dependent (GDD) tuning61,62 adjusts ω by computing the distance dx between an electron in the outer regions of a molecule and the exchange hole in the localized valence orbitals
ωGDD = C〈dx2〉−1/2, | (9) |
Yan et al.63 proposed an alternative empirical XCF reparametrization designed for the description of CT states. From a preliminary TD-DFT calculation using a related XCF (e.g. the PBE064 XCF is used for tuning LRC-ωPBE41), and the average excitation distance DCT is computed for several CT excited states, respectively. The ω parameter for the final calculations is then obtained from the maximum DCT in the set
![]() | (10) |
We start from a partitioning of the orbital coefficient matrix C of a restricted Hartree–Fock (RHF) calculation but note that this approach can be extended straightforwardly to the unrestricted case and Kohn–Sham theory. All C elements are real numbers and C is ordered according to the block structure shown in Fig. 1.
Given an atomic orbital (AO) basis of size Nbasis, the C matrix of dimension NMO(≤Nbasis) is ordered into blocks. Active doubly-occupied vectors are in the first block D, indexed from 1 to ND. A block Df with NDf frozen doubly-occupied orbitals follows, indexed from ND+ 1 to . Similarly, virtual orbitals are partitioned into the V block containing active virtual orbitals
and a block Vf containing frozen virtual orbitals
. Note that for ICT excitations the Df and Nf blocks are simple column vectors. At the beginning of the calculation, occupied (red) and virtual (green) orbitals are reordered to move the frozen vectors (in grey) to the end of each block. In the simple example shown in panel A of Fig. 1, a double excitation is computed by exciting one α and one β electron from MO no. 2 to MO no. 4, and the MOs are sorted in the order 1-4-3-5-6-2. The Df and Vf blocks contain only MOs no. 4 and 2, respectively. The block structure emerging for the C matrix by partitioning into active and frozen parts is depicted in panel B of Fig. 1, using the general notation introduced above. Once the matrix is reordered to have the MOs involved in the excitation at the end of the occupied and virtual block, the constrained optimization just requires input on the number of holes and electrons to constrain, and in principle, multiple electron–hole pairs can be considered.
The energy optimization is then performed with a second-order method, which requires gradients and previous steps to compute an approximation of the inverse Hessian. In our implementation, we leverage the efficient GDM algorithm34 implemented in Q-Chem. The C matrix reordering results in the partitioning of the Fock matrix depicted in Fig. 2. The Fock matrix is computed in the MO basis. The energy gradient with respect to the MO rotations is given by the off-diagonal elements of the symmetric Fock matrix. In the presence of frozen blocks Df and Vf, the Fock matrix is partitioned into 16 blocks, six belonging to the upper triangle. Only the DV block in the upper triangle (highlighted in blue in Fig. 2) enters the gradient, as frozen vectors do not contribute in this constrained optimization. Similarly, gradients and steps from previous iterations are stored only for the free active vectors and are transformed by taking the corresponding blocks from the transformation matrix given the step Δ.
![]() | ||
Fig. 2 F matrix block structure arising from the C matrix partitioning introduced in Fig. 1. The Fock matrix has 16 blocks, six in the upper triangle. Frozen blocks do not contribute to the gradient, such that only the DV block from the upper triangle (highlighted in blue) contributes to the gradient. |
We evaluate the performance of this refined guess strategy (termed FR) combined with the IMOM and SGM algorithms on the lowest-lying ICT excitation of the tetrafluoroethylene-ethylene dimer at various donor–acceptor displacements in the following Section 4.1.
![]() | ||
Fig. 3 Lowest-lying ICT excitation in the tetrafluoroethylene-ethylene dimer as a function of donor–acceptor distance RDA. The convergence of several OO-DFT methods and initial guess MOs is analyzed in panel A, and the accuracy of the FRZ-SGM method is compared against TDA and the EOM-CCSD(fT)71/cc-pVTZ72 reference in panel C. The density-based CT descriptor DCT is plotted in panel B, alongside a sketch of the system investigated in the inset. The markers show results for IMOM calculations computed with the ground-state MO guess orbitals (grey x markers), ALMO guess orbitals (pink triangles), and frozen hole-and-electron guess orbitals (blue dots). The dashed orange lines correspond to the energy and descriptors computed on the ALMO density, and the dashed green lines belong to the frozen hole-and-electron density. The purple line is obtained by using the FRZ guess orbitals to initiate the SGM algorithm with line search (FRZ-SGM). In panel C, the red line is the TDA result, and the reference (grey line) is computed with EOM-CCSD(fT) and the cc-pVTZ basis. |
The energies of the lowest ICT states as a function of the donor–acceptor distance are shown in panel A of Fig. 3. The orange and green dotted lines result from calculations where the ALMO and FRZ constraints are sustained, respectively, whereas the purple line corresponds to a subsequent variational optimization with SGM (FRZ-SGM). We note that starting SGM from the ALMO guess yields identical results in this example. For large donor–acceptor distances, all methods presented here reliably converge to an ICT state that shows the expected 1/RDA behaviour. Evidently, even at short distances, the ALMO and FRZ constraints provide excellent guesses that are only minimally optimized during the subsequent unconstrained variational optimization. At long donor–acceptor distances, this discrepancy is invisible on the scale of the graph. The CT descriptor in panel B confirms convergence to an ICT state.
Different results are obtained when these initial guesses are combined with a subsequent IMOM optimization. Starting from a ground-state (gray x-markers), FRZ guess (blue dots), or ALMO (pink triangles), panel A of Fig. 3 shows that for short donor–acceptor distances, IMOM optimization converges to states that are significantly lower in energy than the initial guesses or the SGM results described above. Predictably, the discrepancy is worst on average for the ground state guess, but also the refined guess strategies cannot guarantee convergence to the targeted ICT state. Indeed, and in line with our previous work,24 panel B demonstrates that these calculations converged to lower-lying non-ICT states, because DCT quantifies how far the electron density is shifted upon excitation, and its value is <2 Å.
We hence conclude that ALMO and FRZ provide excellent initial guesses for ICT states, and SGM provides a reliable convergence to the targeted states once the constraints are lifted in a variational optimization. We further compare the OO-DFT results to TD-DFT calculations (red line) and an accurate equations-of-motion coupled-cluster (EOM-CC) reference calculation (gray line) in panel C of Fig. 3. The reference densities (panel B in Fig. 3) were computed with EOM-CCSD, and the energies (panel C) were adjusted by perturbative triples correction with the EOM-CCSD(fT) method. All DFT methods are red-shifted by an offset, but as Table 1 shows, the OO-DFT methods yield a much smaller mean error (0.64 eV) than TDA (1.27 eV). The improvement achieved by the OO-DFT method is attributed to the accuracy of the range-separated hybrid XCF, combined with an algorithm that ensures convergence to the targeted state. The approximation of weakly interacting fragments introduced in the ALMO method is invalid for short donor–acceptor separations, such that the stabilization achieved upon variational optimization for this situation with significant orbital overlap between the fragments is not captured. The statistical errors for the constrained ALMO results can hence look better than the results of the OO-DFT methods (here: FRZ-SGM) but the former does not correspond to saddle points on the electronic hypersurface and are therefore not valid excited states. The TDA method gives the largest mean error and mean variance in Table 1. We observe, however, a very small variance for FRZ-SGM, meaning that the energies deviate from the reference energies by a rather constant shift. A systematic offset can be beneficial compared to a statistical error in the prediction of deactivation paths if all excited states are affected by the same shift, resulting in fortuitous error compensation. Interestingly, the FRZ energy (green dotted line) lies lower than the target stationary points at short distances, indicating that the saddle-point search must proceed uphill along the electron–hole MO rotation, which is possible since we optimize the squared gradient. Furthermore, the DCT descriptor (panel B) shows how FRZ-SGM is the only method correctly describing the electron–hole interaction at 3.5 Å separation, deviating from the linear trend just as the reference method.
Method | Mean error (eV) | Mean variance (eV) |
---|---|---|
TDA | −1.27 | 0.043 |
ALMO | −0.53 | 0.031 |
FRZ guess | −0.77 | 0.062 |
FRZ-SGM | −0.64 | 0.005 |
We conclude that when started with a suitable initial guess, the SGM method converges reliably to the desired stationary point on the excited-state surface. In combination with a suitable range-separated hybrid XCF, OO-DFT methods yield a small mean error and variance with respect to highly accurate reference calculations.
Fig. 5 shows the TDA results obtained with the PBE064 XCF for the lowest-lying ICT excitation of the D–A dimer in the conformations of the PTZ-ANQ cage, as a function of the D–A dihedral angle (and corresponding distance) discussed above. We define the donor–acceptor distance RDA as the distance between nuclei of the dye centers highlighted in panel A of Fig. 4. The DCT magnitude is color-coded, and the DCT is additionally shown separately in the inset. The excitation energy is fitted with the asymptotic expression
![]() | (11) |
DCT = kRDA + q, | (12) |
Fig. 6 shows the results of the dihedral scan for the remaining electronic-structure methods in Table 2. Panel A is obtained by applying the GDD reparametrization scheme, tuning ωGDD for each conformation in the scan. The resulting curve is smooth, with an asymptote that is blue-shifted by more than 2 eV compared to the non-range-separated XCF. The data point corresponding to 56.3 dihedral is excluded from the fit, since the calculated DCT value strongly deviates from the expected linear trend.
Results obtained from the empirical parametrization63 suggested by Yan et al. are shown in panel B of Fig. 6. The 40 lowest-lying singlet excited states are computed with the PBE0 XCF (results reported in Fig. 5), and the state with the maximum DCT is used to compute ω*. The resulting curve is steeper than for the previous results and results in the largest IPD − EAA and b obtained for all methods. The data points corresponding to the stacked conformer were removed from the fit, but the obtained energies scatter slightly around the fitted trend. To better understand the reason for the smoother curve obtained with the ωGDD-tuning procedure as compared to the tuning based on the maximum DCT, we report the respective range-separation parameters in Table 3. The negligible variation of ωGDD along the scan explains the smooth and monotonous curve obtained with this tuning procedure. On the contrary, major changes of DmaxCT result in strong variations of ω* for neighboring data points in the scan. Notably, for large donor–acceptor separations, the results are almost identical. Future work might reveal if empirical tuning based on the DCT descriptor in the asymptotic limit can generally improve the results. In conclusion, for systems where conformational changes can significantly affect the energy of CT excitations, size-consistent reparametrizations like ωGDD are recommended for accurate TD-DFT energetics of ICT states.
D–A dihedral θ (°) | ωGDD (Å−1) | ω* = 2/DmaxCT (Å−1) | DmaxCT (Å) |
---|---|---|---|
13.5 | 0.242 | 0.467 | 4.28 |
31.3 | 0.240 | 0.152 | 13.12 |
36.3 | 0.240 | 0.153 | 12.82 |
41.3 | 0.242 | 0.142 | 14.12 |
46.3 | 0.239 | 0.188 | 10.63 |
51.3 | 0.237 | 0.187 | 10.72 |
56.3 | 0.235 | 0.229 | 8.74 |
61.3 | 0.234 | 0.237 | 8.45 |
66.8 | 0.233 | 0.215 | 9.31 |
Finally, we also investigated the performance of the ALMO-SGM and FRZ-SGM OO-DFT methods. Tight convergence of the SGM algorithm can be challenging, because converging the gradient to a sufficient degree requires very tight convergence of the squared gradient. In our calculations, we used a rather loose convergence criterion of 10−4, compared to the threshold of 10−5 used for single-point calculations. The results are displayed in panels C and D of Fig. 6 for the ALMO-SGM and FRZ-SGM methods, respectively. The guess energy and DCT descriptors are shown as red circles, whereas the fully converged results are displayed as solid dots. The ALMO guess energy increases when bringing the D and A monomers closer, in contrast to the expected asymptotic trend. Clearly, however, the ALMO density is a suitable guess for further optimization with the SGM method. This is demonstrated by the inset of Fig. 6 panel C, where the results obtained from the guess density are close to the converged results and by the fact that the SGM optimization relaxed smoothly all data points, stabilizing some by several eVs (note the larger energy axis range as compared to the other panels). Conversely, the constrained optimization (FRZ method, Fig. 6 panel D) produces both energies and densities very close to the target stationary state. The asymptotic trend is well reproduced in the dihedral scan for both OO-DFT methods, with the sole exception of the data point corresponding to a dihedral angle of 31.3°, which was found to be problematic also for other methods. Evidently, the density obtained by the constrained FRZ optimization is already very close to the target state, meaning most of the density relaxation involves MOs that are not directly involved in the excitation, but are influenced by the charge separation in the CT state through electrostatic interaction. The ALMO-SGM and FRZ-SGM methods lead to almost identical values for IPD − EAA in the fit (rows 4 and 5 in Table 2), within 0.5 eV of the results obtained with the LRC-ωGDD PBE reparametrization. As a conclusion, we recommend the FRZ-SGM OO-DFT method for the calculation of ICT excitations in large D–A dimers, and the LRC-ωGDDPBE reparametrization for TD-DFT calculations.
We then applied the FRZ-SGM method to the calculation of D–A CT excitations in the full PTZ-ANQ Pd coordination cage. The Pd coordination environment is conserved in the structures interpolated as described above and we computed the CT excitation energies for nine structures of the D–A dihedral scan. The D–A dihedral in Fig. 7 is obtained from planes fitted only through the dye cores (secondary x-axis). By analogy, the RDA separation (primary x-axis) is computed as the distance of the center of nuclear charge of the D and A dye cores. The resulting dihedral spans a wider range compared to ligand-only study in Fig. 6, from −9.6° in the stacked conformer to 77.6° in the orthogonal conformer. The FRZ-SGM method converged reliably to the lowest-lying CT excitation involving the D and A dyes (circles). For the full cage, we decided to fit the calculated ICT energies to a shifted expression
![]() | (13) |
![]() | ||
Fig. 7 D–A CT excitations calculated with FRZ-SGM for the full cage structure as a function of the donor–acceptor distance. The density difference between the ground and ICT excited state is shown at the bottom of each panel as an isosurface plot (isovalue 0.0005 for the blue surface and −0.0005 for the red surface). The energies are fitted with a shifted asymptote (eqn (13)). The D–A CT excitations are marked with dots, and additionally calculated a ligand-to-metal CT excitation are marked with a red-framed star. Inset: DCT values plotted against the donor–acceptor separation distance RDA and fitted with a linear expression. |
We also computed “pure” LMCT excitations (red star marker in Fig. 7) for the stacked and orthogonal conformers, which we unambiguously identified based on a visualization of the difference density. Keeping in mind all uncertainties of our method, our calculations predict LMCT excitations in the same energy range as the D–A CT excitations, both in stacked and orthogonal conformations. Hence, we cannot rule out excited-state relaxation to the LMCT excitations following the experimentally applied 385 nm pulse.
Finally, we estimate the likelihood of inter-cage CT within the interlocked cage dimer. Starting from the crystallographic data reported in ref. 10, we isolated an ANQ-ANQ dimer in a stacked conformation. We then swapped one ANQ ligand for a PTZ ligand, aligning the N atoms to the Pd coordination sites in a π-stacked conformation. The process we followed to generate this stacked conformation as an approximation to the interlocked cage structures is summarized in Fig. 8. The ALMO-SGM calculation shows the presence of a charge-transfer excitation at 3.81 eV, confirming the presence of inter-cage excitations in the energy range of intra-cage CT excitations. However, to confirm the assignment of the PTZ-ANQ excitation to an inter-cage CT would require the computationally prohibitively expensive calculation of the full interlocked cage dimer, since in our model calculation, this ICT state would not be accessible with a 385 nm pulse. In conclusion, we confirm the presence of low-lying D–A CT and LMCT excitations in the supramolecular Pd coordination cage compounds. We find inter-cage CT excitations in an energy range compatible with the cage conformations, but cannot exclude the formation of LMCT excitations in the excited-state deactivation process.
![]() | ||
Fig. 8 Starting from the crystal structure in ref. 10 an ANQ molecular dimer is isolated with monomers coming from two different cages of the interlocked structure. One of the ANQ ligands (acceptor) is then swapped for a PTZ (donor) ligand, preserving the local N-Pd coordination environment. A calculation of the PTZ-ANQ ICT excitation in this conformation is then performed with the ALMO-SGM method. The calculated excitation energy of 3.81 eV, is in the same energy range predicted for CT excitations in the single cage-derived donor–acceptor conformations (see Fig. 7). The predicted ICT excitation energy lies above the energy of the pump pulse used in ref. 10. |
Dye | sTDA | FRZ-SGM | DCT | Class |
---|---|---|---|---|
D102 | 3.48 | 5.45 | 3.21 | Dye–TiO2 |
JK2 | 2.84 | 2.56 | 6.01 | Dye–spacer |
JK2 | 4.03 | 3.17 | 5.90 | Dye–spacer |
JK2 | 4.61 | 3.50 | 7.45 | Dye–TiO2 |
JK2 | 4.69 | 3.57 | 7.41 | Dye–TiO2 |
JK2 | 4.65 | 3.23 | 6.59 | Dye–TiO2 |
JK2 | 4.75 | 3.76 | 7.48 | Dye–TiO2 |
JK2 | 4.80 | 3.82 | 7.52 | Dye–TiO2 |
JK2 | 4.95 | 3.72 | 7.33 | Dye–TiO2 |
JK2 | 4.99 | 3.91 | 7.61 | Dye–TiO2 |
One direct dye–TiO2 excitation was calculated with the FRZ-SGM method for the D102-TiO2 complex, which appears strongly blue-shifted compared to the sTDA result. On the contrary, all excitations computed with FRZ-SGM in the JK2-TiO2 complex are slightly red-shifted in comparison to the sTDA result. The JK2 ligand is characterized by a long aromatic “spacer” connecting the dye to the semiconductor. By inspecting the density differences, we classified all CT excitations computed in Table 4 as “dye–spacer” or “dye–TiO2”. The JK2-semiconductor complex shows two groups of CT states: dye–spacer excitations reside in the visible spectrum, and dye–TiO2 excitations, resulting in the direct injection of a dye electron in the semiconductor band, populate the near-UV spectral region. Fig. 9 shows prototypical density difference plots for both groups of CT states.
The promising FRZ-SGM method was used for the calculation of CT excitations in three large supramolecular systems. Our calculations of PTZ-ANQ excitations in the full coordination cage investigated experimentally by Frank et al.10 predicted excitation energies in the visible range for the stacked conformation, as well as LMCT excitations. To confirm the true character of the experimentally observed excitations, further calculations of the intercalated double-cage dimer are required but are computationally very demanding even for the scalable methods identified in this work. For a quantitative comparison with experiment, adequate treatment of the dynamics and solvation effects is required, but beyond the scope of this work. Nevertheless, our calculated 2.9–3.5 eV window is consistent with the excitation energy pulse of 3.2 eV in the study from Frank et al., thereby providing qualitative validation. Finally, CT excitations in two covalently bonded dye–semiconductor complexes, D102-TiO2 and JK2-TiO2,11 were computed with the sTDA and FRZ-SGM methods. FRZ-SGM reliably converged on the CT excitations also for these systems.
We demonstrate that, when combined with a robust guess refinement strategy, OO-DFT methods are an adequate electronic-structure method for the study of photoinduced phenomena in large supramolecular systems. Being grounded in a rigorous theoretical framework, this method holds the promise to allow for an accuracy that is comparable to ground-state DFT with adequate functional development. With the stable convergence to a desired target state that we demonstrate here, OO-DFT is poised to be the method of choice for the study of excited states in large molecular systems. Necessary extensions for a more direct comparison with experiment, such as the implementation of non-adiabatic coupling vectors and combinations with QM/MM and solvation models, are currently underway in our groups. These developments will lead to an OO-DFT framework that allows for the simulation of excited-state dynamics and a study of photo-deexcitation processes for such large molecular systems that are currently out of reach. TD-DFT calculations in the Tamm–Dancoff approximation proved to be accurate enough for qualitative initial investigations or to set up subsequent OO-DFT calculations if the system-specific optimization of the range-separation parameter followed the global density-dependent tuning procedure. Further development is desirable for the convergence acceleration of the SGM algorithm.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d5cp01867f |
This journal is © the Owner Societies 2025 |