Abstract

Nuclear magnetic resonance (NMR) spectroscopy of paramagnetic molecules provides detailed information about their molecular and electron-spin structure. The paramagnetic NMR spectrum is a very rich source of information about the hyperfine interaction between the atomic nuclei and the unpaired electron density. The Fermi-contact contribution to ligand hyperfine NMR shifts is particularly informative about the nature of the metal–ligand bonding and the structural arrangements of the ligands coordinated to the metal center. In this account, we provide a detailed experimental and theoretical NMR study of compounds of Cr(III) and Cu(II) coordinated with substituted acetylacetonate (acac) ligands in the solid state. For the first time, we report the experimental observation of extremely paramagnetically deshielded 13C NMR resonances for these compounds in the range of 900–1200 ppm. We demonstrate an excellent agreement between the experimental NMR shifts and those calculated using relativistic density-functional theory. Crystal packing is shown to significantly influence the NMR shifts in the solid state, as demonstrated by theoretical calculations of various supramolecular clusters. The resonances are assigned to individual atoms in octahedral Cr(acac)3 and square-planar Cu(acac)2 compounds and interpreted by different electron configurations and magnetizations at the central metal atoms resulting in different spin delocalizations and polarizations of the ligand atoms. Further, effects of substituents on the 13C NMR resonance of the ipso carbon atom reaching almost 700 ppm for Cr(acac)3 compounds are interpreted based on the analysis of Fermi-contact hyperfine contributions.
Short abstract
The ligand NMR shifts in paramagnetic acetylacetonato Cr(III) and Cu(II) complexes have been predicted and measured in the solid state and interpreted by relativistic DFT calculations. The effects of the metal atom, ligand, and crystal packing on the spin delocalization and polarization reflected in the Fermi-contact contribution to the hyperfine interaction are rationalized.
1. Introduction
Nuclear magnetic resonance (NMR) spectroscopy of paramagnetic molecules1,2 has been shown to be an excellent technique for investigating the distribution of spin density3 and the metal–ligand bonding4 in open-shell systems. The resonance frequencies of paramagnetic species typically lie outside the normal range of NMR shifts for their diamagnetic analogs as a result of the additional contribution of hyperfine (de)shielding, which originates in the hyperfine interaction between the nuclear magnetic moment of a ligand atom L and unpaired electron(s). Because the magnetization of unpaired electron(s) is temperature dependent, its transcription to the hyperfine (HF) NMR shift, δLHF, is also temperature dependent. Therefore, measurements at various temperatures are frequently used to estimate the temperature-dependent contribution to the NMR shift and to determine the distribution of spin density.1,2 However, to interpret the experimental NMR observations in detail, theoretical analysis based on first-principles calculations is typically required.5 The total NMR shift of atom L can be calculated as the sum of the orbital (δL, approximately temperature independent) and hyperfine (δLHF, temperature dependent) contributions6,7
| 1 |
The hyperfine NMR shift can be related directly to the parameters of the EPR spin Hamiltonian, as shown for doublet systems by Moon and Patchkovskii6 and later extended to systems with an arbitrary spin degeneracy by Van den Heuvel and Soncini8,9 and others.10−15 In the case of vanishing or negligible zero-field splitting (ZFS), and when the degenerate ground state with 2S + 1 multiplicity is well separated from any excited states, the temperature-dependent part of the isotropic NMR shift obeys the Curie law and is calculated as
| 2 |
Here, we have adopted SI units, kT represents the thermal energy, ℏ is the reduced Planck constant, μe is the Bohr magneton, γL is the gyromagnetic ratio of nucleus L, and the EPR parameters g and A are the g-tensor and the hyperfine coupling tensor (A-tensor), respectively.
Using the EPR parameters, the δLHF can be separated into the Fermi-contact (FC),5,16 spin-dipolar (SD),16,17 and paramagnetic spin–orbit (PSO)16,18 terms used in eq 3, which is governed by the physical terms of the A-tensor.16
| 3 |
The Fermi-contact term, δLFC, can be linked directly to the spin density at the position of the nucleus L being probed and will be shown to be dominant for the 13C NMR shifts of the ligand atoms of Cr(III) and Cu(II) compounds 1a–1d and 2a–2d (for structures, see Figure 1) investigated in this work.
Figure 1.

Structures and atom numbering schemes for compounds 1a–1e (chromium) and 2a–2e (copper).
We recently provided a clear theoretical interpretation of the signs of the electronic g-shifts and hyperfine coupling constants of eq 2 for iridium complexes with a straightforward generalization covering the full d-block.18 Early transition-metal (TM) complexes with low d-electron occupancy (d1–d3) exhibit negative g-shifts, whereas late transition-metal complexes (d7–d9) typically have positive g-shifts. The sign of the isotropic g-shift for the Cr(III) and Cu(II) compounds being investigated (Figure 1) can also be described by a linear spin–orbit approach to the g-tensor in the framework of crystal-field theory.19,20 Recently, the relationship between the electron configuration and the sign of the hyperfine NMR shifts of the ligand has been demonstrated for a series of acetylacetonato vanadium, nickel, and copper complexes.21 Previous EPR and paramagnetic NMR studies of transition-metal complexes of acetylacetonate include those of Cr(III),22,23 Mn(III),23 Fe(III),22,23 Ru(III),22−24 Ni(II),25,26 and Cu(II)23 compounds.
In this study, we focus on representatives of early (chromium, group 6) and late (copper, group 11) 3d transition-metal compounds with various substituents, as shown in Figure 1. Their previous characterizations by NMR spectroscopy relied mostly on 1H NMR experiments, with 13C NMR data, in most cases, missing or incomplete. Here, we report a full set of 1H and 13C NMR shifts for the parental compounds 1a and 2a determined experimentally in the solid state using magic-angle spinning (MAS) experiments with a short echo delay. In addition, relativistic density-functional theory (DFT) calculations27 were performed to assist in assigning the resonances and to interpret the spin distribution in these compounds along with the Fermi-contact contributions to the hyperfine NMR shifts that are related to it.
2. Results and Discussion
2.1. Solid-State 13C NMR Shifts
The Cr(III) and Cu(II) compounds were prepared as summarized in the Supporting Information. Their NMR spectra were recorded in the solid state under magic-angle spinning (MAS) at 30 kHz. The 13C MAS NMR spectra of parent compounds 1a and 2a recorded at 313 K are shown in Figure 2 as examples.
Figure 2.
13C MAS NMR spectra of compounds (a) 1a and (b) 2a obtained in the solid state (30 kHz spinning) at 313 K. The NMR lines assigned to atoms C1, C2, and C3 are colored in red, green, and blue, respectively. The hash marks (#) denote spinning sidebands, which are out of phase due to the mismatch of the echo delay and the rotation period. The asterisk (*) denotes an additional signal in the sample of 1a. The NMR spectra (average of 105–106 scans) were recorded by a spin-echo pulse sequence with a short delay (10 μs) between the excitation and refocusing pulses and a relaxation delay of 25 ms. For more experimental details, see Section 4 and the Supporting Information (Table S1, Figures S1 and S2).
In contrast to previous reports, we succeeded in measuring all of the 13C NMR resonances22 of compound 1a and a very broad resonance21 at around 900 ppm of 2a. The measured 1H (solution) and 13C (solid state) NMR shifts for compounds 1a–d and 2a–d are summarized in Table 1.
Table 1. Experimental 1H (Solution, CDCl3) and 13C (Solid State) NMR Shifts for Compounds 1a–d and 2a–d Measured at 313 K.
| compound | atom | 1H (solution) | 13C (solid state) |
|---|---|---|---|
| 1a | C1 | +1155 | |
| (R = H) | H2/C2 | +26.9 | +316 |
| H3/C3 | +38.1 | –421 | |
| 1b | C1 | a | |
| (R = CN) | C2 | a | |
| H3/C3 | +44.4 | –340 | |
| 1c | C1 | a | |
| (R = Cl) | C2 | a | |
| H3/C3 | +38.9 | –250 | |
| 1d | C1 | a | |
| (R = Br) | C2 | a | |
| H3/C3 | +38.1 | –320 | |
| 2a | C1 | +96 | |
| (R = H) | H2/C2 | –20.1 | –67 |
| H3/C3 | +0.9 | +929 | |
| 2b | C1 | +96 | |
| (R = CN) | C2 | –87 | |
| H3/C3 | +2.4b | +750 | |
| 2c | C1 | +123 | |
| (R = Cl) | C2 | –143 | |
| H3/C3 | +6.2 | +800 | |
| 2d | C1 | +162 | |
| (R = Br) | C2 | –160 | |
| H3/C3 | +7.0 | +804 |
Not observed.
Measured in methanol-d4.
The 13C NMR shifts clearly indicate the different electron configurations and spin distributions in the groups of compounds 1 and 2. In compound 1a, the carbonyl atom C1 is greatly paramagnetically deshielded (δ = +1155 ppm), whereas the methyl C3 is notably shielded (δ = −421 ppm). The signal assignment pattern is reversed for 2a: C1 is slightly shielded compared to its diamagnetic counterparts, but C3 is greatly deshielded (δ = +929 ppm). The experimental (solid state) and DFT-calculated (in vacuo) 13C NMR shifts of atoms C1–C3 in compounds 1a and 2a are shown in Figure 3.
Figure 3.
Experimental (solid state, stripe pattern) and theoretical (DFT-calculated in vacuo using ADF2017, see Section 4) 13C NMR shifts of atoms C1 (red), C2 (green), and C3 (blue) in compounds 1a (left) and 2a (right) at 313 K. The estimated uncertainty of the experimental NMR shifts is shown by standard error bars.
The NMR shifts for compounds 1 reported in this work have been calculated assuming a quartet ground state, eq 2. This is perfectly justified by the vanishingly small experimental ZFS (D < 0.6 cm–1 for 1a)28,29 and our comparison between the NMR shifts calculated with and without ZFS (Table S2). To determine the physical contributions to the hyperfine NMR shifts that dominate the trends in Figure 3, we performed theoretical calculations of the hyperfine coupling tensors at the four-component relativistic DFT level (Table S3 in the Supporting Information).30 The values obtained for the isolated molecules are summarized in Table S4 and confirm the dominant role of the Fermi-contact contributions related to the distribution of electron-spin density in the β-diketonato moiety, which will be discussed in detail in Section 2.3. Note that the NMR shifts for 2a calculated in vacuo (Figure 3) agree nicely with those reported previously (+1153, +149, and −44 ppm).21
The successful observation of all of the 13C NMR resonances of parent compounds 1a and 2a (including those at around 1000 ppm) prompted us to investigate the set of compounds 1 and 2 shown in Figure 1 and to analyze the effects of substituents on the NMR shifts. However, for the chromium complexes 1, we obtained the NMR resonances of C1 (+1155 ppm) and C2 (+316 ppm) only for compound 1a. Our inability to detect NMR signals of these two atoms for the other Cr(III) complexes can be rationalized by either the presence of several molecules in the asymmetric unit of our crystalline samples or very fast transverse relaxation or both. However, as the trends in NMR shifts for 1a and 2a are satisfactorily reproduced by theoretical calculations (see Figure 3), we visualize the theoretical substituent-dependent 13C NMR shift of C2 in Figure 4. Substituting bromine for hydrogen in compound 1 results in an additional deshielding of C2 by almost 700 ppm (compare δC2tot for 1a and 1d in Figure 4), accounted for mainly by the contribution of the hyperfine shift. The electronic origins of the individual hyperfine contributions will be discussed in detail in Sections 2.3 and 2.4.
Figure 4.
Experimental (solid state) and theoretical (DFT-calculated in vacuo using ADF2017) 13C NMR shifts of C2 for compounds 1a–1e (left) and 2a–2e (right) at 313 K.
2.2. DFT Calculation of NMR Shifts at Experimental Conditions
Despite the general agreement of the trends in the NMR shifts shown in Figures 3 and 4, there is a significant mismatch between the experimental solid-state data and the theoretical values calculated in vacuo, as shown in Table 2 and Figure 5. We therefore investigated the individual physical effects influencing the NMR shifts. Because of the planar nature of the Cu(II) compounds and their assumed greater sensitivity to intermolecular contacts including cation−π interactions, compound 2a is analyzed in this section as an example.
Table 2. NMR Shifts for Compounds 1a (S = 3/2) and 2a (S = 1/2) Calculated Using PBE0/TZ2P In Vacuo or in a Supramolecular Clusterc.
| atom | vacuoa | vacuob | clust(3) | clust(3↑↓)d | Δ↑↓ | clust(9) | clust(9) + Δ↑↓ | exp | |
|---|---|---|---|---|---|---|---|---|---|
| 1a | C1 | +1337 | +1304 | +1291 | +1286 | +1155 | |||
| C2 | +248 | +254 | +268 | +266 | +316 | ||||
| C3 | –569 | –567 | –551 | –531 | –421 | ||||
| 2a | C1 | +137 | +107 | +86 | +86 | +0 | +83 | +83 | +96 |
| C2 | –17 | –25 | –32 | –57 | –25 | –39 | –64 | –67 | |
| C3 | +1167 | +1168 | +1092 | +988 | –104 | +1059 | +955 | +929 |
The geometry of the central molecule was optimized in the supramolecular cluster unless stated otherwise; see Section 4.
Geometry optimized in vacuo (EPR and NMR parameters calculated in vacuo using ADF2017).
Geometry optimized in clust(9) (EPR and NMR parameters calculated in vacuo using ADF2019 and ADF2017, respectively); clust(3): two cluster neighbors are diamagnetic (with Ca atoms).
clust(3↑↓): all three molecules are open shell (see Table 3); Δ↑↓: calculated as clust(3↑↓) – clust(3); clust(9): eight cluster neighbors are diamagnetic (with Ga and Ca atoms for 1a and 2a, respectively); for Cartesian coordinates, see the Supporting Information; exp: experimental solid-state NMR data.
Figure 5.
(a) Contributions of individual physical effects to the NMR shift of C3 in compound 2a as described in Table 2. (b) Side view of the supramolecular arrangement of two nearest-neighbor molecules in the small cluster (clust-3) of compound 2a indicating intermolecular cation−π interaction between Cu and C2 (plotted by the OLEX2 program31).
2.2.1. Crystal Packing
First, we investigated the effect of crystal packing on the NMR shift using the cluster approach.32−34 The central molecule of interest was embedded in a cluster of eight surrounding molecules generated from the X-ray structure ACACCU0435 (R = 2.2), clust(9). Calcium atoms were substituted for copper in all of the embedding diamagnetic molecules (for the cluster arrangement, see Cartesian coordinates in the Supporting Information). The NMR shift of C3 in the central molecule of 2a in this large clust(9) changed by 109 ppm, as shown in Table 2 and Figure 5. However, approximately 70% of this supramolecular effect on the C3 NMR shifts for 2a had already been reproduced by the two nearest neighbors in clust(3). Therefore, we analyze only the small clust(3) of 2a in the following section.
2.2.2. Effect of Neighboring Paramagnetic Molecules
To investigate the hyperfine effects of paramagnetic neighbors
on the NMR shifts of the central molecule of 2a in the
solid state,36,37 we resorted to the smallest cluster,
the three molecules shown in Figure 5b. Because the magnetic coupling constant and zero-field
splitting are negligible, the electronic ground state of this clust(3↑↓) is eight-times degenerate (one quartet and
two doublets). This degeneracy permits us to use eight single Slater
determinants (ΨSD) to describe its electronic states.
To choose unique ΨSD (with different NMR values),
we fix the α electron-spin state at the central
probing molecule (labeled as a bold ↑). As a result, four spin
combinations at the two neighboring paramagnetic centers are possible,
one ΨSD with three unpaired spins (↑
↑)
and three ΨSD with one unpaired spin (↓
↑,
↑
↓,
and ↓
↓).
The EPR and NMR parameters calculated
for the individual spin systems using the ZORA approach are summarized
in Table 3. Clearly, the differences between the 13C NMR shifts calculated for the individual ↓
↑,
↑
↓,
and ↓
↓
states are rather marginal (<15
ppm). However, the NMR values for the ↑↑↑ state are notably different. Its 25% contribution to clust(3↑↓) brings the theoretical values significantly
closer to the experiment. The averaged 13C NMR shifts are
included in Table 2 to calculate the shielding correction (Δ↑↓) used further for the calculation on the large clust(9). Summing
up all of the physical effects in Table 2 results in very satisfactory agreement between
the theoretical and experimental 13C NMR data for 2a (Δδ < 30 ppm).
Table 3. Theoretical Boltzmann Populations (in %) and 13C NMR Shifts (in ppm), Calculated Using the ZORA Approach and the PBE0 Functional, for the Individual Spin States of clust(3↑↓) and the Averaged Valuesa.
↑ ↑ |
↓ ↑ |
↑ ↓ |
↓ ↓ |
clust(3↑↓) | |
|---|---|---|---|---|---|
| population | 25 | 25 | 25 | 25 | |
| C1 | +18 | +104 | +105 | +118 | +86 |
| C2 | –192 | –27 | –28 | –22 | –57 |
| C3 | +1367 | +857 | +861 | +868 | +988 |
2.2.3. Comment on the Selection of the Density Functional
The calculation of NMR and EPR parameters has previously been shown to be very sensitive to the density functional, particularly the amount of the exact-exchange admixture used.5,24,38 The effect of the exact-exchange admixture (0–40%) in the PBE-based functional for compounds 1a and 2a is shown in Table S8 and Figure S3 (Supporting Information). This effect amounts to up to 500 ppm (C1) and 800 ppm (C3) for compounds 1a and 2a, respectively. However, the use of a reliable structural model and the simulation of crystal packing by the cluster approach provide very good agreement between the theoretical and experimental NMR data (Table 2). Therefore, we used 25% of the exact-exchange admixture as implemented in the standard PBE0 functional for all of our production calculations. To interpret the differences in the NMR shifts between compounds 1 and 2, we first analyze and discuss their electronic structures.
2.3. Differences between Compounds 1a (d3) and 2a (d9)
2.3.1. Electronic Structure
Compound 1a contains three 3d electrons, each occupying a metal-centered dxy-, dxz-, or dyz-based molecular spin orbital (MSO, S = 3/2), as shown schematically in Figure 6a. These single-occupied orbitals are of π-symmetry relative to the six Cr–O bonds. Because of the π-symmetry, the spin polarization in π-space is expected to be the main mechanism resulting in a Fermi-contact contribution to the ligand hyperfine 13C NMR shifts of 1a. However, an additional donation of α-spin density (Hund’s rule) to the formally vacant dx2–y2 and dz2 would result in an overabundance of β-spin density in the ligand σ-space (see below).
Figure 6.
Schematic representation of the energy and occupation of metal 3d orbitals in compounds (a) 1a and (b) 2a. Note that the ligands occupy x, y, z and x, y positions in the octahedral and square-planar complexes, respectively.
In contrast, 2a has nine 3d electrons. The metal-centered nonbonding molecular spin orbitals (MSOs) of dxy-, dxz-, dyz-, and dz2-type are approximately paired, whereas a single unpaired electron (S = 1/2) resides in dx2–y2; see Figure 6b. The unpaired density is thus found in the σ-space of all of the four Cu–O bonds. Therefore, the σ-bond spin delocalization and hyperconjugation interactions39 are expected to dominate the Fermi-contact mechanism of the ligand hyperfine 13C NMR shifts in 2a. The total spin densities for compounds 1a and 2a, calculated at the scalar-relativistic (1c) level of theory, are visualized in Figure 7.
Figure 7.

Visualization of the scalar-relativistic (1c) spin density (isosurface at ± 0.0002 au) and the principal components of the g-tensor for (a) 1a and (b) 2a, calculated using DFT at the spin–orbit (2c) level; see Section 4. The predominance of the α and β spin density is shown in blue and red, respectively.
As shown in Figure 7, the isotropic g factors (giso) are opposite for the compounds 1a and 2a. Note in passing that our calculated giso values of 1.98 (1a) and 2.10 (2a) are in very good agreement with the experimental data reported previously (1.98 for 1a,28 2.12 for 2a(40)). The signs of the g-shifts (Δgiso) have been interpreted in terms of the electronic structure since the dawn of EPR spectroscopy.19,41 The angular magnetic couplings in the α-spin space involving a singly occupied molecular orbital (SOMO) result in negative g-shifts (Δgiso) for early TM complexes (d1–d3).18,19 This applies to the d3 compound 1a (Δgiso = −20 ppt), where three singly occupied 3d α orbitals can form efficient angular magnetic couplings with two unoccupied 3d α orbitals, whereas three singly unoccupied 3d β orbitals have no occupied β d-orbitals nearby with which to form efficient magnetic couplings (see Figure 6). In contrast, couplings in the β-spin space involving singly unoccupied molecular orbital (SUMO) result in positive g-shifts for the late TM complexes (d6–d9).18,19 This clearly applies to the d9 compound 2a (Δgiso = +96 ppt) because several occupied 3d β orbitals are available for the β SUMO to form angular magnetic couplings, whereas no unoccupied α orbitals are accessible to the α SOMO (see Figure 6). This has been shown to be the case for 3d TM complexes with acetylacetonate involving V, Ni, and Cu.21
2.3.2. Spin Delocalization and Polarization
To analyze the hyperfine coupling pathways for the individual atoms C1–C3 in compounds 1a and 2a in detail, the spin densities are plotted separately for α and β spaces in Figure 8.
Figure 8.
Visualization of the spin density (isosurface at ± 0.0001 au) for compounds (a) 1a and (b) 2a, separated into the α (top) and β (bottom) parts.
In compound 1a, the spin-propagation mechanism involves two distinct pathways. First, the spin polarization of the σCr–O bond, linked with the donation of α-spin density (Hund’s rule) to the vacant dx2–y2 and dz2 orbitals, as shown in Figure 6a, leaves β-spin density in the σ-space of the ligand, particularly at the oxygen atom (spin population −0.020 au). This polarization is partially transferred to the C3 atom by the Cr–O ↔ C1–C3 σ-hyperconjugation mechanism, giving rise to a negative spin population at C3 (−0.004 au) and a negative hyperfine 13C NMR shift (−421 ppm); see Figure 8a (bottom). Second, conjugation of the π-type single-occupied metal-based d-orbital(s) with the ligand π-type orbitals enables the out-of-plane spin-delocalization mechanism (Figure 8a, top), resulting in positive spin populations (predominance of α) in the π-space of C1 (+0.016 au) and C2 (+0.007 au) and positive hyperfine shifts for these carbon atoms. This observation is in agreement with the previously reported data for a series of Cr(III)42 and Ru(III)24 β-diketonates, which indicated considerable π-type delocalization of the metal electrons through the π* orbitals of the ligands.
Because of the different symmetry of the SOMO (dx2–y2 in Figure 6b) in compound 2a, the α-delocalization on the σCu–O bond results in positive spin density at the oxygen atom (+0.077 au); see Figure 8b. Further, this α-delocalization is very efficiently transferred to C3 by the Cu–O ↔ C1–C3 hyperconjugation in σ-space, resulting in a δC3HF > 1000 ppm. In passing, note the expected absence of a π-mechanism for the C3 atom in both compounds. However, in contrast to that in 1a, C1 is polarized more indirectly in the π-space of 2a, therefore, in the opposite way and notably less efficiently (−0.005 au). Similarly, the in-plane spin imbalance at C2 is relatively weak and probably results from the interplay of two mechanisms: (i) σ-hyperconjugation Cu–O ↔ C1–C2 interaction and (ii) π-polarization by two neighboring C1 atoms.
The in-plane spin densities highlighting inverted spin-polarization patterns on the O–C1–C2–H2 coupling pathway for compounds 1a and 2a are shown in Figure 9. Note the completely opposite patterns at the C1–C2 bond and around C2, which is propagated to the opposite spin polarization at the H2 atom. This is reflected in the positive and negative δH2HF values for 1a and 2a, respectively. The theoretical hyperfine NMR shifts (calculated in vacuo) of +34.7 ppm (1a) and −26.9 ppm (2a) correspond nicely to the total experimental values (measured in a CDCl3 solution) of +26.9 and −20.1 ppm, respectively, Table 1. Clearly, the opposite patterns of the spin densities in the acac ligands result from the opposite situations at the central metal atoms and the different mechanisms of spin propagation. All of the observations discussed above are even more pronounced in compounds bearing a polarizable halogen substituent at the C2 atom, as analyzed in detail for the bromo compounds 1d and 2d in the following section.
Figure 9.

Visualization of the in-plane spin density in the acac ligand for compounds (a) 1a and (b) 2a, highlighting the inverted spin-density patterns of the hyperfine coupling pathways toward the C2 and H2 atoms. The atomic spin populations and hyperfine 1H NMR shifts for H2 are included. The predominance of α and β spin density is shown in blue and red, respectively.
2.4. Substituent Effects on the Spin Densities and Hyperfine NMR Shifts
The Fermi-contact contribution, which dominates the hyperfine 13C NMR shifts discussed in the previous sections, is analyzed here for 1a vs 1d and 2a vs 2d to reveal the electronic nature of the substituent-induced NMR shifts observed (see Figure 4).
As discussed in Section 2.3, the spin delocalization to C2 in compound 1a via π-space results in a hyperfine 13C NMR shift of +136 ppm. The presence of a bromine atom at C2 in compound 1d provides more efficient spin polarization, resulting in δC2HF = +829 ppm. The difference of almost +700 ppm between these two compounds results from repolarization in the p-orbitals and additional α-spin polarization of the s-orbitals in the bromine compound, as shown by the spatial distribution of the spin density and the differences in the orbital spin populations between 1d and 1a in Figure 10a.
Figure 10.

Visualization of the out-of-plane spin density on the C2–X bond for compounds (a) 1a (X = H) and 1d (X = Br) and (b) 2a (X = H) and 2d (X = Br). The differences in the spin populations of the AOs between derivatives a (top) and d (bottom) are also presented.
Similarly, the more negative hyperfine 13C NMR shift of C2 for 2d as compared with that for 2a reflects an increased imbalance in the spin distribution around the C2 nucleus favoring β-spin density; see Figure 10b.
The observations and differences in spin distribution summarized above result from the different electronegativities of the X atom (H vs Br) at C2, the polarizabilities of the C2–X bonds, and the related different degrees of hybridization at C2. The more polarizable and longer C2–Br bond involves more carbon 2p-character, leaving the carbon 2s orbital less affected and more symmetrically distributed at all three bonds around C2 and, therefore, available for symmetrical polarization around the carbon nucleus; see Figure 10, bottom. In general, this rearrangement results in a significantly larger Fermi-contact contribution to the hyperfine NMR shifts for the d compounds. Similar arguments apply to the other derivatives with halogens, compounds c and e (cf., Figure 4). A detailed investigation of the effects of substituents on the spin-polarization and hyperfine coupling pathways in a large set of transition-metal complexes is in progress in our laboratories.
3. Conclusions
Carefully measured, calculated, and interpreted ligand NMR data for two series of paramagnetic acetylacetonate complexes are reported. For the first time, solid-state 13C NMR resonances for greatly paramagnetically deshielded carbon atoms in Cr(III) and Cu(II) compounds were obtained by a short spin-echo MAS experiment. The trends in experimental NMR shifts have been reproduced by the DFT calculations for isolated molecules. However, quantitative agreement between the experimental solid-state and theoretical DFT values is shown to be achieved only if the effects of the supramolecular interactions in the crystal and nearest-neighbor paramagnetic molecules are included. Our contribution provides a guide for many future NMR investigations of complex systems.
The differences in 13C NMR resonances between octahedral d3 Cr(acac)3 and square-planar d9 Cu(acac)2 compounds are rationalized by the molecular symmetry, the electronic and spin structure, and different spin-delocalization mechanisms involving conjugation in π-space and hyperconjugation in σ-space. Further, the substituent-induced hyperfine effects are shown to amount up to 700 ppm, calculated as a difference between the 13C shifts of C2 in compounds with the hydrogen (a) and bromine (d) atom. However, a full understanding of this substituent-induced effect requires that additional analyses be performed and it cannot be rationalized by simply plotting the spin densities and total atomic spin populations. The analysis of hyperfine coupling pathways in a heavy-element series of acetylacetonates using hyperfine coupling densities is in progress in our laboratories.
4. Experimental and Theoretical Methods
4.1. Materials and Synthesis
The starting compounds CrCl3·xH2O, CuCl2·xH2O, 2,4-pentanedione, N-bromosuccinimide, NaHCO3, K2CO3, Cu(NO3)2·3H2O, acetic anhydride, and CH3COONa were used as obtained from our suppliers. The solvents in p.a. grade were used as received without further purification. The compounds were synthesized according to slightly modified procedures reported previously, as described in the Supporting Information.
4.2. NMR Spectroscopy
4.2.1. Solid-State NMR
1H and 13C MAS NMR spectra of compounds 1 and 2 were recorded on a Bruker AVANCE-II spectrometer at a magnetic field of 4.7 T (200.06 and 50.31 MHz for 1H and 13C, respectively) using a home-built MAS probe for 1.8 mm Si3N4 rotors. The NMR shifts are given using a TMS scale. The temperature of a spinning sample depends on the spinning speed and on the air flow in the driving and bearing channels. The actual temperature of the sample has been calibrated using the NMR shift of lead nitrate, Pb(NO3)2, which was spun under identical conditions.43
The 1H MAS spectra were recorded with the spin-echo pulse sequence π/2−τ–π–τ–ACQ, where π/2 = 1.5 μs. The sample spinning rate was 40 kHz, and the echo delay was set to one sample rotation period τ = τr = 25 μs. The spectra were recorded from 1 to 2 × 103 repetitions and with a 100–200 ms of relaxation delay.
The 13C MAS spectra were also recorded with a spin-echo pulse sequence (π/2 = 2.5 μs), but a minimal echo delay of 7–10 μs was used to minimize the signal decay of the broad NMR lines with the short relaxation times. A spin-echo mismatch caused the spinning sidebands to appear out of phase; see Figure 2. The sample spinning rate was 30 kHz, and the spectra were recorded from 0.1 to 2.5 × 106 repetitions with a 100–200 ms relaxation delay.
4.2.2. NMR in Solution
The 1H NMR spectra of compounds 1 and 2 were recorded on Bruker AVANCE NEO spectrometers at frequencies of 500 and 700 MHz. The NMR shifts are reported relative to TMS, the reference used. The temperature of the NMR sample was calibrated as reported previously.24
4.3. Quantum Chemical Calculations
4.3.1. Geometry of the Isolated Molecules
The molecular geometries of compounds 1 and 2 were fully optimized in vacuo using the PBE0 functional44,45 and the def2-TZVPP basis set (with the corresponding def2-ECP for heavier atoms)46 as implemented in the Turbomole 7.00 program.47 All calculations using this optimized protocol48,49 were performed with the m5 integration grid and the following convergence criteria: 10–6 for the energy change and 10–3 for the geometry gradient.
4.3.2. Geometry in the Solid State
The starting structures for compounds 1a and 2a were extracted from the X-ray data with reference codes ACACCR07 (R = 4.2)50 and ACACCU04 (R = 2.2),35 respectively. The central molecule of interest was embedded in a cluster of eight surrounding diamagnetic molecules (Ga and Ca were substituted for Cr and Cu, respectively) generated from the crystal packing. We used an approach calibrated in our previous studies of octahedral and square-planar transition-metal complexes to optimize the molecular geometry of the central molecule.5,48 The surrounding cluster was kept fixed (for 1a) except the positions of all of the hydrogen atoms (for 2a) that were relaxed (for Cartesian coordinates, see the Supporting Information). The geometry was optimized using density-functional theory (DFT) with the PBE0 functional44,45 and the def2-TZVPP basis set46 for all atoms of the central molecule and the def2-SVP basis set for all atoms of the surrounding cluster, with the corresponding relativistic effective core potentials (def2-ECPs)51 for the metal center. The dispersion correction D3-BJ52,53 was employed to treat weak supramolecular interactions in the cluster.
4.3.3. Two-Component Calculation of the NMR and EPR Parameters
The total NMR shifts were calculated as the sum of the orbital (δLorb) and hyperfine (δL) contributions, according to eq 1. The orbital shift of the ligand atom L is obtained theoretically as the difference between the shielding constants of a reference diamagnetic compound (σref) and a spectator atom L (δLorb)
| 4 |
The NMR and EPR parameters were calculated using the zeroth-order regular approximation (ZORA)54,55 Hamiltonian, as implemented in the ADF program package (versions 2017 and 2019).56−58 The orbital NMR shielding constants and zero-field splittings for open-shell systems were determined only at the scalar (spin-free, 1c) ZORA level. In the production calculations, we used the PBE0 functional (for other functionals, see the Supporting Information), the TZ2P basis sets,59 and either vacuum or the cluster model derived from the X-ray geometry (see above). The calculated orbital shifts (δLorb) were referenced relative to adamantane as a secondary reference (δref = 1.82 ppm for 1H and δref = 38.48 ppm for 13C) and are reported relative to TMS (primary reference). The hyperfine shifts (δL) were calculated at the 2c (SO-ZORA) level using the same functional and basis sets as used for the calculation of the orbital contributions. The zero-field splitting (ZFS) parameters D and E were included in the testing calculations using the pNMRShift program.60,61 For details and the effects of ZFS on the calculated hyperfine NMR shifts, see the Supporting Information. The reported values were obtained by averaging the corresponding values for all chemically equivalent atoms. The net atomic and orbital spin populations were calculated at the 1c level of theory in the ADF program (ZORA/PBE0) using the Mulliken approach62 consistent with our previous reports on paramagnetic Ru(III) compounds.5,24
4.3.4. Four-Component Calculation of the Fermi-Contact, Spin-Dipolar, and Paramagnetic Spin–Orbit Contributions to the Hyperfine NMR Shift
The separation of δLHF into individual physical terms according to eq 3 was performed as described previously16,63 at the four-component level of theory employing the ReSpect program package.30
Acknowledgments
R.M and I.H. are grateful to Dr. Jasper Adamson, Tallinn, for his help in initiating this Czech–Estonian collaborative project. This work has received support from the Czech Science Foundation (Grant numbers 21-06991S and 18-05421S to R.M). CIISB, Instruct-CZ Center of Instruct-ERIC EU consortium, funded by the MEYS CR infrastructure project LM2018127, is gratefully acknowledged for financial support of the measurements made at the Core Facility Josef Dadok National NMR Center. The computational resources were supported by MEYS CR from the Large Infrastructures for Research, Experimental Development and Innovations project “e-Infrastructure CZ”-LM2018140. Financial support was also received from the Slovak Grant Agency APVV (Grant No. APVV-19-0516 to S.K.) and the European Regional Development Fund (Grant No. TK134 to I.H).
Supporting Information Available
The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.inorgchem.1c00204.
Solid-state NMR spectroscopy, DFT calculations, synthesis, and cartesian coordinates (PDF)
Author Contributions
∇ J.N. and L.J. contributed equally to this work.
The authors declare no competing financial interest.
Supplementary Material
References
- La Mar G. N.; DeW Horrocks W.; Holm R. H.. NMR of Paramagnetic Molecules; Academic Press: New York, 1973. [Google Scholar]
- Bertini I.; Luchinat C.; Parigi G.; Ravera E.. NMR of Paramagnetic Molecules: Applications to Metallobiomolecules and Models; Elsevier: Amsterdam, 2016. [Google Scholar]
- Kaupp M.; Köhler F. H. Combining NMR Spectroscopy and Quantum Chemistry as Tools to Quantify Spin Density Distributions in Molecular Magnetic Compounds. Coord. Chem. Rev. 2009, 253, 2376–2386. 10.1016/j.ccr.2008.12.020. [DOI] [Google Scholar]
- Pell A. J.; Pintacuda G.; Grey C. P. Paramagnetic NMR in Solution and the Solid State. Prog. Nucl. Magn. Reson. Spectrosc. 2019, 111, 1–271. 10.1016/j.pnmrs.2018.05.001. [DOI] [PubMed] [Google Scholar]
- Novotný J.; Sojka M.; Komorovsky S.; Nečas M.; Marek R. Interpreting the Paramagnetic NMR Spectra of Potential Ru(III) Metallodrugs: Synergy between Experiment and Relativistic DFT Calculations. J. Am. Chem. Soc. 2016, 138, 8432–8445. 10.1021/jacs.6b02749. [DOI] [PubMed] [Google Scholar]
- Moon S.; Patchkovskii S.. First-Principles Calculations of Paramagnetic NMR Shifts. In Calculation of NMR and EPR Parameters. In Theory and Applications; Kaupp M.; Bühl M.; Malkin V. G., Eds.; Wiley-VCH: Weinheim, 2004; pp 325–328. [Google Scholar]
- Komorovsky S.; Repisky M.; Ruud K.; Malkina O. L.; Malkin V. G. Four-Component Relativistic Density Functional Theory Calculations of NMR Shielding Tensors for Paramagnetic Systems. J. Phys. Chem. A 2013, 117, 14209–14219. 10.1021/jp408389h. [DOI] [PubMed] [Google Scholar]
- Van den Heuvel W.; Soncini A. NMR Chemical Shift in an Electronic State with Arbitrary Degeneracy. Phys. Rev. Lett. 2012, 109, 073001 10.1103/PhysRevLett.109.073001. [DOI] [PubMed] [Google Scholar]
- Van den Heuvel W.; Soncini A. NMR Chemical Shift as Analytical Derivative of the Helmholtz Free Energy. J. Chem. Phys. 2013, 138, 054113 10.1063/1.4789398. [DOI] [PubMed] [Google Scholar]
- Hrobárik P.; Reviakine R.; Arbuznikov A. V.; Malkina O. L.; Malkin V. G.; Köhler F. H.; Kaupp M. Density Functional Calculations of NMR Shielding Tensors for Paramagnetic Systems with Arbitrary Spin Multiplicity: Validation on 3d Metallocenes. J. Chem. Phys. 2007, 126, 024107 10.1063/1.2423003. [DOI] [PubMed] [Google Scholar]
- Gendron F.; Sharkas K.; Autschbach J. Calculating NMR Chemical Shifts for Paramagnetic Metal Complexes from First-Principles. J. Phys. Chem. Lett. 2015, 6, 2183–2188. 10.1021/acs.jpclett.5b00932. [DOI] [PubMed] [Google Scholar]
- Rouf S. A.; Mareš J.; Vaara J. 1H Chemical Shifts in Paramagnetic Co(II) Pyrazolylborate Complexes: A First-Principles Study. J. Chem. Theory Comput. 2015, 11, 1683–1691. 10.1021/acs.jctc.5b00193. [DOI] [PubMed] [Google Scholar]
- Vaara J.; Rouf S. A.; Mareš J. Magnetic Couplings in the Chemical Shift of Paramagnetic NMR. J. Chem. Theory Comput. 2015, 11, 4840–4849. 10.1021/acs.jctc.5b00656. [DOI] [PubMed] [Google Scholar]
- Vaara J.Chemical Shift in Paramagnetic Systems. In High Resolution NMR Spectroscopy: Understanding Molecules and Their Electronic Structures. In Science and Technology of Atomic Molecular Condensed Matter and Biological Systems; Contreras R. H.; Contreras R. H., Eds.; Elsevier, 2013; pp 41–67. [Google Scholar]
- Autschbach J.NMR Calculations for Paramagnetic Molecules and Metal Complexes. In Annual Reports in Computational Chemistry; Dixon D. A., Ed.; Elsevier, 2015; Chapter 1, Vol. 11, pp 3–36. [Google Scholar]
- Novotný J.; Přichystal D.; Sojka M.; Komorovsky S.; Nečas M.; Marek R. Hyperfine Effects in Ligand NMR: Paramagnetic Ru(III) Complexes with 3-Substituted Pyridines. Inorg. Chem. 2018, 57, 641–652. 10.1021/acs.inorgchem.7b02440. [DOI] [PubMed] [Google Scholar]
- Chyba J.; Novák M.; Munzarová P.; Novotný J.; Marek R. Through-Space Paramagnetic NMR Effects in Host–Guest Complexes: Potential Ruthenium(III) Metallodrugs with Macrocyclic Carriers. Inorg. Chem. 2018, 57, 8735–8747. 10.1021/acs.inorgchem.7b03233. [DOI] [PubMed] [Google Scholar]
- Bora P. L.; Novotný J.; Ruud K.; Komorovsky S.; Marek R. Electron-Spin Structure and Metal–Ligand Bonding in Open-Shell Systems from Relativistic EPR and NMR: A Case Study of Square-Planar Iridium Catalysts. J. Chem. Theory Comput. 2019, 15, 201–214. 10.1021/acs.jctc.8b00914. [DOI] [PubMed] [Google Scholar]
- Mabbs F. E.; Collison D.. Electron Paramagnetic Resonance of d Transition Metal Compounds; Elsevier Science: Amsterdam, 1992; Vol. 16, pp 338–441. [Google Scholar]
- Abragam A.; Bleaney B.. Electron Paramagnetic Resonance of Transition Ions; Oxford Classic Texts in the Physical Sciences; Oxford University Press: Oxford, New York, 2012. [Google Scholar]
- Andersen A. B. A.; Pyykkönen A.; Jensen H. J. A.; McKee V.; Vaara J.; Nielsen U. G. Remarkable Reversal of 13C-NMR Assignment in D1, D2 Compared to D8, D9 Acetylacetonate Complexes: Analysis and Explanation Based on Solid-State MAS NMR and Computations. Phys. Chem. Chem. Phys. 2020, 22, 8048–8059. 10.1039/D0CP00980F. [DOI] [PubMed] [Google Scholar]
- Pritchard B.; Autschbach J. Theoretical Investigation of Paramagnetic NMR Shifts in Transition Metal Acetylacetonato Complexes: Analysis of Signs, Magnitudes, and the Role of the Covalency of Ligand–Metal Bonding. Inorg. Chem. 2012, 51, 8340–8351. 10.1021/ic300868v. [DOI] [PubMed] [Google Scholar]
- Rastrelli F.; Bagno A. Predicting the NMR Spectra of Paramagnetic Molecules by DFT: Application to Organic Free Radicals and Transition-Metal Complexes. Chem. - Eur. J. 2009, 15, 7990–8004. 10.1002/chem.200802443. [DOI] [PubMed] [Google Scholar]
- Jeremias L.; Novotný J.; Repisky M.; Komorovsky S.; Marek R. Interplay of Through-Bond Hyperfine and Substituent Effects on the NMR Chemical Shifts in Ru(III) Complexes. Inorg. Chem. 2018, 57, 8748–8759. 10.1021/acs.inorgchem.8b00073. [DOI] [PubMed] [Google Scholar]
- Lennartson A.; Christensen L. U.; McKenzie C. J.; Nielsen U. G. Solid State 13C and 2H NMR Investigations of Paramagnetic [Ni(II)(Acac)2L2] Complexes. Inorg. Chem. 2014, 53, 399–408. 10.1021/ic402354r. [DOI] [PubMed] [Google Scholar]
- Rouf S. A.; Jakobsen V. B.; Mareš J.; Jensen N. D.; McKenzie C. J.; Vaara J.; Nielsen U. G. Assignment of Solid-State 13C and 1H NMR Spectra of Paramagnetic Ni(II) Acetylacetonate Complexes Aided by First-Principles Computations. Solid State Nucl. Magn. Reson. 2017, 87, 29–37. 10.1016/j.ssnmr.2017.07.003. [DOI] [PubMed] [Google Scholar]
- Kaupp M.; Bühl M.; Malkin V. G.. Calculation of NMR and EPR Parameters: Theory and Applications; Wiley-VCH: Weinheim, 2004. [Google Scholar]
- Singer L. S. Paramagnetic Resonance Absorption in Some Cr +3 Complexes. J. Chem. Phys. 1955, 23, 379–388. 10.1063/1.1741973. [DOI] [Google Scholar]
- Sugisaki K.; Toyota K.; Sato K.; Shiomi D.; Takui T. Behaviour of DFT-Based Approaches to the Spin–Orbit Term of Zero-Field Splitting Tensors: A Case Study of Metallocomplexes, M III (Acac) 3 (M = V, Cr, Mn, Fe and Mo). Phys. Chem. Chem. Phys. 2017, 19, 30128–30138. 10.1039/C7CP05533A. [DOI] [PubMed] [Google Scholar]
- Repisky M.; Komorovsky S.; Kadek M.; Konecny L.; Ekström U.; Malkin E.; Kaupp M.; Ruud K.; Malkina O. L.; Malkin V. G. ReSpect: Relativistic Spectroscopy DFT Program Package. J. Chem. Phys. 2020, 152, 184101 10.1063/5.0005094. [DOI] [PubMed] [Google Scholar]
- Dolomanov O. V.; Bourhis L. J.; Gildea R. J.; Howard J. aK.; Puschmann H. OLEX2: A Complete Structure Solution, Refinement and Analysis Program. J. Appl. Crystallogr. 2009, 42, 339–341. 10.1107/S0021889808042726. [DOI] [Google Scholar]
- Maliňáková K.; Novosadová L.; Lahtinen M.; Kolehmainen E.; Brus J.; Marek R. 13C Chemical Shift Tensors in Hypoxanthine and 6-Mercaptopurine: Effects of Substitution, Tautomerism, and Intermolecular Interactions. J. Phys. Chem. A 2010, 114, 1985–1995. 10.1021/jp9100619. [DOI] [PubMed] [Google Scholar]
- Babinský M.; Bouzková K.; Pipíška M.; Novosadová L.; Marek R. Interpretation of Crystal Effects on NMR Chemical Shift Tensors: Electron and Shielding Deformation Densities. J. Phys. Chem. A 2013, 117, 497–503. 10.1021/jp310967b. [DOI] [PubMed] [Google Scholar]
- Bouzková K.; Babinský M.; Novosadová L.; Marek R. Intermolecular Interactions in Crystalline Theobromine as Reflected in Electron Deformation Density and 13 C NMR Chemical Shift Tensors. J. Chem. Theory Comput. 2013, 9, 2629–2638. 10.1021/ct400209b. [DOI] [PubMed] [Google Scholar]
- Golchoubian H. Redetermination of Crystal Structure of Bis(2,4-Pentanedionato)Copper(II). Asian J. Chem. 2008, 20, 5834–5838. [Google Scholar]
- Bühl M.; Ashbrook S. E.; Dawson D. M.; Doyle R. A.; Hrobárik P.; Kaupp M.; Smellie I. A. Paramagnetic NMR of Phenolic Oxime Copper Complexes: A Joint Experimental and Density Functional Study. Chem. - Eur. J. 2016, 22, 15328–15339. 10.1002/chem.201602567. [DOI] [PubMed] [Google Scholar]
- Mali G.; Mazaj M. Hyperfine Coupling Constants in Cu-Based Crystalline Compounds: Solid-State NMR Spectroscopy and First-Principles Calculations with Isolated-Cluster and Extended Periodic-Lattice Models. J. Phys. Chem. C 2021, 125, 4655–4664. 10.1021/acs.jpcc.0c09651. [DOI] [Google Scholar]
- Hrobárik P.; Repiský M.; Komorovský S.; Hrobáriková V.; Kaupp M. Assessment of Higher-Order Spin–Orbit Effects on Electronic g-Tensors of D1 Transition-Metal Complexes by Relativistic Two- and Four-Component Methods. Theor. Chem. Acc. 2011, 129, 715–725. 10.1007/s00214-011-0951-7. [DOI] [Google Scholar]
- Martin B.; Autschbach J. Kohn-Sham Calculations of NMR Shifts for Paramagnetic 3d Metal Complexes: Protocols, Delocalization Error, and the Curious Amide Proton Shifts of a High-Spin Iron(II) Macrocycle Complex. Phys. Chem. Chem. Phys. 2016, 18, 21051–21068. 10.1039/C5CP07667F. [DOI] [PubMed] [Google Scholar]
- Adato I.; Eliezer I. Effect of the Solvent on the ESR Parameters of Copper Acetylacetonate. J. Chem. Phys. 1971, 54, 1472–1476. 10.1063/1.1675040. [DOI] [Google Scholar]
- Symons M. C. R.Chemical and Biochemical Aspects of Electron Spin Resonance Spectroscopy; Van Nostrand Reinhold Inc: New York, US, 1978. [Google Scholar]
- Lintvedt R. L.; Fatta N. M. Nephelauxetic and Spectrochemical Series for 1,3-Diketonates. Ligand Field Spectra of Some Tris(1,3-Diketonato)Chromium(III) Chelates. Inorg. Chem. 1971, 10, 478–481. 10.1021/ic50097a008. [DOI] [Google Scholar]
- Guan X.; Stark R. E. A General Protocol for Temperature Calibration of MAS NMR Probes at Arbitrary Spinning Speeds. Solid State Nucl. Magn. Reson. 2010, 38, 74–76. 10.1016/j.ssnmr.2010.10.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Adamo C.; Barone V. Toward Reliable Density Functional Methods without Adjustable Parameters: The PBE0 Model. J. Chem. Phys. 1999, 110, 6158–6170. 10.1063/1.478522. [DOI] [Google Scholar]
- Adamo C.; Scuseria G. E.; Barone V. Accurate Excitation Energies from Time-Dependent Density Functional Theory: Assessing the PBE0 Model. J. Chem. Phys. 1999, 111, 2889–2899. 10.1063/1.479571. [DOI] [Google Scholar]
- Schäfer A.; Huber C.; Ahlrichs R. Fully Optimized Contracted Gaussian Basis Sets of Triple Zeta Valence Quality for Atoms Li to Kr. J. Chem. Phys. 1994, 100, 5829–5835. 10.1063/1.467146. [DOI] [Google Scholar]
- TURBOMOLE V7.0 2015; TURBOMOLE GmbH. http://www.turbomole.com, 2007.
- Vícha J.; Patzschke M.; Marek R. A Relativistic DFT Methodology for Calculating the Structures and NMR Chemical Shifts of Octahedral Platinum and Iridium Complexes. Phys. Chem. Chem. Phys. 2013, 15, 7740–7754. 10.1039/c3cp44440f. [DOI] [PubMed] [Google Scholar]
- Vícha J.; Novotný J.; Straka M.; Repisky M.; Ruud K.; Komorovsky S.; Marek R. Structure, Solvent, and Relativistic Effects on the NMR Chemical Shifts in Square-Planar Transition-Metal Complexes: Assessment of DFT Approaches. Phys. Chem. Chem. Phys. 2015, 17, 24944–24955. 10.1039/C5CP04214C. [DOI] [PubMed] [Google Scholar]
- Chrzanowski L. S.; von Lutz M.; Spek A. L. α-Tris(2,4-Pentanedionato-κ 2 O, O ′)Cobalt(III) at 240, 210, 180, 150 and 110 K. Acta Crystallogr., Sect. C: Cryst. Struct. Commun. 2007, 63, m283–m288. 10.1107/S0108270107022950. [DOI] [PubMed] [Google Scholar]
- Andrae D.; Häußermann U.; Dolg M.; Stoll H.; Preuß H. Energy-Adjusted Ab Initio Pseudopotentials for the Second and Third Row Transition Elements. Theor. Chim. Acta 1990, 77, 123–141. 10.1007/BF01114537. [DOI] [Google Scholar]
- Grimme S.; Antony J.; Ehrlich S.; Krieg H. A Consistent and Accurate Ab Initio Parametrization of Density Functional Dispersion Correction (DFT-D) for the 94 Elements H-Pu. J. Chem. Phys. 2010, 132, 154104 10.1063/1.3382344. [DOI] [PubMed] [Google Scholar]
- Grimme S.; Ehrlich S.; Goerigk L. Effect of the Damping Function in Dispersion Corrected Density Functional Theory. J. Comput. Chem. 2011, 32, 1456–1465. 10.1002/jcc.21759. [DOI] [PubMed] [Google Scholar]
- van Lenthe E.; Snijders J. G.; Baerends E. J. The Zero-order Regular Approximation for Relativistic Effects: The Effect of Spin–Orbit Coupling in Closed Shell Molecules. J. Chem. Phys. 1996, 105, 6505–6516. 10.1063/1.472460. [DOI] [Google Scholar]
- Saue T. Relativistic Hamiltonians for Chemistry: A Primer. ChemPhysChem 2011, 12, 3077–3094. 10.1002/cphc.201100682. [DOI] [PubMed] [Google Scholar]
- te Velde G.; Bickelhaupt F. M.; Baerends E. J.; Fonseca Guerra C.; van Gisbergen S. J. A.; Snijders J. G.; Ziegler T. Chemistry with ADF. J. Comput. Chem. 2001, 22, 931–967. 10.1002/jcc.1056. [DOI] [Google Scholar]
- Guerra C. F.; Snijders J. G.; Velde G.; te Baerends E. J. Towards an Order-N DFT Method. Theor. Chem. Acc. 1998, 99, 391–403. 10.1007/s002140050353. [DOI] [Google Scholar]
- Baerends E. J.; Ziegler T.; Atkins A. J.; Autschbach J.; Baseggio O.; Bashford D.; Bérces A.; Bickelhaupt F. M.; Bo C.; Boerrigter P. M.; Cavallo L.; Daul C.; Chong D. P.; Chulhai D. V.; Deng L.; Dickson R. M.; Dieterich J. M.; Ellis D. E.; van Faassen M.; Fan L.. ADF2019, SCM, Theoretical Chemistry; Vrije Universiteit: Amsterdam: The Netherlands, 2019.
- Güell M.; Luis J. M.; Solà M.; Swart M. Importance of the Basis Set for the Spin-State Energetics of Iron Complexes. J. Phys. Chem. A 2008, 112, 6384–6391. 10.1021/jp803441m. [DOI] [PubMed] [Google Scholar]
- Autschbach J.; Patchkovskii S.; Pritchard B. Calculation of Hyperfine Tensors and Paramagnetic NMR Shifts Using the Relativistic Zeroth-Order Regular Approximation and Density Functional Theory. J. Chem. Theory Comput. 2011, 7, 2175–2188. 10.1021/ct200143w. [DOI] [PubMed] [Google Scholar]
- Martin B.; Autschbach J. Temperature Dependence of Contact and Dipolar NMR Chemical Shifts in Paramagnetic Molecules. J. Chem. Phys. 2015, 142, 054108 10.1063/1.4906318. [DOI] [PubMed] [Google Scholar]
- Mulliken R. S. Electronic Population Analysis on LCAO–MO Molecular Wave Functions, I. J. Chem. Phys. 1955, 23, 1833–1840. 10.1063/1.1740588. [DOI] [Google Scholar]
- Haase P. A. B.; Repisky M.; Komorovsky S.; Bendix J.; Sauer S. P. A. Relativistic DFT Calculations of Hyperfine Coupling Constants in 5d Hexafluorido Complexes: [ReF6]2– and [IrF6]2–. Chem. - Eur. J. 2018, 24, 5124–5133. 10.1002/chem.201704653. [DOI] [PMC free article] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.










