Next Article in Journal
On the Size of Superconducting Islands on the Density-Wave Background in Organic Metals
Previous Article in Journal
Prospects of Using Machine Learning and Diamond Nanosensing for High Sensitivity SARS-CoV-2 Diagnosis
Previous Article in Special Issue
NMR Magnetic Shielding in Transition Metal Compounds Containing Cadmium, Platinum, and Mercury
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Assessment of a Computational Protocol for Predicting Co-59 NMR Chemical Shift

by
Matheus G. R. Gomes
1,
Andréa L. F. De Souza
2,
Hélio F. Dos Santos
3,
Wagner B. De Almeida
4 and
Diego F. S. Paschoal
1,*
1
NQTCM: Núcleo de Química Teórica e Computacional de Macaé, Polo Ajuda, Instituto Multidisciplinar de Química, Centro Multidisciplinar UFRJ-Macaé, Universidade Federal do Rio de Janeiro, Macaé 27971-525, RJ, Brazil
2
LACASO: Laboratório de Catálise Aplicada e Síntese Orgânica, Departamento de Química Orgânica, Instituto de Química, Universidade Federal do Rio de Janeiro, Cidade Universitária, Rio de Janeiro 21941-909, RJ, Brazil
3
NEQC: Núcleo de Estudos em Química Computacional, Departamento de Química—ICE, Universidade Federal de Juiz de Fora, Campus Universitário, Juiz de Fora 36036-900, MG, Brazil
4
Laboratório de Química Computacional e Modelagem Molecular (LQC-MM), Departamento de Química Inorgânica, Instituto de Química, Universidade Federal Fluminense (UFF), Outeiro de São João Batista s/n, Campus do Valonguinho, Centro, Niterói 24020-141, RJ, Brazil
*
Author to whom correspondence should be addressed.
Magnetochemistry 2023, 9(7), 172; https://doi.org/10.3390/magnetochemistry9070172
Submission received: 9 May 2023 / Revised: 27 June 2023 / Accepted: 29 June 2023 / Published: 2 July 2023
(This article belongs to the Special Issue NMR Spectroscopy in Transition Metal Complexes)

Abstract

:
In the present study, we benchmark computational protocols for predicting Co-59 NMR chemical shift. Quantum mechanical calculations based on density functional theory were used, in conjunction with our NMR-DKH basis sets for all atoms, including Co, which were developed in the present study. The best protocol included the geometry optimization at BLYP/def2-SVP/def2-SVP/IEF-PCM(UFF) and shielding constant calculation at GIAO-LC-ωPBE/NMR-DKH/IEF-PCM(UFF). This computational scheme was applied to a set of 34 Co(III) complexes, in which, Co-59 NMR chemical shift ranges from +1162 ppm to +15,100 ppm, and these were obtained in distinct solvents (water and organic solvents). The resulting mean absolute deviation (MAD), mean relative deviation (MRD), and coefficient of determination (R2) were 158 ppm, 3.0%, and 0.9966, respectively, suggesting an excellent alternative for studying Co-59 NMR.

Graphical Abstract

1. Introduction

The use of cobalt (Co) coordination compounds in medicinal chemistry has increased in recent years. Studies in the literature have shown Co(III) complexes with antiviral [1,2,3,4,5], anti-inflammatory [5,6,7], antibacterial [5,8], and anticancer [9,10] activities. Regarding anticancer activities, Co(III) compounds act as cytotoxic ligand carriers to hypoxic regions, due to its redox properties [11]. Furthermore, cobalt is a less toxic metal than nonessential metals, such as platinum, and is present in biomolecules such as cobalamin, which represents an advantage its use in cancer treatments [5].
Nuclear Magnetic Resonance (NMR) spectroscopy provides fundamental information about the geometry and electronic structure of transition metal complexes [12]. As metals are important in different biological processes [13], NMR spectroscopy becomes a powerful technique for the speciation of transition metal complexes in solution and in biological systems [14]. The cobalt-59 (Co-59) nucleus is an NMR-active nucleus with a nuclear spin quantum number I = 7/2, and it has a natural isotopic abundance of 100% [15]. It is an important NMR probe because the signals are easily detectable in both solution and solid-states [16,17]. However, the high nuclear spin (I = 7/2) associated with a relatively large quadrupole moment can make the NMR peaks very broad [16,17], which implies low-resolution spectra. The Co-59 NMR chemical shift (δ59Co) spreads over a very broad range, about 18,000 ppm [16,17]. Moreover, the Co-59 NMR chemical shift is known for its sensitivity to the chemical environment in diamagnetic low-spin d6 Co(III) compounds [18], which highlights the value of Co-59 NMR in solving molecular structures, understanding isomerism, and assigning stereochemistry [19].
The computational study of NMR properties involving metallic nuclei is dependent on several factors, such as exchange-correlation in density functional theory (DFT), basis sets, solvent, and relativistic effects [12,20]. Some computational studies of Co-59 NMR are available in the literature [21,22,23,24]. Chan and Au-Yeung [22] proposed a computational scheme for predicting the δ59Co. Among a set of 13 Co(III) complexes, with a Co-59 NMR chemical shift covering a range of 11,000 ppm, at the GIAO-B3LYP/TZVP(Co)/IGLO-II(Ligands) level, the mean relative deviation (MRD) was less than 30%. Godbout and Oldfield [23] studied a set of six Co(III) complexes at the B3LYP/Wachters’(Co)/6-31G(d)(Ligands) level. The authors found a coefficient of determination (R2) of 0.98 for the correlation between the calculated Co-59 NMR shielding constants (σ59Co) and the experimental δ59Co, with a slope of -0.83. However, their calculated δ59Co from the experimental data had a mean absolute deviation (MAD) greater than 1000 ppm. Despite the high deviations among the experimental data, the model was able to quantitatively represent the trends. A set of four cobalt complexes were studied with the inclusion of implicit and explicit solvent effects by Bühl et al. [24]. Considering the CPMD (Carr-Parrinelo Molecular Dynamics) and QM/MM-BOMD (Quantum-Mechanical/Born-Oppenheimer Molecular Dynamics) approaches, the authors found a MAD of 400 ppm and 600 ppm, respectively. Furthermore, the authors showed that the solvent effects on both geometry and δ59Co were well-described using the PCM (Polarizable Continuum Model) implicit model. Four-component-relativistic (4c-Rel) calculations have been employed to calculate NMR chemical shifts in Co complexes [25,26,27,28]. Nonrelativistic (NRel) and relativistic (Rel) calculations of the NMR of light nuclei in the neighboring of metallic center were carried out by Semenov et al. [25]. The results showed that the shielding constant for 15N provides a relativistic deshielding correction of up to 9.3 ppm, showing that, even for a light nucleus, relativistic effects are important for an accurate description of the chemical shift and the solvent effect becomes more significant with larger dielectric constants. In their situation, there was a resulting increase of up to 13 ppm for the δ15N. Furthermore, Samultsev et al. [26] studied the σ59Co in [Co(NH3)5(OH2)]3+ complex and found a variation of 4.7% between the NRel and 4c-Rel values. When the Co metal center is replaced by iridium (Ir), the variation was 67.4%, demonstrating that the Rel corrections are more important for heavier nuclei. In another study, Samultsev et al. [27] calculated the δ59Co for a set of 27 Co complexes employing NRel and 4c-Rel approaches. The authors quantified the average contributions of relativistic and solvent effects on the calculated shielding constants as 4% and 1.4%, respectively. However, a direct comparison between calculated and experimental values of δ59Co was not performed. Recently, Samultsev et al. [28] studied the NMR shielding constants for Fe, Co, Ni, Pd, and Pt glycinates using the 4c-Rel calculation (4c-PBE0/dyall.ae3z level). They showed that the relativistic corrections resulted in an increase of 494 ppm and 483 ppm in the σ59Co of the fac-[Co(gly)3] and mer-[Co(gly)3] complexes, respectively. However, the net effect on the δ59Co was not reported.
Considering that there is still a gap in the literature regarding an accurate prediction of the Co-59 NMR chemical shift (δ59Co) in cobalt complexes, the present study describes the results of a broad benchmarking of DFT protocols to predict δ59Co for Co(III) complexes in solution. This is part of a continuous project aiming to calculate the NMR properties of transition metal complexes using our NMR-DKH basis sets, as described for Pt-195 [29,30,31] and Tc-99 [32] nuclei.

2. Theoretical Methodology

2.1. NMR-DKH Basis Set Development for Co

The NMR-DKH basis sets were proposed previously by Paschoal et al. for H-He, Li-Ne, Na-Ar, K-Ca, Ga-Kr, Rb-Sr, In-Xe, Pt [29], and Tc [32] atoms. These segmented all-electron relativistically contracted Douglas–Kroll–Hess (DKH) Gaussian basis sets were developed specifically for the NMR calculations, presenting excellent results in the study of Pt-195 [29,30,31], Xe-129 [33], and Tc-99 [32] NMR. In the present contribution, the same methodology was used in the development of a new NMR-DKH basis set for the Co atom [29,32].
Initially, for the construction of the new NMR-DKH basis set for Co, the maximum exponents per angular momentum (αl, l = s, p, d, f) were obtained with the Equation (1) [34,35,36,37,38,39]:
α l = k l   2 f l 2 π r l 2
where kl values are 1, 4/3, and 8/5, and fl is 33, 100, and 1000 for s, p, and d functions, respectively. In this equation, kl is a scaling factor used in generating the exponent of each angular momentum to produce enough tight exponents to describe the core. The innermost radial expectation values ( r _ l   , in Bohr), obtained from the multiconfigurational Dirac–Fock (MCDF) numerical calculations of the Co atom in the ground state ( ( _ ^ 4 ) F _ ( 9 / 2 ) ), were 0.056156329, 0.222395260, and 1.02444740 bohr for l = s, p, and d functions, respectively. The MCDF calculations were carried out with the GRASP90 program [40]. The following values were calculated for αl (in Bohr−2): 201,875.076137984 (s), 2288.26618462583 (p), and 51.2453777725453 (d).
In the next step, a series of descending primitives was generated according to Equation (2) [34,35,36,37,38,39]:
ζ = α l   χ i
where i is a positive integer, ζ corresponds to the Gaussian primitive exponent, and χ is a parameter used to determine the spacing and number of primitives. The χ values of χs = 2.50, χp = 2.75, and χd = 3.00 were considered. A total of 81 primitive basis functions were generated (18s11p6d). The basis set was contracted as a triple-ζ basis set, in which only the first set of each angular momentum was contracted. The contraction coefficients were obtained from the coefficients of the atomic orbitals calculated at unrestricted Hartree–Fock (UHF) level with the inclusion of scalar relativistic corrections through the second-order Douglas–Kroll–Hess (DKH2) approximation [41,42,43,44,45,46,47]. This calculation was performed in GAUSSIAN 16 Rev. C.01 program [48].
Finally, the basis set was augmented with the addition of three sets of f-polarization functions, with the exponents being adjusted in order to minimize the atomic energy at the UHF-DKH2 level in the presence of an electric field (z = 0.01 a.u.) [49]. This calculation was also performed in GAUSSIAN 16 Rev. C.01 program [48]. Subsequently, the two sets of f-polarization functions with the highest exponent were also contracted.
As a result, the new triple-ζ-doubly polarized (TZ2P) NMR-DKH basis set for the Co has a set of 102 primitives (GTO) and 59 contracted (CGTO) basis functions, with the following contraction scheme: ( 18 s 11 p 6 d 3 f ) [ 12 s 6 p 3 d 2 f ] . The NMR-DKH basis set for Co is found in the Supplementary Material or downloaded from the Basis Set Exchange portal (https://www.basissetexchange.org/) [50].

2.2. Benchmarking the Computational Protocols

Computational protocols were benchmarking for prediction of the Co-59 NMR chemical shift in Co(III) complexes. Initially, a set of five cobalt complexes (Figure 1) were selected—[Co(NH3)6]3+ (Cpx01) [51,52], [CoCl(NH3)5]2+ (Cpx02) [52,53], [Co(NO2)(NH3)5]2+ (Cpx03) [52,54], [Co(SCN)(NH3)5]2+ (Cpx04) [52,55], and [Co(NCS)(NH3)5]2+ (Cpx05) [52,55]—which present experimental data for Co-59 NMR chemical shift and structure (X-ray). The [Co(CN)6]3− complex (Ref) [56] was also selected because it is the internal reference in Co-59 NMR measurements [52].
The geometries of the Co(III) complexes were optimized and characterized as minimum on the potential energy surface (PES) through harmonic frequency calculations (all frequencies real). For the geometry optimization and NMR calculation, the solvent effects (the same used in the experiments) were included, using the Integral Equation Formalism for the Polarizable Continuum Model (IEF-PCM), with the Radii set from the UFF force field [57]. The Co-59 shielding constant (σ59Co) was calculated using the Gauge-Independent Atomic Orbital (GIAO) [58,59,60,61,62] approach and the Co-59 NMR chemical shift (δ59Co) was calculated according to Equation (3) [20]:
δ C 59 o = σ ref σ calc
where σref is the calculated shielding constant for the [Co(CN)6]3− (internal reference) in D2O and σcalc is the shielding constant calculated for the Co(III) complex under study.
The benchmarking scheme starts with the structures for the six Co(III) complexes (Figure 1) where the DFT-Functional/def2-SVP(Co)/def2-SVP(Ligands)/IEF-PCM(UFF) calculations were evaluated. A set of 21 DFT functionals were tested: BP86 [63,64], BLYP [63,65,66], PBE [67,68], PW91 [67,69,70], M06-L [71], TPSS [72], BB95 [63,73], B3PW91 [67,69,70,74], B3LYP [65,74,75], PBE0 [76], BHANDHLYP [77], M06 [78], M06-2X [78], TPSSh [72,79,80], B1B95 [73], BMK [81], LC-BLYP [63,65,66,82], LC-ωPBE [82,83,84,85], CAM-B3LYP [86], ωB97xD [87], and B97D3 [88]. In the next step, the Co-59 NMR chemical shift was calculated at GIAO-PBE/NMR-DKH/IEF-PCM(UFF) for each geometry obtained at distinct DFT levels. The DFT level that led to the best NMR agreement with experimental results, was selected for geometry optimization of all complexes. Once the protocol for geometry was defined, the DFT functional for predicting the Co-59 NMR chemical shift was assessed at GIAO-DFT-Functional/NMR-DKH/IEF-PCM(UFF), considering the same 21 DFT functionals previously tested for geometries. In total, 378 calculations were evaluated (Figure S1).
The calculations were performed in GAUSSIAN 16 Rev. C.01 program [48].

2.3. Validation of the Best Computational Protocol

The best computational protocol (labeled as Model 1) was applied for prediction of the Co-59 NMR chemical shift (Equation (3)) of other 29 Co(III) complexes. Thus, included in the study was a total of 34 Co(III) complexes (29 used in the validation + 5 used in the benchmarking), with a Co-59 NMR chemical shift ranging from +1162 ppm to +15,100 ppm, and six distinct solvents (water—H2O, dimethylsulfoxide—DMSO, acetonitrile—MeCN, methanol—MeOH, acetone, chloroform—CHCl3, and benzene). Model 1 was also applied for predicting the Co-59 NMR chemical shift of four Co(III) complexes that provided experimental data in five distinct solvents (H2O, formic acid—FA, DMSO, n,n-dimethylformamide—DMF, and MeOH). It is important to point out that almost all of the experimental data for the Co-59 NMR chemical shift of Co(III) complexes were obtained in water. Some data obtained in polar organic solvents are available, but data in non-polar organic solvents are very scarce.

3. Results and Discussion

3.1. Benchmarking the Computational Protocols

(a)
The structures of Co(III) complexes
Considering that the NMR properties are very sensitive to the structure, an evaluation of the DFT functional in the geometry and in the δ59Co was conducted. The def2-SVP basis sets, which showed good results for geometries of transition metal complexes [29,30,31,32], were considered for both cobalt and ligand atoms.
The performance of each computational protocol was assessed with the MRD and MAD calculated by Equations (4) and (5), respectively. These are:
M R D = 1 n k k = 1 n k R D i , j / k R D i , j / k = s i e x p t s i , j / k c a l c s i e x p t × 100 %
and
M A D = 1 n k k = 1 n k A D i , j / k A D i , j / k = s i e x p t s i , j / k c a l c
where RD and AD correspond to the relative and absolute deviation, respectively, i is the considered property (structural parameter or chemical shift), j is the protocol, and k is the Co complex under consideration. For example, s _ ( Co N , B 3 LYP / Cpx 01 )   corresponds to the Co N bond length calculated at B3LYP/def2-SVP(Co)/def2-SVP(Ligands)/IEF-PCM(UFF) protocol for complex 1 (Cpx 01—[Co(NH3)6]3+).
Tables S1–S6 show the calculated values for the bond lengths and bond angles with different DFT-Functionals for the six complexes in Figure 1. A total of 10 Co—L (L = ligand atom) bonds and 27 L—Co—L angles were evaluated. The MRD varied between 0.34% (PBE and BB95) and 1.28% (M06-2X) for the Ref—[Co(CN)6]3−, 1.18% (LC-ωPBE) and 2.02% (BLYP) for the Cpx01—[Co(NH3)6]3+, 1.09% (M06) and 1.96% (BLYP) for the Cpx02—[CoCl(NH3)5]2+, 0.82% (BB95 and TPSSh) and 1.72% (LC-ωPBE) for the Cpx03—[Co(NO2)(NH3)5]2+, 1.06% (B3PW91) and 2.57% (M06-2X) for the Cpx04—[Co(SCN)(NH3)5]2+, and 0.97% (BHandHLYP) and 3.77% (B97D3) for the Cpx05—[Co(NCS)(NH3)5]2+. From Tables S1–S6, it can be seen that all calculated bond lengths have an RD < 5%, with the highest RD = 4.84% for the Co–SCN bond ([Co(NH3)5(SCN)]2+) with the BLYP and TPSS functionals. Regarding the bond angles, all calculated L–Co–L angles had an RD < 3%. The largest deviations were found for the Co–N2–C angle in the Cpx05, when a GGA or meta-GGA functional is considered, with an RD between 10% and 18%, approximately.
For all 10 bond lengths evaluated, the MRD varied between 1.15% (LC-BLYP) and 2.40% (BLYP), and for all 27 bond angles, the MRD varied between 1.04% (TPSSh) and 1.82% (B97D3) (Figure 2). Considering all 37 structural parameters, the MRD ranged between 1.08% (B3PW91) and 1.77% (M06-L) (Figure 2). Although a direct comparison between the experimental structural data, which are obtained in solid-state, and the calculated data, which are obtained in solution, is not the most appropriate, the analysis carried out is important, in that it demonstrates that the calculated structures are well-described and are not very sensitive to the used DFT functional. Thus, the best protocol chosen to describe the structure will be based on the property of interest, i.e., the δ59Co.
For the δ59Co (Table 1), we observed large variations with small changes in the geometries. From Figure 3, the sensitivity of the chemical shift and the geometry can be seen. Although the calculated absolute deviations (AD) for the δ59Co are greater than 1000 ppm, the smallest MADs are found for geometries optimized with pure GGA or meta-GGA functionals. The BLYP geometries gave the smallest MAD, 1573 ppm (MRD = 19.0%). It is important to bear in mind that all DFT functionals presented an excellent description of the structural parameters, with an MRD for the BLYP functional of 1.72%.
If the calculated δ59Co for each Co(III) complex is evaluated separately, the geometries of BLYP/def2-SVP/def2-SVP/IEF-PCM(UFF) presented the lowest RD for all complexes: 18.1% for Cpx01—[Co(NH3)6]3+, 19.9% for Cpx02—[Co(NH3)5Cl]2+, 21.0% for Cpx03—[Co(NH3)5(NO2)]2+, 19.2% for Cpx04—[Co(NH3)5(SCN)]2+, and 16.9% for Cpx05—[Co(NH3)5(NCS)]2+. Therefore, the protocol BLYP/def2-SVP/def2-SVP/IEF-PCM(UFF) was chosen as the most suitable for the geometry optimization of Co(III) complexes. It is worth clarifying that BLYP geometries agree with X-ray data within 1.72%, which is the second-worst performance among the 21 DFT functionals tested (Figure 2). Nonetheless, BLYP geometries gave the best NMR agreement with the experiment. This apparent paradox might be understood if we realized that the geometries were optimized in solution and compared to the solid-state data (X-ray), some differences are expected. On the other hand, the NMR was calculated in solution and compared to the solution experimental data, therefore, the NMR calculation is more suitable to set the best geometry, as mentioned previously.
(b)
Protocol for predicting the Co-59 NMR chemical shift (δ59Co)
The role of the DFT functional to predict δ59Co was assessed at GIAO-DFT-Functional/NMR-DKH/IEF-PCM(UFF)//BLYP/def2-SVP/def2-SVP/IEF-PCM(UFF), considering the same set of 21 DFT functionals used previously. The BLYP/def2-SVP/def2-SVP/IEF-PCM(UFF) optimized geometries were considered in this benchmarking. The calculated values of δ59Co for the Co(III) complexes (Figure 1) are presented in Table 2. The results show that for Cpx01 ([Co(NH3)6]3+), the lowest AD (37 ppm) is obtained with the CAM-B3LYP functional. The LC-BLYP functional has the lowest AD for Cpx02 ([Co(NH3)5Cl]2+) and Cpx04 ([Co(NH3)5(SCN)]2+) (10 ppm and 27 ppm, respectively). The LC-ωPBE functional gave the lowest AD for Cpx03 ([Co(NH3)5(NO2)]2+) and Cpx05 ([Co(NH3)5(NCS)]2+) (60 ppm and 13 ppm, respectively).
From the MAD analysis (Table 2), it is observed that δ59Co is also very sensitive to the DFT functional, with MAD varying between 49 ppm (LC-ωPBE) and 19,049 ppm (M06-2X). The GGA and meta-GGA functionals showed MAD ranging from 1631 ppm (BB95) to 3499 ppm (M06-L). It is interesting to note that when the Minnesota functionals are considered, a significant increase in MAD is observed, with an increase in the % of HF exchange of 3499 ppm for the M06-L (0% of HF exchange), 6488 ppm for the M06 (27% of HF exchange), and reaching 19,049 ppm with the M06-2X (54% of HF exchange). The BMK (42% of HF exchange) and BHandHLYP (50% of HF exchange) also showed high MAD, 5680 ppm and 3436 ppm, respectively. Only six DFT functionals had an MAD below 1000 ppm: the hybrid functionals B3LYP (364 ppm) and B3PW91 (292 ppm), and the long-range (LR) corrected functionals ωB97xD (558 ppm), CAM-B3LYP (106 ppm), LC-BLYP (93 ppm), and LC-ωPBE (49 ppm). This indicates that LR correction plays an important role in predicting δ59Co. Considering the MRD, the six best DFT functionals showed an MRD < 5%, with the LC-ωPBE functional presenting an MRD = 0.6%.
Finally, in order to check the performance of the protocols for all Co(III) complexes, the standard deviation (SD) of the AD was also evaluated (Table 2). The lowest SD (30 ppm) was obtained for the LC-ωPBE functional, indicating that the GIAO-LC-ωPBE/NMR-DKH/IEF-PCM(UFF)//BLYP/def2-SVP/def2-SVP/IEF-PCM(UFF) protocol, named as Model 1, presents a similar accuracy for all complexes and is, therefore, the best choice for predicting δ59Co in Co(III) complexes.

3.2. Validation of the Computational Protocol

The best computational protocol, namely, GIAO-LC-ωPBE/NMR-DKH/IEF-PCM(UFF)//BLYP/def2-SVP/IEF-PCM(UFF)—Model 1, was applied to a new set of 29 Co(III) complexes, not included in the initial set, aiming at validating the protocol. Then, a total of 34 Co(III) complexes were studied in the present paper, with a wide range of δ59Co, from +1162 ppm to +15,100 ppm. It should also be noted that among the 34 complexes studied, 22 had their δ59Co measured in water (H2O), 10 in polar organic solvents (DMSO, MeOH, MeCN, or acetone), and two in non-polar organic solvents (CHCl3 and benzene).
The calculated δ59Co with Model 1 are presented in Table 3 where the MAD and MRD for all 34 Co(III) complexes were 158 ppm and 3.0%, respectively. Figure 4 shows the correlation between the experimental and calculated δ59Co with Model 1 including all 34 Co(III) complexes studied. The coefficient of determination (R2) of 0.9966, the slope of 0.9837 ± 0.0102, and the y-intercept of 67.4421 ± 78.7321 illustrate the quality and the predictive capacity of the proposed computational protocol.
It is interesting to note the quality of Model 1 for the structural characterization of Co(III) complexes. The Cpx04—[Co(NH3)5(SCN)]2+) and Cpx05—[Co(NH3)5(NCS)]2+) present linkage isomerism, with the Co-59 nucleus providing more deshielding in Cpx04, a trend that was well described by Model 1. Conversely, Cpx07—[Co(NH3)5(HIm)]3+ and Cpx08—[Co(NH3)5(MeIm)]3+ present the only difference in the replacement of an -NH by a -NCH3 group in imidazole ligand, resulting in a change of only 7 ppm in the experimental δ59Co. The values calculated with Model 1 also adequately described this trend with a difference of only 33 ppm between the calculated values.
Among the 34 Co(III) complexes included in the present work, six were recently studied by Samultsev et al. [27] using the 4c-Rel approximation. Our calculated results show that the values predicted by Model 1 (20 ≤ AD ≤ 147 ppm) were better than those predicted in ref. [27] for all six complexes (see Table 4).
In a separate analysis to examine the impact of the solvent used, the 22 complexes in water presented a MAD of 156 ppm and a MRD of 3.2%, while the 10 complexes in polar organic solvents (DMSO, MeCN, MeOH, or acetone) showed MAD and MRD of only 139 ppm and 2.6%, respectively. For the two complexes studied in non-polar organic solvents, an MAD of 274 ppm with an MRD of only 2.2% were obtained. This result shows that although Model 1 was obtained from a benchmarking considering complexes (Cpx01 to 05) studied in water, its application is valid for other distinct solvents used within the PCM approach.
Furthermore, for five of the 34 complexes studied, Cpx07, Cpx08, Cpx11—trans-[Co(en)2(N3)2]+, Cpx12—trans-[Co(en)2Cl2]+, and Cpx13—trans-[Co(en)2(NO2)2]+, experimental data for δ59Co are available in more than one solvent. Experimental data show that changing the solvent generates a small variation in δ59Co, with the largest difference observed, 79 ppm (~1%), between the values in H2O and DMSO for Cpx11. Model 1 was applied to these five complexes considering the different solvents (Table 4). For Cpx07 and Cpx08, with H2O or MeOH, it is experimentally observed that there is a small increase in δ59Co for both complexes when MeOH is considered, which is predicted with Model 1 for both complexes. For Cpx11 (data in H2O, FA, DMSO, and MeOH), Cpx12 (data in H2O, DMSO, and MeOH), and Cpx13 (data in H2O, DMSO, DMF, MeOH, and MeCN), the absolute deviations found with Model 1 considering the different solvents varied between 114 ppm and 191 ppm for Cpx11, 371 ppm and 446 ppm for Cpx12, and 1 ppm and 72 ppm for Cpx13. Even for Cpx12, which had a higher AD, the trend of the calculated values followed the same trend observed with the solvent variation. Therefore, Model 1 was able to adequately describe δ59Co, regardless of the considered solvent.
Model 1 was applied to cobaloximes (Cpx27 to Cpx32), which are important Co(III) compounds used as a model for vitamin B12 in studies of their properties and mechanisms of action. The six studied cobaloximes present δ59Co between 3270 ppm and 5371 ppm, with experimental values obtained in DMSO (Cpx27 and Cpx28), H2O (Cpx29 and Cpx30), and acetone (Cpx31 and Cpx32). The calculated δ59Co values (Table 3) for the cobaloximes also showed good agreement with the experimental values, with an AD varying between 124 ppm and 333 ppm, corresponding to an MAD of 189 ppm and an MRD of 4.6%. In addition, Model 1 was able to adequately describe the trend of δ59Co as the axial cobaloximes ligands are changed (Figure 5).

4. Concluding Remarks

The present study aimed to propose a computational protocol based on the DFT level to calculate the Co-59 NMR chemical shift (δ59Co). An initial set of five Co(III) complexes and the internal reference in Co-59 NMR were selected for the DFT benchmarking, which included 21 DFT functionals and the basis set def2-SVP for all atoms. The geometries were optimized at all DFT-Functionals/def2-SVP/IEF-PCM(UFF) levels and further used for the prediction of δ59Co at GIAO-PBE/NMR-DKH/IEF-PCM(UFF). Note that these two first steps included 252 calculations, from which the best scheme was selected: GIAO-PBE/NMR-DKH/IEF-PCM(UFF)//BLYP/def2-SVP/def2-SVP/IEF-PCM(UFF). The MAD and MRD for this protocol were 1573 ppm and 19.0%, respectively. With the aim of improving the protocol, the 21 DFT functionals were used to calculate δ59Co, instead of PBE (126 calculations). The results demonstrated an important role of the long-range correction to the δ59Co values, with the LC-ωPBE leading to the best agreement with the experimental data, MAD = 30 ppm and MRD = 0.5%. The final protocol was labeled as Model 1: GIAO-LC-ωPBE/NMR-DKH/IEF-PCM(UFF)//BLYP/def2-SVP/IEF-PCM(UFF).
Model 1 was applied to a new set of 29 Co(III) complexes, not included in the original set. Considering all 34 complexes, that present experimental data in water and in organic solvents (polar and non-polar), the δ59Co varied between +1162 ppm and +15,110 ppm. The MAD, MRD, and R2 were 158 ppm, 3.0%, and 0.9966, respectively. Lastly, for five of the 34 complexes studied, which present experimental data in different solvents (H2O, FA, DMSO, DMF, MeOH, and MeCN), Model 1 was also able to adequately describe the δ59Co.
The results obtained in the present study suggest Model 1 as an excellent alternative to calculate the δ59Co in Co(III) complexes, with an absolute error that is low enough to assign Co(III) complex structures.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/magnetochemistry9070172/s1, Table S1. Calculated bond lengths ( ) and bond angles (°) for Ref—[Co(CN)6]3− at DFT-Functional/def2-SVP/def2-SVP/IEF-PCM(UFF) level. The mean relative deviation (MRD) in relation to the experimental X-ray values considering all structural parameters evaluated is also presented. Table S2. Calculated bond lengths ( ) and bond angles (°) for Cpx01—[Co(NH3)6]3+ at DFT-Functional/def2-SVP/def2-SVP/IEF-PCM(UFF) level. The mean relative deviation (MRD) in relation to the experimental X-ray values considering all structural parameters evaluated is also presented. Table S3. Calculated bond lengths ( ) and bond angles (°) for Cpx02—[Co(NH3)5Cl]2+ at DFT-Functional/def2-SVP/def2-SVP/IEF-PCM(UFF) level. The mean relative deviation (MRD) in relation to the experimental X-ray values considering all structural parameters evaluated is also presented. Table S4. Calculated bond lengths ( ) and bond angles (°) for Cpx03—[Co(NH3)5(NO2)]2+ at DFT-Functional/def2-SVP/def2-SVP/IEF-PCM(UFF) level. The mean relative deviation (MRD) in relation to the experimental X-ray values considering all structural parameters evaluated is also presented. Table S5. Calculated bond lengths ( ) and bond angles (°) for Cpx04—[Co(NH3)5(SCN)]2+ at DFT-Functional/def2-SVP/def2-SVP/IEF-PCM(UFF) level. The mean relative deviation (MRD) in relation to the experimental X-ray values considering all structural parameters evaluated is also presented. Table S6. Calculated bond lengths ( ) and bond angles (°) for Cpx05—[Co(NH3)5(NCS)]2+ at DFT-Functional/def2-SVP/def2-SVP/IEF-PCM(UFF) level. The mean relative deviation (MRD) in relation to the experimental X-ray values considering all structural parameters evaluated is also presented. Figure S1. Benchmarking flowchart applied to obtain Model 1. Figure S2. 3D structure (with labels) of Co(III) complexes considered in the initial set of the benchmarking. Ref—[Co(CN)6]3−, Cpx01—[Co(NH3)6]3+, Cpx02—[Co(NH3)5Cl]2+, Cpx03—[Co(NH3)5(NO2)]2+, Cpx04—[Co(NH3)5(SCN)]2+, and Cpx05—[Co(NH3)5(NCS)]2+. NMR-DKH basis set for the Co atom. Optimized structures (.xyz format) at BLYP/def2-SVP/def2-SVP/def2-SVP/IEF-PCM(UFF) level of the 34 Co(III) complexes and internal reference complex.

Author Contributions

Conceptualization, M.G.R.G. and D.F.S.P.; Methodology, M.G.R.G. and D.F.S.P.; Software, H.F.D.S. and D.F.S.P.; Validation, M.G.R.G. and D.F.S.P.; Formal Analysis, M.G.R.G., H.F.D.S., W.B.D.A. and D.F.S.P.; Investigation, M.G.R.G. and D.F.S.P.; Resources, H.F.D.S. and D.F.S.P.; Writing, M.G.R.G., A.L.F.D.S., H.F.D.S., W.B.D.A. and D.F.S.P.; Supervision, A.L.F.D.S. and D.F.S.P.; Project Administration, D.F.S.P. All authors have read and agreed to the published version of the manuscript.

Funding

DFSP and MGRG would like to thank the Brazilian agency FAPERJ (E-26/201.141/2019—BOLSA, E-26/010.002261/2019—EMERGENTES, E-26/210.070/2022—DCTR, and E-26/201.336/2022—BOLSA) for the financial support. This study was financed in part by the Coordenação de Aperfeiçoamento de Pessoal de Nível Superior—Brasil (CAPES)—Finance Code 001. HFDS thanks the FAPEMIG (APQ-00591-15) and CNPq (307018/2021-0) for continuing support to the NEQC laboratory.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data will be made available on request.

Acknowledgments

DFSP thanks the FAPERJ for supporting to the NQTCM laboratory.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Delehanty, J.B.; Bongard, J.E.; Thach, D.C.; Knight, D.A.; Hickey, T.E.; Chang, E.L. Antiviral Properties of Cobalt(III)-Complexes. Bioorg. Med. Chem. 2008, 16, 830–837. [Google Scholar] [CrossRef] [PubMed]
  2. Mjos, K.D.; Orvig, C. Metallodrugs in Medicinal Inorganic Chemistry. Chem. Rev. 2014, 114, 4540–4563. [Google Scholar] [CrossRef]
  3. Anthony, E.J.; Bolitho, E.M.; Bridgewater, H.E.; Carter, O.W.L.; Donnelly, J.M.; Imberti, C.; Lant, E.C.; Lermyte, F.; Needham, R.J.; Palau, M.; et al. Metallodrugs Are Unique: Opportunities and Challenges of Discovery and Development. Chem. Sci. 2020, 11, 12888–12917. [Google Scholar] [CrossRef]
  4. Chang, E.L.; Simmers, C.; Knight, D.A. Cobalt Complexes as Antiviral and Antibacterial Agents. Pharmaceuticals 2010, 3, 1711–1728. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Heffern, M.C.; Yamamoto, N.; Holbrook, R.J.; Eckermann, A.L.; Meade, T.J. Cobalt Derivatives as Promising Therapeutic Agents. Curr. Opin. Chem. Biol. 2013, 17, 189–196. [Google Scholar] [CrossRef] [Green Version]
  6. Tsiliou, S.; Kefala, L.-A.; Hatzidimitriou, A.G.; Kessissoglou, D.P.; Perdih, F.; Papadopoulos, A.N.; Turel, I.; Psomas, G. Cobalt(II) Complexes with Non-Steroidal Anti-Inflammatory Drugs and α-Diimines. J. Inorg. Biochem. 2016, 160, 125–139. [Google Scholar] [CrossRef]
  7. Perontsis, S.; Dimitriou, A.; Fotiadou, P.; Hatzidimitriou, A.G.; Papadopoulos, A.N.; Psomas, G. Cobalt(II) Complexes with the Non-Steroidal Anti-Inflammatory Drug Diclofenac and Nitrogen-Donor Ligands. J. Inorg. Biochem. 2019, 196, 110688. [Google Scholar] [CrossRef]
  8. Khan, M.; Khan, N.; Ghazal, K.; Shoaib, S.; Ali, I.; Rauf, M.K.; Badshah, A.; Tahir, M.N.; Rehman, A.-U. Synthesis, Characterization, Crystal Structure, in-Vitro Cytotoxicity, Antibacterial, and Antifungal Activities of Nickel(II) and Cobalt(III) Complexes with Acylthioureas. J. Coord. Chem. 2020, 73, 1790–1805. [Google Scholar] [CrossRef]
  9. Sobiesiak, M.; Cieślak, M.; Królewska, K.; Kaźmierczak-Barańska, J.; Pasternak, B.; Budzisz, E. Thiosemicarbazone-Derived Copper(II), Cobalt(II) and Nickel(II) Complexes as Potential Anticancer Agents: Nuclease Activity, Cytotoxicity and Apoptosis Studies. New J. Chem. 2016, 40, 9761–9767. [Google Scholar] [CrossRef]
  10. Munteanu, C.R.; Suntharalingam, K. Advances in Cobalt Complexes as Anticancer Agents. Dalt. Trans. 2015, 44, 13796–13808. [Google Scholar] [CrossRef]
  11. King, A.P.; Gellineau, H.A.; Ahn, J.-E.; MacMillan, S.N.; Wilson, J.J. Bis(Thiosemicarbazone) Complexes of Cobalt(III). Synthesis, Characterization, and Anticancer Potential. Inorg. Chem. 2017, 56, 6609–6623. [Google Scholar] [CrossRef] [PubMed]
  12. Bühl, M. Chapter 3. DFT Computations of Transition-Metal Chemical Shifts. Annu. Rep. NMR Spectrosc. 2008, 64, 77–126. [Google Scholar] [CrossRef]
  13. Ronconi, L.; Sadler, P.J. Applications of Heteronuclear NMR Spectroscopy in Biological and Medicinal Inorganic Chemistry. Coord. Chem. Rev. 2008, 252, 2239–2277. [Google Scholar] [CrossRef] [PubMed]
  14. Zou, T.; Sadler, P.J. Speciation of Precious Metal Anti-Cancer Complexes by NMR Spectroscopy. Drug Discov. Today Technol. 2015, 16, 7–15. [Google Scholar] [CrossRef] [Green Version]
  15. Harris, R.K.; Becker, E.D.; Cabral de Menezes, S.M.; Goodfellow, R.; Granger, P. NMR Nomenclature: Nuclear Spin Properties and Conventions for Chemical Shifts. Solid State Nucl. Magn. Reson. 2002, 22, 458–483. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Chan, J.C.C.; Au-Yeung, S.C.F. Cobalt-59 NMR Spectroscopy. Annu. Rep. NMR Spectrosc. 2000, 41, 1–54. [Google Scholar] [CrossRef]
  17. Yamasaki, A. Cobalt-59 Nuclear Magnetic Resonance Spectroscopy in Coordination Chemistry. J. Coord. Chem. 1991, 24, 211–260. [Google Scholar] [CrossRef]
  18. Yamasaki, A.; Yajima, F.; Fujiwara, S. Nuclear Magnetic Resonance Studies on Cobalt Complexes. I. Cobalt-59 Nuclear Magnetic Resonance Spectra of Cobalt(III) Complexes. Inorg. Chim. Acta 1968, 2, 39–42. [Google Scholar] [CrossRef]
  19. Barnard, I.; Koch, K.R. 59Co NMR, a Facile Tool to Demonstrate EEE, EEZ, EZZ and ZZZ Configurational Isomerism in Fac-[Co(L-ΚS,O)3] Complexes Derived from Asymmetrically Substituted N,N-Dialkyl-N′-Aroylthioureas. Inorg. Chim. Acta 2019, 495, 119019. [Google Scholar] [CrossRef]
  20. Rusakova, I.L. Quantum Chemical Approaches to the Calculation of NMR Parameters: From Fundamentals to Recent Advances. Magnetochemistry 2022, 8, 50. [Google Scholar] [CrossRef]
  21. Chan, J.C.C.; Au-Yeung, S.C.F.; Wilson, P.J.; Webb, G.A. SOS-DFPT-IGLO Calculations of 59Co NMR Shielding Parameters of Hexacoordinated Diamagnetic Co(III) Complexes. J. Mol. Struct. THEOCHEM 1996, 365, 125–130. [Google Scholar] [CrossRef]
  22. Chan, J.C.C.; Au-Yeung, S.C.F. A Comparative Study of the Calculation of 59Co NMR Shielding Constants of Hexacoordinated Diamagnetic Co(III) Complexes Using DFT-IGLO and DFT-GIAO Methods. J. Mol. Struct. THEOCHEM 1997, 393, 93–96. [Google Scholar] [CrossRef]
  23. Godbout, N.; Oldfield, E. Density Functional Study of Cobalt-59 Nuclear Magnetic Resonance Chemical Shifts and Shielding Tensor Elements in Co(III) Complexes. J. Am. Chem. Soc. 1997, 119, 8065–8069. [Google Scholar] [CrossRef]
  24. Bühl, M.; Grigoleit, S.; Kabrede, H.; Mauschick, F.T. Simulation of 59Co NMR Chemical Shifts in Aqueous Solution. Chem. A Eur. J. 2006, 12, 477–488. [Google Scholar] [CrossRef]
  25. Semenov, V.A.; Samultsev, D.O.; Krivdin, L.B. Four-Component Relativistic Computational NMR Study of Ferrous, Cobalt and Nickel Bisglycinates. Mendeleev Commun. 2020, 30, 476–478. [Google Scholar] [CrossRef]
  26. Samultsev, D.O.; Semenov, V.A.; Krivdin, L.B. Four-component Relativistic Calculations of NMR Shielding Constants of the Transition Metal Complexes. Part 1: Pentaammines of Cobalt, Rhodium, and Iridium. Magn. Reson. Chem. 2022, 60, 463–468. [Google Scholar] [CrossRef] [PubMed]
  27. Samultsev, D.O.; Semenov, V.A.; Rusakova, I.L.; Krivdin, L.B. Four-Component Relativistic Calculations of NMR Shielding Constants of the Transition Metal Complexes—Part 2: Nitrogen-Coordinated Complexes of Cobalt. Int. J. Mol. Sci. 2022, 23, 13178. [Google Scholar] [CrossRef]
  28. Samultsev, D.O.; Semenov, V.A.; Krivdin, L.B. Four-Component Relativistic Calculations of NMR Shielding Constants of the Transition Metal Complexes—Part 3: Fe, Co, Ni, Pd, and Pt Glycinates. Magnetochemistry 2023, 9, 83. [Google Scholar] [CrossRef]
  29. Paschoal, D.; Guerra, C.F.; de Oliveira, M.A.L.; Ramalho, T.C.; Dos Santos, H.F. Predicting Pt-195 NMR Chemical Shift Using New Relativistic All-Electron Basis Set. J. Comput. Chem. 2016, 37, 2360–2373. [Google Scholar] [CrossRef]
  30. Carvalho, J.; Paschoal, D.; Fonseca Guerra, C.; Dos Santos, H.F. Nonrelativistic Protocol for Calculating the 1J(195Pt-15N) Coupling Constant in Pt(II)-Complexes Using All-Electron Gaussian Basis-Set. Chem. Phys. Lett. 2020, 745, 137279. [Google Scholar] [CrossRef]
  31. E Silva, J.H.C.; Dos Santos, H.F.; Paschoal, D.F.S. Predicting Pt-195 Nmr Chemical Shift and 1J(195Pt-31P) Coupling Constant for Pt(0) Complexes Using the NMR-DKH Basis Sets. Magnetochemistry 2021, 7, 148. [Google Scholar] [CrossRef]
  32. de Andrade, T.F.C.B.; Dos Santos, H.F.; Fonseca Guerra, C.; Paschoal, D.F.S. Computational Prediction of Tc-99 NMR Chemical Shifts in Technetium Complexes with Radiopharmaceutical Applications. J. Phys. Chem. A 2022, 126, 5434–5448. [Google Scholar] [CrossRef] [PubMed]
  33. Paschoal, D.F.S.; Dos Santos, H.F. Predicting the Structure and NMR Coupling Constant 1J(129Xe-19F) of XeF6 Using Quantum Mechanics Methods. Phys. Chem. Chem. Phys. 2021, 23, 7240–7246. [Google Scholar] [CrossRef]
  34. Rolfes, J.D.; Neese, F.; Pantazis, D.A. All-electron Scalar Relativistic Basis Sets for the Elements Rb–Xe. J. Comput. Chem. 2020, 41, 1842–1849. [Google Scholar] [CrossRef] [PubMed]
  35. Aravena, D.; Neese, F.; Pantazis, D.A. Improved Segmented All-Electron Relativistically Contracted Basis Sets for the Lanthanides. J. Chem. Theory Comput. 2016, 12, 1148–1156. [Google Scholar] [CrossRef] [PubMed]
  36. Pantazis, D.A.; Neese, F. All-Electron Scalar Relativistic Basis Sets for the 6p Elements. Theor. Chem. Acc. 2012, 131, 1292. [Google Scholar] [CrossRef]
  37. Pantazis, D.A.; Neese, F. All-Electron Scalar Relativistic Basis Sets for the Actinides. J. Chem. Theory Comput. 2011, 7, 677–684. [Google Scholar] [CrossRef]
  38. Pantazis, D.A.; Neese, F. All-Electron Scalar Relativistic Basis Sets for the Lanthanides. J. Chem. Theory Comput. 2009, 5, 2229–2238. [Google Scholar] [CrossRef]
  39. Pantazis, D.A.; Chen, X.-Y.; Landis, C.R.; Neese, F. All-Electron Scalar Relativistic Basis Sets for Third-Row Transition Metal Atoms. J. Chem. Theory Comput. 2008, 4, 908–919. [Google Scholar] [CrossRef]
  40. Dyall, K.G.; Grant, I.P.; Johnson, C.T.; Parpia, F.A.; Plummer, E.P. GRASP: A General-Purpose Relativistic Atomic Structure Program. Comput. Phys. Commun. 1989, 55, 425–456. [Google Scholar] [CrossRef]
  41. Douglas, M.; Kroll, N.M. Quantum Electrodynamical Corrections to the Fine Structure of Helium. Ann. Phys. 1974, 82, 89–155. [Google Scholar] [CrossRef]
  42. Hess, B.A. Applicability of the No-Pair Equation with Free-Particle Projection Operators to Atomic and Molecular Structure Calculations. Phys. Rev. A 1985, 32, 756–763. [Google Scholar] [CrossRef] [PubMed]
  43. Hess, B.A. Relativistic Electronic-Structure Calculations Employing a Two-Component No-Pair Formalism with External-Field Projection Operators. Phys. Rev. A 1986, 33, 3742–3748. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Jansen, G.; Hess, B.A. Revision of the Douglas-Kroll Transformation. Phys. Rev. A 1989, 39, 6016–6017. [Google Scholar] [CrossRef]
  45. Visscher, L.; Dyall, K.G. Dirac–Fock atomic electronic structure calculations using different nuclear charge distributions. At. Data Nucl. Data Tables 1997, 67, 207–224. [Google Scholar] [CrossRef]
  46. Barysz, M.; Sadlej, A.J. Two-Component Methods of Relativistic Quantum Chemistry: From the Douglas–Kroll Approximation to the Exact Two-Component Formalism. J. Mol. Struct. THEOCHEM 2001, 573, 181–200. [Google Scholar] [CrossRef]
  47. de Jong, W.A.; Harrison, R.J.; Dixon, D.A. Parallel Douglas–Kroll Energy and Gradients in NWChem: Estimating Scalar Relativistic Effects Using Douglas–Kroll Contracted Basis Sets. J. Chem. Phys. 2001, 114, 48. [Google Scholar] [CrossRef]
  48. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Petersson, G.A.; Nakatsuji, H.; et al. Gaussian 16, Revision C.01; Gaussian, Inc.: Wallingford, CT, USA, 2016. [Google Scholar]
  49. Paschoal, D.; Costa, M.F.; Dos Santos, H.F. NLO-X (X = I-III): New Gaussian Basis Sets for Prediction of Linear and Nonlinear Electric Properties. Int. J. Quantum Chem. 2014, 114, 796–804. [Google Scholar] [CrossRef]
  50. Pritchard, B.P.; Altarawy, D.; Didier, B.; Gibson, T.D.; Windus, T.L. New Basis Set Exchange: An Open, Up-to-Date Resource for the Molecular Sciences Community. J. Chem. Inf. Model. 2019, 59, 4814–4820. [Google Scholar] [CrossRef] [PubMed]
  51. Kruger, G.J.; Reynhardt, E.C. Hexaamminecobalt(III) Chloride. Acta Crystallogr. Sect. B Struct. Crystallogr. Cryst. Chem. 1978, 34, 915–917. [Google Scholar] [CrossRef]
  52. Chan, J.C.C.; Au-Yeung, S.C.F. Interpretation of 59 Co NMR Shielding Using the Hard and Soft Acid–Base Concept. Insight into the Relative Magnitude of the Nephelauxetic and the Spectrochemical Effect. J. Chem. Soc. Faraday Trans. 1996, 92, 1121–1128. [Google Scholar] [CrossRef]
  53. Messmer, G.G.; Amma, E.L. Redetermination of the Crystal Structure of Chloropentamminecobalt(III) Dichloride. Acta Crystallogr. Sect. B Struct. Crystallogr. Cryst. Chem. 1968, 24, 417–422. [Google Scholar] [CrossRef]
  54. Cotton, F.A.; Edwards, W.T. The Crystal and Molecular Structure of Nitropentamminocobalt(III) Bromide. Acta Crystallogr. Sect. B Struct. Crystallogr. Cryst. Chem. 1968, 24, 474–477. [Google Scholar] [CrossRef] [Green Version]
  55. Snow, M.R.; Boomsma, R.F. The Crystal Structures and Isomerization of the Linkage Isomers Thiocyanato- and Isothiocyanato-Pentaamminecobalt(III) Dichloride, [Co(SCN)(NH3)5]Cl2.H2O, and [Co(NCS)(NH3)5]Cl2. Acta Crystallogr. Sect. B Struct. Crystallogr. Cryst. Chem. 1972, 28, 1908–1913. [Google Scholar] [CrossRef]
  56. Iwata, M.; Saito, Y. The Crystal Structure of Hexamminecobalt(III) Hexacyanocobaltate(III): An Accurate Determination. Acta Crystallogr. Sect. B Struct. Crystallogr. Cryst. Chem. 1973, 29, 822–832. [Google Scholar] [CrossRef]
  57. Scalmani, G.; Frisch, M.J. Continuous Surface Charge Polarizable Continuum Models of Solvation. I. General Formalism. J. Chem. Phys. 2010, 132, 114110. [Google Scholar] [CrossRef]
  58. London, F. Théorie Quantique Des Courants Interatomiques Dans Les Combinaisons Aromatiques. J. Phys. le Radium 1937, 8, 397–409. [Google Scholar] [CrossRef] [Green Version]
  59. McWeeny, R. Perturbation Theory for the Fock-Dirac Density Matrix. Phys. Rev. 1962, 126, 1028–1034. [Google Scholar] [CrossRef]
  60. Ditchfield, R. Self-Consistent Perturbation Theory of Diamagnetism. Mol. Phys. 1974, 27, 789–807. [Google Scholar] [CrossRef]
  61. Wolinski, K.; Hinton, J.F.; Pulay, P. Efficient Implementation of the Gauge-Independent Atomic Orbital Method for NMR Chemical Shift Calculations. J. Am. Chem. Soc. 1990, 112, 8251–8260. [Google Scholar] [CrossRef]
  62. Cheeseman, J.R.; Trucks, G.W.; Keith, T.A.; Frisch, M.J. A Comparison of Models for Calculating Nuclear Magnetic Resonance Shielding Tensors. J. Chem. Phys. 1996, 104, 5497–5509. [Google Scholar] [CrossRef]
  63. Becke, A.D. Density-Functional Exchange-Energy Approximation with Correct Asymptotic Behavior. Phys. Rev. A 1988, 38, 3098–3100. [Google Scholar] [CrossRef]
  64. Perdew, J.P. Density-Functional Approximation for the Correlation Energy of the Inhomogeneous Electron Gas. Phys. Rev. B 1986, 33, 8822–8824. [Google Scholar] [CrossRef] [PubMed]
  65. Lee, C.; Yang, W.; Parr, R.G. Development of the Colle-Salvetti Correlation-Energy Formula into a Functional of the Electron Density. Phys. Rev. B 1988, 37, 785–789. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Miehlich, B.; Savin, A.; Stoll, H.; Preuss, H. Results Obtained with the Correlation Energy Density Functionals of Becke and Lee, Yang and Parr. Chem. Phys. Lett. 1989, 157, 200–206. [Google Scholar] [CrossRef]
  67. Perdew, J.P.; Burke, K.; Wang, Y. Generalized Gradient Approximation for the Exchange-Correlation Hole of a Many-Electron System. Phys. Rev. B 1996, 54, 16533–16539. [Google Scholar] [CrossRef] [Green Version]
  68. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865, Erratum in Phys. Rev. Lett. 1997, 78, 1396. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  69. Perdew, J.P.; Chevary, J.A.; Vosko, S.H.; Jackson, K.A.; Pederson, M.R.; Singh, D.J.; Fiolhais, C. Atoms, Molecules, Solids, and Surfaces: Applications of the Generalized Gradient Approximation for Exchange and Correlation. Phys. Rev. B 1992, 46, 6671–6687. [Google Scholar] [CrossRef]
  70. Perdew, J.P.; Chevary, J.A.; Vosko, S.H.; Jackson, K.A.; Pederson, M.R.; Singh, D.J.; Fiolhais, C. Atoms, Molecules, Solids, and Surfaces: Applications of the Generalized Gradient Approximation for Exchange and Correlation. Phys. Rev. B 1992, 46, 6671–6687, Erratum in Phys. Rev. B 1993, 48, 4978. [Google Scholar] [CrossRef] [PubMed]
  71. Zhao, Y.; Truhlar, D.G. A New Local Density Functional for Main-Group Thermochemistry, Transition Metal Bonding, Thermochemical Kinetics, and Noncovalent Interactions. J. Chem. Phys. 2006, 125, 194101. [Google Scholar] [CrossRef] [Green Version]
  72. Tao, J.; Perdew, J.P.; Staroverov, V.N.; Scuseria, G.E. Climbing the Density Functional Ladder: Nonempirical Meta–Generalized Gradient Approximation Designed for Molecules and Solids. Phys. Rev. Lett. 2003, 91, 146401. [Google Scholar] [CrossRef] [Green Version]
  73. Becke, A.D. Density-functional Thermochemistry. IV. A New Dynamical Correlation Functional and Implications for Exact-exchange Mixing. J. Chem. Phys. 1996, 104, 1040–1046. [Google Scholar] [CrossRef] [Green Version]
  74. Becke, A.D. Density-functional Thermochemistry. III. The Role of Exact Exchange. J. Chem. Phys. 1993, 98, 5648–5652. [Google Scholar] [CrossRef] [Green Version]
  75. Stephens, P.J.; Devlin, F.J.; Chabalowski, C.F.; Frisch, M.J. Ab Initio Calculation of Vibrational Absorption and Circular Dichroism Spectra Using Density Functional Force Fields. J. Phys. Chem. 1994, 98, 11623–11627. [Google Scholar] [CrossRef]
  76. Adamo, C.; Barone, V. Toward Reliable Density Functional Methods without Adjustable Parameters: The PBE0 Model. J. Chem. Phys. 1999, 110, 6158–6170. [Google Scholar] [CrossRef]
  77. Becke, A.D. A New Mixing of Hartree–Fock and Local Density-functional Theories. J. Chem. Phys. 1993, 98, 1372–1377. [Google Scholar] [CrossRef]
  78. Zhao, Y.; Truhlar, D.G. The M06 Suite of Density Functionals for Main Group Thermochemistry, Thermochemical Kinetics, Noncovalent Interactions, Excited States, and Transition Elements: Two New Functionals and Systematic Testing of Four M06-Class Functionals and 12 Other Function. Theor. Chem. Acc. 2008, 120, 215–241. [Google Scholar] [CrossRef] [Green Version]
  79. Staroverov, V.N.; Scuseria, G.E.; Tao, J.; Perdew, J.P. Comparative Assessment of a New Nonempirical Density Functional: Molecules and Hydrogen-Bonded Complexes. J. Chem. Phys. 2003, 119, 12129–12137. [Google Scholar] [CrossRef]
  80. Staroverov, V.N.; Scuseria, G.E.; Tao, J.; Perdew, J.P. Comparative Assessment of a New Nonempirical Density Functional: Molecules and Hydrogen-Bonded Complexes. J. Chem. Phys. 2003, 119, 12129, Erratum in J. Chem. Phys. 2004, 121, 11507. [Google Scholar] [CrossRef]
  81. Boese, A.D.; Martin, J.M.L. Development of Density Functionals for Thermochemical Kinetics. J. Chem. Phys. 2004, 121, 3405–3416. [Google Scholar] [CrossRef]
  82. Iikura, H.; Tsuneda, T.; Yanai, T.; Hirao, K. A Long-Range Correction Scheme for Generalized-Gradient-Approximation Exchange Functionals. J. Chem. Phys. 2001, 115, 3540–3544. [Google Scholar] [CrossRef]
  83. Vydrov, O.A.; Scuseria, G.E. Assessment of a Long-Range Corrected Hybrid Functional. J. Chem. Phys. 2006, 125, 234109. [Google Scholar] [CrossRef] [PubMed]
  84. Vydrov, O.A.; Heyd, J.; Krukau, A.V.; Scuseria, G.E. Importance of Short-Range versus Long-Range Hartree-Fock Exchange for the Performance of Hybrid Density Functionals. J. Chem. Phys. 2006, 125, 074106. [Google Scholar] [CrossRef]
  85. Vydrov, O.A.; Scuseria, G.E.; Perdew, J.P. Tests of Functionals for Systems with Fractional Electron Number. J. Chem. Phys. 2007, 126, 154109. [Google Scholar] [CrossRef]
  86. Yanai, T.; Tew, D.P.; Handy, N.C. A New Hybrid Exchange–Correlation Functional Using the Coulomb-Attenuating Method (CAM-B3LYP). Chem. Phys. Lett. 2004, 393, 51–57. [Google Scholar] [CrossRef] [Green Version]
  87. Chai, J.-D.; Head-Gordon, M. Long-Range Corrected Hybrid Density Functionals with Damped Atom–Atom Dispersion Corrections. Phys. Chem. Chem. Phys. 2008, 10, 6615. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  88. Grimme, S.; Ehrlich, S.; Goerigk, L. Effect of the Damping Function in Dispersion Corrected Density Functional Theory. J. Comput. Chem. 2011, 32, 1456–1465. [Google Scholar] [CrossRef]
  89. Kirby, C.W.; Puranda, C.M.; Power, W.P. Cobalt-59 Nuclear Magnetic Relaxation Studies of Aqueous Octahedral Cobalt(III) Complexes. J. Phys. Chem. 1996, 100, 14618–14624. [Google Scholar] [CrossRef]
  90. Medek, A.; Frydman, V.; Frydman, L. Solid and Liquid Phase 59Co NMR Studies of Cobalamins and Their Derivatives. Proc. Natl. Acad. Sci. USA 1997, 94, 14237–14242. [Google Scholar] [CrossRef]
Figure 1. Co(III) complexes considered in the initial set of the benchmarking. The geometries were optimized at BLYP/def2-SVP/def2-SVP/IEF-PCM(UFF) level: (a) Ref—[Co(CN)6]3−, (b) Cpx01—[Co(NH3)6]3+, (c) Cpx02—[Co(NH3)5Cl]2+, (d) Cpx03—[Co(NH3)5(NO2)]2+, (e) Cpx04—[Co(NH3)5(SCN)]2+, and (f) Cpx05—[Co(NH3)5(NCS)]2+.
Figure 1. Co(III) complexes considered in the initial set of the benchmarking. The geometries were optimized at BLYP/def2-SVP/def2-SVP/IEF-PCM(UFF) level: (a) Ref—[Co(CN)6]3−, (b) Cpx01—[Co(NH3)6]3+, (c) Cpx02—[Co(NH3)5Cl]2+, (d) Cpx03—[Co(NH3)5(NO2)]2+, (e) Cpx04—[Co(NH3)5(SCN)]2+, and (f) Cpx05—[Co(NH3)5(NCS)]2+.
Magnetochemistry 09 00172 g001
Figure 2. Mean relative deviation (MRD, %) for the structural parameters of Co(III) complexes calculated at DFT-Functional/def2-SVP/def2-SVP/IEF-PCM(UFF) level.
Figure 2. Mean relative deviation (MRD, %) for the structural parameters of Co(III) complexes calculated at DFT-Functional/def2-SVP/def2-SVP/IEF-PCM(UFF) level.
Magnetochemistry 09 00172 g002
Figure 3. Mean relative deviation (MRD, %) for the structural parameters of Co(III) complexes calculated at DFT-Functional/def2-SVP/def2-SVP/IEF-PCM(UFF) level, and MRD for the δ59Co calculated at GIAO-PBE/NMR-DKH/IEF-PCM(UFF) level. The DFT functionals in the X-axis refer to the level used for geometry optimization.
Figure 3. Mean relative deviation (MRD, %) for the structural parameters of Co(III) complexes calculated at DFT-Functional/def2-SVP/def2-SVP/IEF-PCM(UFF) level, and MRD for the δ59Co calculated at GIAO-PBE/NMR-DKH/IEF-PCM(UFF) level. The DFT functionals in the X-axis refer to the level used for geometry optimization.
Magnetochemistry 09 00172 g003
Figure 4. Correlation between the experimental and calculated (Model 1) δ59Co (ppm) for all 34 Co(III) complexes studied in the present paper. The level of theory was GIAO-LC-ωPBE/NMR-DKH/IEF-PCM(UFF)//BLYP/def2-SVP/def2-SVP/IEF-PCM(UFF).
Figure 4. Correlation between the experimental and calculated (Model 1) δ59Co (ppm) for all 34 Co(III) complexes studied in the present paper. The level of theory was GIAO-LC-ωPBE/NMR-DKH/IEF-PCM(UFF)//BLYP/def2-SVP/def2-SVP/IEF-PCM(UFF).
Magnetochemistry 09 00172 g004
Figure 5. Calculated δ59Co (ppm) with Model 1, GIAO-LC-ωPBE/NMR-DKH/IEF-PCM(UFF)//BLYP/def2-SVP/def2-SVP/IEF-PCM(UFF), for all six cobaloximes studied in the present paper.
Figure 5. Calculated δ59Co (ppm) with Model 1, GIAO-LC-ωPBE/NMR-DKH/IEF-PCM(UFF)//BLYP/def2-SVP/def2-SVP/IEF-PCM(UFF), for all six cobaloximes studied in the present paper.
Magnetochemistry 09 00172 g005
Table 1. Calculated δ59Co (ppm) at GIAO-PBE/NMR-DKH/IEF-PCM(UFF)//DFT-Functional/def2-SVP/def2-SVP/IEF-PCM(UFF) for Co(III) complexes.
Table 1. Calculated δ59Co (ppm) at GIAO-PBE/NMR-DKH/IEF-PCM(UFF)//DFT-Functional/def2-SVP/def2-SVP/IEF-PCM(UFF) for Co(III) complexes.
DFT-FunctionalCpx01Cpx02Cpx03Cpx04Cpx05MADMRD
GGABP8663076689565163526545197323.8%
BLYP66737092604167976941157319.0%
PBE63686703569562186566197223.8%
PW9162996643562262426493202224.4%
meta-GGAM06-L64766824588865056707180121.7%
TPSS62486567554161916409209025.3%
BB9565066684578364006701186722.5%
HybridB3PW9161206501552961276308216426.1%
B3LYP61676605560062886391207125.0%
PBE059466288536859386139234628.3%
BHANDHLYP54635865487455375683279733.8%
Hybrid meta-GGATPSSh60856443544060706207223227.0%
B1B9559176280533758706099238128.8%
BMK58156305526458746158239829.0%
M0658336207532458276039243529.4%
M06-2X54215732479854725680286134.6%
LR correctedCAM-B3LYP57316097514057325929255630.9%
LC-BLYP53625666476552675521296535.8%
LC-ωPBE57916085718956355949215225.6%
B97D363186530581164476423197623.8%
ωB97xD60306381539159726111230527.8%
Experimental81528852764384108350--
Ref = [Co(CN)6]3−, Cpx01 = [Co(NH3)6]3+, Cpx02 = [Co(NH3)5Cl]2+, Cpx03 = [Co(NH3)5(NO2)]2+, Cpx04 = [Co(NH3)5(SCN)]2+, Cpx05 = [Co(NH3)5(NCS)]2+. Experimental δ59Co values, measured in D2O in relation to the internal reference complex ([Co(CN)6]3− in D2O), obtained from Chan et al. [52]. MAD = mean absolute deviation, Equation (5); MRD = mean relative deviation, Equation (4).
Table 2. Calculated δ59Co (ppm) at GIAO-DFT-Functional/NMR-DKH/IEF-PCM(UFF)//BLYP/def2-SVP/def2-SVP/IEF-PCM(UFF) for Co(III) complexes.
Table 2. Calculated δ59Co (ppm) at GIAO-DFT-Functional/NMR-DKH/IEF-PCM(UFF)//BLYP/def2-SVP/def2-SVP/IEF-PCM(UFF) for Co(III) complexes.
DFT-FunctionalCpx01Cpx02Cpx03Cpx04Cpx05MADSDMRD
GGABP8664676872587166026736177212421.4%
BLYP60546387554561496303219415426.5%
PBE66737092604167976941157312119.0%
PW9165226923593466466780172012420.8%
meta-GGAM06-L47454941449647594971349926142.2%
TPSS54395752506655295663279218233.7%
BB9566177021597167296912163113519.7%
HybridB3PW9183359138782287888759287963.4%
B3LYP77778337728380588106369864.4%
PBE0900110,053897810,0139561124024415.0%
BHANDHLYP10,82812,84710,78412,36011,767343649941.4%
Hybrid meta-GGATPSSh60796546571863176432206314124.9%
B1B95950810,725963410,48310,179182424922.1%
BMK13,13215,01812,73914,85014,068568057468.5%
M0614,19915,83913,52915,20115,079648843778.3%
M06-2X25,26029,62123,32229,96228,46019,0442257229.3%
LR correctedCAM-B3LYP81628732785683778225100731.2%
LC-BLYP8382886277498437828388781.1%
LC-ωPBE8176882677038505833744300.5%
B97D369797426633771417232125810715.2%
ωB97xD860492558314900189975531076.7%
Experimental81528852764384108350---
Ref = [Co(CN)6]3−, Cpx01 = [Co(NH3)6]3+, Cpx02 = [Co(NH3)5Cl]2+, Cpx03 = [Co(NH3)5(NO2)]2+, Cpx04 = [Co(NH3)5(SCN)]2+, Cpx05 = [Co(NH3)5(NCS)]2+. Experimental δ59Co values, measured in D2O in relation to the internal reference complex ([Co(CN)6]3− in D2O), obtained from Chan et al. [52]. MAD = mean absolute deviation, Equation (5); MRD = mean relative deviation, Equation (4). SD = standard deviation of absolute deviations (AD).
Table 3. Calculated δ59Co (ppm) at GIAO-LC-ωPBE/NMR-DKH/IEF-PCM(UFF)//BLYP/def2-SVP/def2-SVP/IEF-PCM(UFF) for the set of 34 Co(III) complexes studied in the present paper. The values shown in square brackets are found in ref. [27] considering a Rel-4c approach.
Table 3. Calculated δ59Co (ppm) at GIAO-LC-ωPBE/NMR-DKH/IEF-PCM(UFF)//BLYP/def2-SVP/def2-SVP/IEF-PCM(UFF) for the set of 34 Co(III) complexes studied in the present paper. The values shown in square brackets are found in ref. [27] considering a Rel-4c approach.
CpxStructureSolventCalc.Expt.CpxStructureSolventCalc.Expt.
01Magnetochemistry 09 00172 i001
[Co(NH3)6]3+
D2O8176
[9223.1]
8152 a18Magnetochemistry 09 00172 i002
[Co(NH3)5(CO3)]+
D2O89239000 b
02Magnetochemistry 09 00172 i003
[Co(NH3)5Cl]2+
D2O8826
[7685.5]
8852 a19Magnetochemistry 09 00172 i004
cis-[Co(NH3)4(CO3)]+
D2O93609662 c
03Magnetochemistry 09 00172 i005
[Co(NH3)5(NO2)]2+
D2O7703
[8409.0]
7643 a20Magnetochemistry 09 00172 i006
[Co(CN)5(NO2)]3−
D2O15661400 b
04Magnetochemistry 09 00172 i007
[Co(NH3)5(SCN)]2+
D2O85058410 a21Magnetochemistry 09 00172 i008
[Co(NH3)(CN)5]2−
D2O10721162 b
05Magnetochemistry 09 00172 i009
[Co(NH3)5(NCS)]2+
D2O83378350 a22Magnetochemistry 09 00172 i010
mer-[Co(CN)3(NH3)3]
D2O39173947 b
06Magnetochemistry 09 00172 i011
[Co(NH3)5(N3)]2+
D2O8695
[9000.3]
8842 a23Magnetochemistry 09 00172 i012
cis-[Co(CN)2(en)2]+
D2O45454364 b
07Magnetochemistry 09 00172 i013
[Co(NH3)5(HIm)]3+
MeOH81498208 a24Magnetochemistry 09 00172 i014
trans-[Co(NH3)4(NO2)Cl]+
D2O78938180 b
08Magnetochemistry 09 00172 i015
[Co(NH3)5(MeIm)]3+
MeOH81828215 b25Magnetochemistry 09 00172 i016
[Co(en)(CN)4]
D2O19342006 b
09Magnetochemistry 09 00172 i017
[CoBr(NH3)6]2+
D2O8939
[9119.4]
8919 b26Magnetochemistry 09 00172 i018
[Co(OH2)6]3+
D2O15,48515,100 b
10Magnetochemistry 09 00172 i019
[Co(NH3)5I]2+
D2O8935
[8639.2]
8849 a27Magnetochemistry 09 00172 i020
trans-[Co(DH)2(CN)(py)]
DMSO43634150 d
11Magnetochemistry 09 00172 i021
trans-[Co(en)2(N3)2]+
DMSO84708350 a28Magnetochemistry 09 00172 i022
trans-[Co(DH)2(CN)2]
DMSO34243270 d
12Magnetochemistry 09 00172 i023
trans-[Co(en)2Cl2]+
DMSO93138870 a29Magnetochemistry 09 00172 i024
trans-[Co(DH)2(CH3)(OH2)]
D2O40694220 d
13Magnetochemistry 09 00172 i025
trans-[Co(en)2(NO2)2]+
DMSO63966395 a30Magnetochemistry 09 00172 i026
trans-[Co(DH)2(NH3)2]
D2O57045371 d
14Magnetochemistry 09 00172 i027
trans-[Co(en)2(NCS)2]+
MeOH79547870 a31Magnetochemistry 09 00172 i028
trans-[Co(DH)2(CH3)(MeIm)]
Acetone37443620 d
15Magnetochemistry 09 00172 i029
trans-[Co(NH3)4(NO2)]+
D2O69037157 c32Magnetochemistry 09 00172 i030
trans-[Co(DH)2(CH3)(py)]
Acetone38073645 d
16Magnetochemistry 09 00172 i031
cis-[Co(NH3)4(NO2)2]+
D2O75097227 c33Magnetochemistry 09 00172 i032
[Co(acac)3]
CHCl312,83512,650 b
17Magnetochemistry 09 00172 i033
fac-[Co(CN)3(NH3)3]
D2O29433289 c34Magnetochemistry 09 00172 i034
[Co(dbzm)3]
Benzene12,89412,530 b
34 Co(III) complexes: MAD = 158 ppm/MRD = 3.0%/R2 = 0.9966
Experimental δ59Co values, measured in relation to the internal reference complex ([Co(CN)6]3− in D2O), obtained from: a Chan et al. [52]; b Yamasaki [17]; c Kirby et al. [89]; d Medek et al. [90]. HIm = imidazole; MeIm = methylimidazole; en = ethylenediamine; py = pyridine; DH = dimethylglioximato; acac = acetylacetonato; dbzm = dibenzoylmethanato.
Table 4. Calculated δ59Co (ppm) at GIAO-LC-ωPBE/NMR-DKH/IEF-PCM(UFF)//BLYP/def2-SVP/def2-SVP/IEF-PCM(UFF) for the five Co(III) complexes that present experimental data in distinct solvents.
Table 4. Calculated δ59Co (ppm) at GIAO-LC-ωPBE/NMR-DKH/IEF-PCM(UFF)//BLYP/def2-SVP/def2-SVP/IEF-PCM(UFF) for the five Co(III) complexes that present experimental data in distinct solvents.
δ59Co (ppm)
CpxCo(III) ComplexesSolventModel 1Expt.AD (ppm)RD (%)
07[Co(NH3)5(HIm)]3+D2O81328170380.5%
MeOH81498208590.7%
08[Co(NH3)5(MeIm)]3+D2O81618178170.2%
MeOH81828215330.4%
11trans-[Co(en)2(N3)2]+D2O847383591141.4%
FA847182801912.3%
DMSO847083501201.4%
MeOH846882991692.0%
12trans-[Co(en)2Cl2]+D2O933189603714.1%
DMSO931388704435.0%
MeOH929688504465.0%
13trans-[Co(en)2(NO2)2]+D2O63966324721.1%
DMSO6396639510.0%
DMF6397640030.0%
MeOH63976381160.3%
MeCN63976366310.5%
Experimental δ59Co values, measured in relation to the internal reference complex ([Co(CN)6]3− in D2O), obtained from Chan et al. [52]. AD = absolute deviation; RD = relative deviation. HIm = imidazole; MeIm = methylimidazole; en = ethylenediamine.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Gomes, M.G.R.; De Souza, A.L.F.; Dos Santos, H.F.; De Almeida, W.B.; Paschoal, D.F.S. Assessment of a Computational Protocol for Predicting Co-59 NMR Chemical Shift. Magnetochemistry 2023, 9, 172. https://doi.org/10.3390/magnetochemistry9070172

AMA Style

Gomes MGR, De Souza ALF, Dos Santos HF, De Almeida WB, Paschoal DFS. Assessment of a Computational Protocol for Predicting Co-59 NMR Chemical Shift. Magnetochemistry. 2023; 9(7):172. https://doi.org/10.3390/magnetochemistry9070172

Chicago/Turabian Style

Gomes, Matheus G. R., Andréa L. F. De Souza, Hélio F. Dos Santos, Wagner B. De Almeida, and Diego F. S. Paschoal. 2023. "Assessment of a Computational Protocol for Predicting Co-59 NMR Chemical Shift" Magnetochemistry 9, no. 7: 172. https://doi.org/10.3390/magnetochemistry9070172

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop