Next Article in Journal
Cannabis sativa CBD Extract Shows Promising Antibacterial Activity against Salmonella typhimurium and S. newington
Previous Article in Journal
Influence of Homogenizing Methodology on Mechanical and Tribological Performance of Powder Metallurgy Processed Titanium Composites Reinforced by Graphene Nanoplatelets
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

DFT Calculations of 31P NMR Chemical Shifts in Palladium Complexes

by
Svetlana A. Kondrashova
,
Fedor M. Polyancev
and
Shamil K. Latypov
*
Arbuzov Institute of Organic and Physical Chemistry, FRC Kazan Scientific Center of RAS, 420088 Kazan, Russia
*
Author to whom correspondence should be addressed.
Molecules 2022, 27(9), 2668; https://doi.org/10.3390/molecules27092668
Submission received: 14 March 2022 / Revised: 4 April 2022 / Accepted: 19 April 2022 / Published: 21 April 2022

Abstract

:
In this study, comparative analysis of calculated (GIAO method, DFT level) and experimental 31P NMR shifts for a wide range of model palladium complexes showed that, on the whole, the theory reproduces the experimental data well. The exceptions are the complexes with the P=O phosphorus, for which there is a systematic underestimation of shielding, the value of which depends on the flexibility of the basis sets, especially at the geometry optimization stage. The use of triple-ζ quality basis sets and additional polarization functions at this stage reduces the underestimation of shielding for such phosphorus atoms. To summarize, in practice, for the rapid assessment of 31P NMR shifts, with the exception of the P=O type, a simple PBE0/{6-311G(2d,2p); Pd(SDD)}//PBE0/{6-31+G(d); Pd(SDD)} approximation is quite acceptable (RMSE = 8.9 ppm). Optimal, from the point of view of “price–quality” ratio, is the PBE0/{6-311G(2d,2p); Pd(SDD)}//PBE0/{6-311+G(2d); Pd(SDD)} (RMSE = 8.0 ppm) and the PBE0/{def2-TZVP; Pd(SDD)}//PBE0/{6-311+G(2d); Pd(SDD)} (RMSE = 6.9 ppm) approaches. In all cases, a linear scaling procedure is necessary to minimize systematic errors.

Graphical Abstract

1. Introduction

Transition metal (TM) complexes are catalysts for a number of reactions that are important from the point of view of practical applications [1,2,3,4]. The rational design of such systems requires information about the structure, type of coordination, and electronic structure as well. To this end, NMR chemical shifts (CS) of atoms directly coordinating the metal are sensitive to the electronic structure and, thus, can be used to obtain hidden information about the electronic state of the complex. Therefore, one needs to have a tool that allows the calculation of NMR shifts based on the principles of quantum chemistry. However, in the case of TM complexes, this is not a trivial task, and until recently, there was no such computational method that could be considered safe and reliable. This is due to the fact that in TM the requirements to account for electron correlation effects [5,6] sharply increase. Moreover, in the case of heavier metals, it may be necessary to take into account relativistic effects as well [7,8,9].
Nevertheless, it has recently been shown [10,11,12] that for nickel complexes, the 13C and 31P CSs can be calculated within the non-relativistic approximation at the Kohn–Sham (KS) density functional theory [13]. Moreover, the approach allowed several anomalies to be explained with the interpretation of NMR data [10] and to revise some structures [11].
From the point of view of catalysis and catalyst development, palladium- and platinum-based complexes play important roles [14,15,16,17,18,19,20,21,22,23]. Therefore, in order to deal with such systems, it is necessary to have a similar tool to calculate 31P CSs in the complexes of these metals. In this case, however, there is a certain pessimism, since it is believed that for 4d and, especially for 5d metals, relativistic effects may need to be taken into account [8]. Unfortunately, fully relativistic calculations for such systems are still difficult due to the enormous demands on computational resources. In this regard, it is interesting to see whether it will be sufficient to take these effects into account, at least within the framework of relatively simple scalar approximations using the quasi-relativistic effective core potential (ECP) [24]. Therefore, at this stage, we try to evaluate the scopes and limitations of density functional theory (DFT) methods for estimating 31P CSs in palladium complexes using this approach to describe the electronic structure of the metal. In fact, the present study was motivated by practical needs and the main goal is to propose a simple DFT method that could be safely applied to calculate 31P NMR shifts in Pd complexes.

2. Results and Discussion

2.1. General Overview

To calculate the 31P NMR shifts in Pd complexes, a combination similar to the one used in nickel complexes was applied as an initial approximation [11]—namely, the PBE0 functional with the Pople’s basis set of double-ζ quality (6-31+G(d)) on elements for geometry optimization and the basis set of triple-ζ quality (6-311G(2d,2p)) for shielding calculation—while ECP SDD was applied to describe Pd (PBE0/{6-311G(2d,2p); Pd(SDD)}//PBE0/{6-31+G(d); Pd(SDD)}; hereafter, this combination will be referred to as comb_1. To assess the quality of the calculation, Pd complexes based on organophosphorus ligands (39 in total, structures with numbering in Figures S1–S4 in the Supporting Information (SI)), which represent almost all known complexes of this type and have 31P CSs covering the wide range of possible values (from −50 to 200 ppm), were used as components of a test set. Calculations were carried out within the framework of the GIAO method [25], all 31P CSs were referenced to H3PO4, and the linear scaling procedure was applied [26,27,28] with coefficients similar to that for nickel complexes [11]. To begin with, the calculations were carried out without taking into account the solvent effects (Table S1).
It was found that, in general, for most of the complexes, the theory reproduces 31P CSs well, and there is a correlation between the calculated and experimental data close to linear (Figure 1).
A more detailed analysis allowed us to estimate the quality of calculations for each of the types of complexes under consideration. Specifically, in Pd complexes based on σ-donor ligands (Figure S1), the calculated 31P CSs correlate well with the experimental data (● in Figure 1, R2 = 0.982). In charged palladium complexes based on σ-donor ligands (Figure S2), the situation is somewhat more complicated. Here, the best agreement of the calculation with the experiment is achieved (1) for systems with a weakly coordinating counterion (BF4) if the counterion is not taken into account in the calculation; (2) for systems with a strongly coordinating counterion (Cl and Br) to take them into account if they complement the complex to a square–planar; otherwise, as with a weak counterion, it is necessary to carry out the calculation without counterion. With this approach, for such systems, there is good agreement between the calculation and experiment (R2 = 0.988, in Figure 1). The most important finding is that the systematic deviation of the calculated 31P CS from the experimental ones is observed in Pd complexes based on ligands containing P=O groups (Figure S3). Specifically, there is an overestimation by ~15–30 ppm for CSs of only P=O phosphorus atoms ( in Figure 1). At the same time, for this type of phosphorus, the dependence is also close to linear. As for palladium complexes based on π-donor ligands (Figure S4), the theory reproduces the experiment well (R2 = 0.973, in Figure 1). Thus, it can be argued that, at the PBE0/{6-311G(2d,2p); Pd(SDD)}//PBE0/{6-31+G(d); Pd(SDD)} level, a good agreement can be reached between the calculated and experimental 31P NMR shifts for most complexes. In principle, there are certain problems only with the P=O type of phosphorus.

2.2. Optimization of the Computational Approach

In an attempt to resolve this problem, the main factors that can affect the results of the calculation were considered. These are the medium effects, the impact of high-spin states, the influence of the quality of basis set/functional, the possible impact of relativistic effects, etc. Of the possible effects, one that is easily verifiable is the effect of a solvent. However, calculations taking into account the solvent within the polarizable continuum model (PCM) do not lead to improvements for the P=O phosphorus atoms (Table S1). Moreover, contributions due to high-spin states can be also excluded, since these molecular systems are very high in energy (35–60 kcal/mole, Table S3).
Thus, there is no “simple” way to correct the deviation for the P=O type of phosphorus. In order to create a cost-effective method, we fixed the functional to the PBE0 and screened different basis-set/functional combinations so as to obtain adequate agreement between calculated and experimental 31P shifts. This requires a detailed systematic analysis, which is more convenient to carry out on a smaller number of “training” models, including both problematic complexes (22, 23, 26, 2830, 32) and several normal systems (2, 9, 17, 38) with different types of ligands (σ, π, charged) in the entire range of 13P CSs (Figure 2).
It was previously shown that improving the quality of the basis set at the geometry optimization stage leads to significant improvement in the calculated 13C NMR shifts in nickel complexes [10]. Therefore, by analogy, the first step was an attempt to “enhance” the basis functions within the Pople’s style basis sets. Indeed, the addition of one more polarization function for heavy elements leads to an increase in the correlation coefficient between calculated and experimental CSs (Table 1, Figure 3). The use of a basis set of triple-ζ quality also improves the R2 (Table 1, Figure 3). The interplay of these modifications on the basis sets leads to the best results (Table 1, Figure 3), and the systematic underestimation of the P=O CSs decreases by half (by ~5–15 ppm), although it cannot be completely corrected even when using this resource-intensive combination (Table S4).
On the other hand, the enlarging of the basis sets on the elements at the CS calculation stage by adding diffuse functions leads only to a minimal improvement, although the calculations become very time-consuming (Figure 3). Thus, further augmentation of the basis sets at the shielding calculation stage does not seem rational.
As a last resort, in an attempt to improve the results of calculations, we further augmented the basis sets at the optimization stage by adding another polarization function on the elements. Indeed, this led to some improvement in the agreement of the calculated CS with the experimental ones for problematic phosphorus atoms (Table 1, entry 6), although the time costs increased significantly (vide infra).
Thus, we concluded that the more flexible the basis set is on heavy elements at the geometry optimization stage, the better the structure is, and as a result, the better the 31P NMR shifts. However, the reason for the imperfect agreement between the calculation and experiment may also be in the quality of the basis sets on the metal. For example, for nickel complexes, diffusion functions on the metal are important for obtaining the correct geometry of the complex and, accordingly, its 13C NMR shifts [10]. Unfortunately, in the case of palladium, the choice of basis sets is very limited. Of the currently available the def2-TZVPD ECP with diffuse functions on valence orbitals is the most extended basis set defined [21,29]. Thus, calculations were carried out using this ECP on palladium and 6-311G(2d,2p)//6-311+G(2d) basis sets on the elements at the shielding//optimization calculation stages, respectively. However, it was found that, in general, the use of this ECP does not lead to an improvement, compared with the comb_1 combination with SDD ECP on Pd (Table 1, entry 7).
A rare alternative to ECP on palladium is the Sapporo-DKH3-DZP all-electron (AE) basis set [30,31], which is a compact Gaussian-type basis set for relativistic molecular calculations. However, the use of this AE basis set on palladium leads even to worse results. For example, if to use the basis set for Pd at the geometry optimization stage (Table 1, entry 8) the systematic underestimation of the P=O CSs increases (20 versus 18 ppm). Moreover, an application of this basis set for Pd at both stages of calculations leads to the worst results. Specifically, in addition to P=O (∆δ = 28 ppm), the P=P (∆δ = 65 ppm) phosphorus also starts to deviate from the main correlation (Table 1, entry 9). Thus, it appears that this basis set for Pd is not good for NMR shielding calculations.
Up to this point, we had checked only Pople’s basis sets on elements. As an alternative, three more popular basis sets, which are often used to calculate magnetic properties in organic systems, were tested. There are Dunning’s [32,33], Jensen’s [34,35,36], and Alrichs’s [29,37,38] basis sets.
First, we analyzed the performance of the basis set for the CS calculation stage using the best geometry optimized at the PBE0/6-311+G(3df) level. It was revealed that the use of Dunning’s correlation-consistent basis sets of triple-ζ quality (cc-pVTZ) for shielding calculation leads to some improvement, particularly for problematic P=O phosphorus (Table 1, entry 10). The Jensen’s polarization-consistent (pc-2) basis sets, which were specially developed for DFT calculations, showed worse results for P=O phosphorus (∆δ = 19 ppm). Additionally, the best results are observed if the Alrichs’s (def2-TZVP) basis sets were used for elements (∆δ = 9 ppm).
Next, we considered how these basis sets perform at the geometry optimization stage while the PBE0/6-311G(2d,2p) level was applied as a standard for CS calculation (Table 1, entry 16–19). We started with the cc-pVDZ (double-ζ) for optimization. The use of this basis set for geometry optimization leads to a very poor correlation between calculation and experiment (R2 = 0.911). Changing from the basis set of double- to triple-ζ quality (cc-pVTZ) leads to notable improvement, and the R2 becomes close to that obtained with PBE0/6-311+G(2d) geometry. The use of pc-2 or def2-TZVP basis sets results in 31P shifts for problematic phosphorus that are only slightly better than the corresponding Pople’s combination (comb_4), although the time costs increase substantially (vide infra).
Thus, the def2-TZVP//6-311+G(3df) combination looks the best. However, it is very difficult to work with in terms of both geometry optimization and CS calculation. Therefore, taking into account practical aspects, options with lighter basis sets were considered both at the geometry optimization stage and shielding calculation. It was found that, for example, the PBE0/6-311+G(2d) geometry with the def2-TZVP basis at the shielding calculation reproduces the experimental values quite well (∆δ = 13 ppm). At the same time, easing the basis set at the shielding calculation stage just to TZV leads to a significant deterioration in the results (∆δ = 81 ppm).
Several other combinations were also tested (Table 1, entries 20–22). Of those considered, only the use of the def2-TZVP basis at both stages of the calculation leads to a good correlation between the calculation and experiment (R2 = 0.964). The rest of the combinations do not look promising. Moreover, some variants proved to be bad, although, for example, in the case of 13C shifts in nickel complexes, they performed exceptionally well [10]. It should be noted that, in this case, both the P=O and phosphorus for systems with multiple bonds begin to deviate strongly from the general dependence (Table 1, entry 20).
Thus, based on the analysis performed on the training set, we came to the conclusion that several combinations are promising, leading to approximately the same correlation coefficients between the calculation and experiments. However, on the other hand, these approximations vary greatly in the time spent on calculations. Therefore, taking into account these factors, some of them seem to be more preferable from a practical point of view.

2.3. Practical Aspects, Quality of Calculations, and Computational Costs

The aim of this study is not just to find an approach that will allow 31P shifts to be estimated in Pd complexes with sufficient accuracy but to develop a tool that can be relatively easily applied for large molecules of practical interest using modest computer resources. Unfortunately, not all the approximations discussed above satisfy these requirements. Therefore, in this section, we discuss a comparative analysis of computational costs in order to choose the leading combinations in terms of the “price–quality” ratio.
First of all, let us consider the optimization stage since it is the most time-consuming. Complex 22 was taken as an example, and all optimizations were carried out from the same starting geometry. If we take the PBE0/6-31+G(d) approximation as a reference for time costs, then among the Pople’s type basis sets, geometry optimization using the 6-311+G(2d) and 6-311+G(3df) basis sets takes ~3 and ~10 times longer, respectively (Figure 4a). For the geometry optimization stage, this is essential, since, when performing calculations on a multiprocessor system, for example, it may take several days.
In turn, using Dunning’s basis sets of double- or triple-ζ quality is ~2 and ~7 times longer than using a 6-31+G(d) basis. Finally, the Jensen’s and Alrichs’s basis sets are ~7–9 times more time-consuming than the 6-31+G(d) basis set. Thus, at the geometry optimization stage, these basis sets are several times more time-consuming than the Pople’s 6-311+G(2d) basis sets at close values of R2 (Figure 3).
As for the CS calculation stage, when adding diffuse functions to the starting basis 6-311G(2d,2p), the “time–cost” increases by an order of magnitude only with a slight increase in quality (Figure 4b). Moreover, calculations using Dunning’s and Jensen’s basis sets of triple-ζ quality are lengthened by ~14 times, compared with 6-311G(2d,2p) at lower quality. Using the Alrichs’s basis set leads to an increase in time by about an order of magnitude, but it leads to a gain in quality, compared with the Pople’s basis set.
Thus, taking into account practical aspects, preference can be given to two combinations, in order to check their quality on a more extended test set—namely, the 6-311+G(2d) level looks optimal for the geometry optimization stage, while for the shielding calculation, the use of 6-311G(2d,2p) and def2-TZVP basis sets looks promising.

2.4. Checking the Leading Combinations on the Entire Test Set of Pd Complexes

The calculations performed for the entire test set using the PBE0/{6-311G(2d,2p); Pd(SDD)}//PBE0/{6-311+G(2d); Pd(SDD)} combination showed that improvement is observed for all problematic phosphorus (R2 = 0.963 versus 0.954), while for “normal” complexes, the changes are minimal (Figure 5). In other words, if there is no P=O phosphorus in the complex, then the lighter comb_1 combination is also sufficient.
At the same time, a more careful analysis reveals limitations typical for this level of calculations in the form of systematic errors when moving to lower fields. However, it is easy to minimize such errors by performing a linear scaling procedure [26,27,28] according to Equation (1) using the regression analysis parameters for the test set (Table 2).
δscaled = (δ − Intercept)/Slope,
where δ is the calculated CS for a particular nucleus, and intercept and slope are scaling factors.
Thus, the application of this procedure allows for the evaluation of 31P shifts in palladium complexes with RMSE = 8.0 ppm.
Similar calculations using (PBE0/{def2-TZVP; Pd(SDD)}//PBE0/{6-311+G(2d); Pd(SDD)}) approximation for the entire test set lead to even higher correlation coefficients (R2 = 0.972). The linear scaling procedure allows for estimates of 31P CS with the best RMSE = 6.9 ppm (Figure S5).

2.5. Impact of Coordination Type on 31P NMR Shifts

The NMR shifts of the atoms directly involved in the formation of the complex with the TM strongly depend on the particular structure. Therefore, a comparison of calculated and experimental 31P NMR shifts can be used as a reliable tool to confirm or rule out hypothetical structures. The effectiveness of the tool can be demonstrated by the example of structure determination of complexes based on triorganylcyclotriphosphines [39]. In general, in such compounds, cyclophosphines can bind with TM in the η2 or η1,η1 fashions. According to X-ray data, in the case of Pd, the P=P double bond coordinates with transition metals in the η2 form, while for Pt, they undergo a ring-opening process during the reaction.
Molecules 27 02668 i001
Indeed, according to calculations, for complex 35, for example, both types correspond to an energy minimum (Figure 6). However, only for the η2 form (35, Figure 6) is the difference in the calculated (comb_3) versus the experimental 31P NMR shifts small (ca. 9 ppm), while for the η1,η1 form, with the ring-opening process of complex 35 (a), the deviation is more than 130 ppm (Figure 6). Thus, comparing theoretical results with those obtained by experiment is an easy way to determine its coordination type in solution as a π complex.

3. Materials and Methods

Calculations

The quantum chemical calculations were carried out within the framework of the Kohn Sham density functional theory [13], with the Gaussian 03 [40] (Revision B.04) and the Gaussian 16 [41] (Revision A.03) software packages by using PBE0 [42] functional and a number of families of basis sets (Pople’s [43,44,45,46,47,48,49,50], Ahlrichs’s [29,37,38], Dunning’s [32,33], Jensen’s [34,35,36] sets). For the Pd center, the quasi-relativistic Stuttgart–Dresden ECP28MWB was used with corresponding (8s7p6d)/(6s5p3d) GTO valence basis set [24] (denoted as “(SDD)”), Sapporo-DKH3-DZP-diffuse (deprecated) basis set [30,31], and def2-TZVPD basis set [21,29]. Wherever possible, geometry optimization was started from an X-ray structure. For most of the complexes, the calculations were carried out for all possible conformers/isomers, and results for the lowest energy forms were used in the analysis. To take into account the medium effects, calculations were also carried out in the framework of the Polarizable Continuum Model [51] (denoted as “PCM”), with chloroform as solvent. To determine the energies of triplet states, geometry optimizations were performed at the spin-unrestricted formalism (UPBE0/6-31+G(d)). 31P NMR CSs were calculated by the GIAO method [25]. All 31P data were referenced to H3PO4.
The pc-2, def2-TZVP, def2-TZVPD, and Sapporo-DKH3-DZP-diffuse basis sets were downloaded from the EMSL basis set library for the Gaussian package [52].
The Gaussian 03 calculations were carried out on a PC with IntelCore i7-3970X CPU, 3.5 GHz. Calculations with 6-311+G(3df), pc-2, cc-pVTZ, def2-TZVP, and Sapporo-DKH3-DZP-diffuse (deprecated) basis sets were carried out with the Gaussian 16 on 20 CPUs, IntelXeon ES-2650 2.20 GHz.

4. Conclusions

Comparative analysis of calculated (GIAO method, DFT level) and experimental 31P NMR shifts for a wide range of model palladium complexes showed that, on the whole, the theory reproduces the experimental data well. The exceptions are the complexes with P=O phosphorus, for which there is a systematic underestimation of shielding, the value of which depends on the flexibility of the basis sets, especially at the geometry optimization stage. The use of basis sets of triple-ζ quality and additional polarization functions at this stage reduces the underestimation of shielding for such phosphorus compounds.
To summarize, in practice, for the rapid assessment of phosphorus CS, with the exception of the P=O type, a rather simple PBE0/{6-311G(2d,2p); Pd(SDD)}//PBE0/{6-31+G(d); Pd(SDD)} approximation is quite acceptable (RMSE = 8.9 ppm). Optimal from the point of view of “price-quality” ratio is the PBE0/{6-311G(2d,2p); Pd(SDD)}//PBE0/{6-311+G(2d); Pd(SDD)} (RMSE = 8.0 ppm) and the PBE0/{def2-TZVP; Pd(SDD)}//PBE0/{6-311+G(2d); Pd(SDD)} (RMSE = 6.9 ppm) approaches. The use of the PBE0/6-311+G(3df) level at the geometry optimization stage leads to some further improvement in the calculation, but the requirements for computer resources are increasing dramatically. In all cases, a linear scaling procedure is necessary to minimize systematic errors.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules27092668/s1, Structures of all model complexes [53,54,55,56,57,58,59,60,61,62,63,64,65,66,67,68,69,70] with experimental 31P NMR shifts and all calculation results. Table S1: Experimental and calculated 31P NMR shifts (ppm) for all model complexes. Table S2: Empirical scaling factors obtained by the linear regression analysis of calculated and experimental δ31P for the title Pd-complexes. Table S3: Triplet-singlet states energy gap (UPBE0/6-31+G(d)) for the “training” set of Pd complexes (2, 9, 17, 2223, 26, 2830, 32, 38). Table S4: Experimental and calculated (by different combinations) 31P NMR shifts (ppm) for the “training” set of Pd complexes (2, 9, 17, 2223, 26, 2830, 32, 38).

Author Contributions

Conceptualization, S.K.L.; methodology, S.K.L. and F.M.P.; validation, S.K.L.; formal analysis, S.A.K., F.M.P. and S.K.L.; investigation, S.A.K., F.M.P. and S.K.L.; data curation, S.A.K., F.M.P. and S.K.L.; writing—original draft preparation, S.K.L.; writing—review and editing, S.K.L. and S.A.K.; visualization, S.A.K.; supervision, S.K.L.; project administration, S.K.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All data are contained within the article and Supporting Information.

Acknowledgments

The authors gratefully acknowledge the Assigned Spectral–Analytical Center of the FRC Kazan Scientific Center of RAS.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples of the compounds are not available from the authors.

References

  1. Korch, K.M.; Watson, D.A. Cross-coupling of heteroatomic electrophiles. Chem. Rev. 2019, 119, 8192–8228. [Google Scholar] [CrossRef] [PubMed]
  2. Negishi, E.-I. Magical power of transition metals: Past, present, and future. Angew. Chem. Int. Ed. 2011, 50, 6738–6764. [Google Scholar] [CrossRef] [PubMed]
  3. Kinzel, N.W.; Werlé, C.; Leitner, W. Transition metal complexes as catalysts for the electroconversion of CO2: An organometallic perspective. Angew. Chem. Int. Ed. 2021, 60, 11628–11686. [Google Scholar] [CrossRef] [PubMed]
  4. Takaya, J. Catalysis using transition metal complexes featuring main group metal and metalloid compounds as supporting ligands. Chem. Sci. 2021, 12, 1964–1981. [Google Scholar] [CrossRef]
  5. Bühl, M.; Kaupp, M.; Malkina, O.L.; Malkin, V.G. The DFT route to NMR chemical shifts. J. Comput. Chem. 1999, 20, 91–105. [Google Scholar] [CrossRef]
  6. Autschbach, J.; Ziegler, T. Double perturbation theory: A powerful tool in computational coordination chemistry. Coord. Chem. Rev. 2003, 238–239, 83–126. [Google Scholar] [CrossRef]
  7. Vıcha, 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 approaches1. Phys. Chem. Chem. Phys. 2015, 17, 24944–24955. [Google Scholar] [CrossRef] [Green Version]
  8. Bagno, A.; Saielli, G. Relativistic DFT calculations of the NMR properties and reactivity of transition metal methane σ-complexes: Insights on C–H bond activation. Phys. Chem. Chem. Phys. 2011, 13, 4285–4291. [Google Scholar] [CrossRef]
  9. Vícha, J.; Novotny, J.; Komorovsky, S.; Straka, M.; Kaupp, M.; Marek, R. Relativistic heavy-neighbor-atom effects on NMR shifts: Concepts and trends across the periodic table. Chem. Rev. 2020, 120, 7065–7103. [Google Scholar] [CrossRef]
  10. Kondrashova, S.A.; Polyancev, F.M.; Ganushevich, Y.S.; Latypov, S.K. DFT approach for predicting 13C NMR shifts of atoms directly coordinated to nickel. Organometallics 2021, 40, 1614–1625. [Google Scholar] [CrossRef]
  11. Latypov, S.K.; Kondrashova, S.A.; Polyancev, F.M.; Sinyashin, O.G. Quantum chemical calculations of 31P NMR chemical shifts in nickel complexes: Scope and limitations. Organometallics 2020, 39, 1413–1422. [Google Scholar] [CrossRef]
  12. Payard, P.A.; Perego, L.A.; Grimaud, L.; Ciofini, I. A DFT protocol for the prediction of 31P NMR chemical shifts of phosphine ligands in first-row transition-metal complexes. Organometallics 2020, 39, 3121–3130. [Google Scholar] [CrossRef]
  13. Kohn, W.; Sham, L.J. Self-consistent equations including exchange and correlation effects. Phys. Rev. 1965, 140, A1133. [Google Scholar] [CrossRef] [Green Version]
  14. Jadhav, B.D.; Pardeshi, S.K. Synthesis and catalytic application of Pd complex catalysts: Atom-efficient cross-coupling of triarylbismuthines with haloarenes and acid chlorides under mild conditions. Appl. Organomet. Chem. 2017, 31, e3591. [Google Scholar] [CrossRef]
  15. Zhu, J.; Lindsay, V.N. Benzimidazolyl palladium complexes as highly active and general bifunctional catalysts in sustainable cross-coupling reactions. ACS Catal. 2019, 9, 6993–6998. [Google Scholar] [CrossRef]
  16. Schroeter, F.; Soellner, J.; Strassner, T. Cross-coupling catalysis by an anionic palladium complex. ACS Catal. 2017, 7, 3004–3009. [Google Scholar] [CrossRef]
  17. Cao, Q.; Bailie, D.S.; Fu, R.; Muldoon, M.J. Cationic palladium (II) complexes as catalysts for the oxidation of terminal olefins to methyl ketones using hydrogen peroxide. Green Chem. 2015, 17, 2750–2757. [Google Scholar] [CrossRef] [Green Version]
  18. Scattolin, T.; Voloshkin, V.A.; Martynova, E.; Broeck, S.M.V.; Beliš, M.; Cazin, C.S.; Nolan, S.P. Synthesis and catalytic activity of palladium complexes bearing N-heterocyclic carbenes (NHCs) and 1,4,7-triaza-9-phosphatricyclo[5.3.2.1]tridecane (CAP) ligands. Dalton Trans. 2021, 50, 9491–9499. [Google Scholar] [CrossRef]
  19. Zhou, T.; Szostak, M. Palladium-catalyzed cross-couplings by C–O bond activation. Catal. Sci. Technol. 2020, 10, 5702–5739. [Google Scholar] [CrossRef]
  20. Miyaura, N.; Suzuki, A. Palladium-catalyzed cross-coupling reactions of organoboron compounds. Chem. Rev. 1995, 95, 2457–2483. [Google Scholar] [CrossRef] [Green Version]
  21. Labinger, J.A. Platinum-catalyzed C–H functionalization. Chem. Rev. 2017, 117, 8483–8496. [Google Scholar] [CrossRef] [PubMed]
  22. Fürstner, A.; Davies, P.W. Catalytic carbophilic activation: Catalysis by platinum and gold π acids. Angew. Chem. Int. Ed. 2007, 46, 3410–3449. [Google Scholar] [CrossRef] [PubMed]
  23. Hughes, A.E.; Haque, N.; Northey, S.A.; Giddey, S. Platinum group metals: A review of resources, production and usage with a focus on catalysts. Resources 2021, 10, 93. [Google Scholar] [CrossRef]
  24. Andrae, D.; Haeussermann, U.; Dolg, M.; Stoll, H.; Preuss, H. Energy-adjusted ab initio pseudopotentials for the second and third row transition elements. Theor. Chem. Acc. 1990, 77, 123–141. [Google Scholar] [CrossRef]
  25. Hansen, A.E.; Bouman, T.D. Localized orbital/local origin method for calculation and analysis of NMR shieldings. Applications to 13C shielding tensors. J. Chem. Phys. 1985, 82, 5035–5047. [Google Scholar] [CrossRef]
  26. Pierens, G.K. 1H and 13C NMR scaling factors for the calculation of chemical shifts in commonly used solvents using density functional theory. J. Comput. Chem. 2014, 35, 1388–1394. [Google Scholar] [CrossRef]
  27. Lodewyk, M.W.; Siebert, M.R.; Tantillo, D.J. Computational prediction of 1H and 13C chemical shifts: A useful tool for natural product, mechanistic, and synthetic organic chemistry. Chem. Rev. 2012, 112, 1839–1862. [Google Scholar] [CrossRef] [PubMed]
  28. Konstantinov, I.A.; Broadbelt, L.J. Regression formulas for density functional theory calculated 1H and 13C NMR chemical shifts in toluene-d8. J. Phys. Chem. A 2011, 115, 12364–12372. [Google Scholar] [CrossRef]
  29. Weigend, F.; Ahlrichs, R. Balanced basis sets of split valence, triple zeta valence and quadruple zeta valence quality for H to Rn: Design and assessment of accuracy. Phys. Chem. Chem. Phys. 2005, 7, 3297–3305. [Google Scholar] [CrossRef]
  30. Noro, T.; Sekiya, M.; Koga, T.; Saito, S.L. Relativistic contracted Gaussian-type basis functions for atoms K through Xe. Chem. Phys. Lett. 2009, 481, 229–233. [Google Scholar] [CrossRef]
  31. Osanai, Y.O.U.; Sekiya, M.; Noro, T.; Koga, T. Valence and correlating basis sets for the second transition-metal atoms from Y to Cd. Mol. Phys. 2003, 101, 65–71. [Google Scholar] [CrossRef]
  32. Dunning, T.H., Jr. Gaussian basis sets for use in correlated molecular calculations. I. The atoms boron through neon and hydrogen. J. Chem. Phys. 1989, 90, 1007–1023. [Google Scholar] [CrossRef]
  33. Woon, D.E.; Dunning, T.H. Gaussian basis sets for use in correlated molecular calculations. III. The atoms aluminum through argon. J. Chem. Phys. 1993, 98, 1358–1371. [Google Scholar] [CrossRef] [Green Version]
  34. Jensen, F. Polarization consistent basis sets: Principles. J. Chem. Phys. 2001, 115, 9113–9125. [Google Scholar] [CrossRef]
  35. Jensen, F. Polarization consistent basis sets. II. Estimating the Kohn-Sham basis set limit. J. Chem. Phys. 2002, 116, 7372–7379. [Google Scholar] [CrossRef] [Green Version]
  36. Jensen, F.; Helgaker, T. Polarization consistent basis sets. V. The elements Si-Cl. J. Chem. Phys. 2004, 121, 3463–3470. [Google Scholar] [CrossRef] [Green Version]
  37. Schaefer, A.; Horn, H.; Ahlrichs, R. Fully optimized contracted Gaussian-basis sets for atoms Li to Kr. J. Chem. Phys. 1992, 97, 2571–2577. [Google Scholar] [CrossRef]
  38. Schaefer, 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. [Google Scholar] [CrossRef]
  39. Fenske, D.; Schottmueller, H. New Research of Reaction Behaviour of Triorganylcyclotriphosphines. The Crystal Structures of [(PPh3)2Pt(PtBu)3], [(PPh3)2Pd(PtBu)2, [(CO)4Cr{(PiPr)3}2], [RhCl(PPh3)(PtBu)3], [(NiCO)6(μ2-CO)3{(PtBu)2}2], and [(CpFeCO)2(μ-CO)(μ-PHtBu)]+∙[FeCl3(thf)]. Z. Anorg. Allg. Chem. 1998, 624, 443–451. [Google Scholar] [CrossRef]
  40. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Montgomery, J.A., Jr.; Vreven, T.; Kudin, K.N.; Burant, J.C.; et al. Gaussian 03, Revision B.04; Gaussian, Inc.: Pittsburgh, PA, USA, 2003. [Google Scholar]
  41. 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 A.03; Gaussian, Inc.: Wallingford, CT, USA, 2016. [Google Scholar]
  42. 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]
  43. Hehre, W.J.; Ditchfield, R.; Pople, J.A. Self-Consistent Molecular Orbital Methods. XII. Further Extensions of Gaussian-Type Basis Sets for Use in Molecular Orbital Studies of Organic Molecules. J. Chem. Phys. 1972, 56, 2257–2261. [Google Scholar] [CrossRef]
  44. Clark, T.; Chandrasekhar, J.; Spitznagel, G.W.; Schleyer, P.V.R. Efficient diffuse function-augmented basis sets for anion calculations. III. The 3-21+G basis set for first-row elements, Li–F. J. Comput. Chem. 1983, 4, 294–301. [Google Scholar] [CrossRef]
  45. Francl, M.M.; Pietro, W.J.; Hehre, W.J.; Binkley, J.S.; Gordon, M.S.; DeFrees, D.J.; Pople, J.A. Self-consistent molecular orbital methods. XXIII. A polarization-type basis set for second-row elements. J. Chem. Phys. 1982, 77, 3654–3665. [Google Scholar] [CrossRef] [Green Version]
  46. Frisch, M.J.; Pople, J.A.; Binkley, J.S. Self-consistent molecular orbital methods 25. Supplementary functions for Gaussian basis sets. J. Chem. Phys. 1984, 80, 3265–3269. [Google Scholar] [CrossRef]
  47. Krishnan, R.; Binkley, J.S.; Seeger, R.; Pople, J.A. Self-consistent molecular orbital methods. XX. A basis set for correlated wave functions. J. Chem. Phys. 1980, 72, 650–654. [Google Scholar] [CrossRef]
  48. McLean, A.D.; Chandler, G.S. Contracted Gaussian basis sets for molecular calculations. I. Second row atoms, Z = 11–18. J. Chem. Phys. 1980, 72, 5639–5648. [Google Scholar] [CrossRef]
  49. Spitznagel, G.W.; Clark, T.; von Ragué Schleyer, P.; Hehre, W.J. An evaluation of the performance of diffuse function-augmented basis sets for second row elements, Na-Cl. J. Comput. Chem. 1987, 8, 1109–1116. [Google Scholar] [CrossRef]
  50. Ditchfield, R.; Hehre, W.J.; Pople, J.A. Self-Consistent Molecular-Orbital Methods. IX. An Extended Gausian-Type Basis for Molecular-Orbital Studies of Organic Molecules. J. Chem. Phys. 1971, 54, 724–728. [Google Scholar] [CrossRef]
  51. Miertus, S.; Scrocco, E.; Tomasi, J. Electrostatic interaction of a solute with a continuum. A direct utilization of AB initio molecular potentials for the prevision of solvent effects. Chem. Phys. 1981, 55, 117–129. [Google Scholar] [CrossRef]
  52. Pritchard, B.P.; Altarawy, D.; Didier, B.; Gibson, T.D.; Windus, T.L. A 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]
  53. Portnoy, M.; Milstein, D. Chelate effect on the structure and reactivity of electron-rich palladium complexes and its relevance to catalysis. Organometallics 1993, 12, 1655–1664. [Google Scholar] [CrossRef]
  54. Raebiger, J.W.; Miedaner, A.; Curtis, C.J.; Miller, S.M.; Anderson, O.P.; DuBois, D.L. Using ligand bite angles to control the hydricity of palladium diphosphine complexes. J. Am. Chem. Soc. 2004, 126, 5502–5514. [Google Scholar] [CrossRef]
  55. Nixon, J.F.; Sillett, G.J. 31P and 195Pt NMR studies on fluxional η1-ligated 1,2,4-triphosphacyclopentadienyl palladium(II) and platinum(II) complexes. J. Organomet. Chem. 1993, 461, 237–245. [Google Scholar] [CrossRef]
  56. Kane, J.C.; Wong, E.H.; Yap, G.P.; Rheingold, A.L. Synthesis and structural studies of molybdenum and palladium complexes of 1,5-diaza-3,7-diphosphacyclooctane ligands. Polyhedron 1999, 18, 1183–1188. [Google Scholar] [CrossRef]
  57. Gouygou, M.; Tissot, O.; Daran, J.C.; Balavoine, G.G. Biphosphole: A C2 Symmetry Chiral Bidentate Ligand. Synthesis and Characterization Of Nickel, Palladium, Platinum, and Rhodium Complexes. Organometallics 1997, 16, 1008–1015. [Google Scholar] [CrossRef]
  58. Wächtler, E.; Priver, S.H.; Wagler, J.; Heine, T.; Zhechkov, L.; Bennett, M.A.; Bhargava, S.K. Metallophilic contacts in 2-C6F4PPh2 bridged heterobinuclear complexes: A crystallographic and computational study. Inorg. Chem. 2015, 54, 6947–6957. [Google Scholar] [CrossRef]
  59. David, M.A.; Wicht, D.K.; Glueck, D.S.; Yap, G.P.; Liable-Sands, L.M.; Rheingold, A.L. Palladium-catalyzed phosphaketene decarbonylation: Diphosphaureylene intermediates in diphosphene formation. Organometallics 1997, 16, 4768–4770. [Google Scholar] [CrossRef]
  60. Sevillano, P.; Habtemariam, A.; Parsons, S.; Castiñeiras, A.; García, M.E.; Sadler, P.J. Gold (I)-induced chelate ring-opening of palladium (II) and platinum (II) triphos complexes. J. Chem. Soc. Dalton Trans. 1999, 16, 2861–2870. [Google Scholar] [CrossRef]
  61. Zhuravel, M.A.; Grewal, N.S.; Glueck, D.S.; Lam, K.C.; Rheingold, A.L. Cyclometalation of Dimesitylphosphine in Cationic Palladium (II) and Platinum (II) Complexes: P–H vs. C–H Activation. Organometallics 2000, 19, 2882–2890. [Google Scholar] [CrossRef]
  62. Mizuta, T.; Okano, A.; Sasaki, T.; Nakazawa, H.; Miyoshi, K. Palladium (II) and platinum (II) complexes of a tetraphosphamacrocycle. X-ray crystal structures of phosphorus analogs of a (tetramethylcyclam) metal complex. Inorg. Chem. 1997, 36, 200–203. [Google Scholar] [CrossRef]
  63. Mason, L.J.; Perrault, E.M.; Miller, S.M.; Helm, M.L. Group 10 metal complexes of a cyclic diphosphine: The crystal structures of bis (cis-P,P′-diphenyl-1,4-diphospha-cyclohexane)M(II)chloride, M= palladium, platinum. Inorg. Chem. Com. 2006, 9, 946–948. [Google Scholar] [CrossRef]
  64. Melaimi, M.; Thoumazet, C.; Ricard, L.; Le Floch, P. Syntheses of a 2,6-bis-(methylphospholyl) pyridine ligand and its cationic Pd(II) and Ni(II) complexes—Application in the palladium-catalyzed synthesis of arylboronic esters. J. Organomet. Chem. 2004, 689, 2988–2994. [Google Scholar] [CrossRef]
  65. Kohler, M.C.; Grimes, T.V.; Wang, X.; Cundari, T.R.; Stockland, R.A., Jr. Arylpalladium phosphonate complexes as reactive intermediates in phosphorus—Carbon bond forming reactions. Organometallics 2009, 28, 1193–1201. [Google Scholar] [CrossRef] [Green Version]
  66. Stockland, R.A.; Levine, A.M.; Giovine, M.T.; Guzei, I.A.; Cannistra, J.C. Reductive elimination from metal phosphonate complexes: Circumvention of competing protonolysis reactions. Organometallics 2004, 23, 647–656. [Google Scholar] [CrossRef]
  67. Stockland, R.A., Jr.; Maher, D.L.; Anderson, G.K.; Rath, N.P. Demethylation of trimethylphosphite promoted by dichlorodiphosphineplatinum and palladium complexes. Structures of the metallophosphonate complexes [Pt(P∧P){P(O)(OMe)2}2](P∧P= dppe, dppp). Polyhedron 1999, 18, 1067–1075. [Google Scholar] [CrossRef]
  68. Gómez-Ruiz, S.; Hey-Hawkins, E. Oxidative cleavage of tetraaryltetraphosphane-1, 4-diides by nickel (ii) and palladium (ii): Formation of unusual Ni0 and Pd0 diaryldiphosphene complexes. Dalton Trans. 2007, 48, 5678–5683. [Google Scholar] [CrossRef]
  69. Phillips, I.G.; Ball, R.G.; Cavell, R.G. Reactions of perfluoromethyl-substituted cyclopolyphosphines with zerovalent group 10 metal complexes. Crystal and molecular structure of a complex with a coordinated diphosphene, [Pd(η2-CF3P=PCF3)(PPh3)2]. Inorg. Chem. 1992, 31, 1633–1641. [Google Scholar] [CrossRef]
  70. Chatt, J.; Hitchcock, P.B.; Pidcock, A.; Warrens, C.P.; Dixon, K.R. The nature of the co-ordinate link. Part 11. Synthesis and phosphorus-31 nuclear magnetic resonance spectroscopy of platinum and palladium complexes containing side-bonded (E)-diphenyldiphosphene. X-Ray crystal and molecular structures of [Pd{(E)-PhP=PPh}(Ph2PCH2CH2PPh2)] and [Pd{[(E)-PhP=PPh][W(CO)5]2}(Ph2PCH2CH2PPh2)]. J. Chem. Soc. Dalton Trans. 1984, 10, 2237–2244. [Google Scholar]
Figure 1. Correlation of calculated (PBE0/{6-311G(2d,2p); Pd(SDD)}//PBE0/{6-31+G(d); Pd(SDD)}) vs. experimental 31P CSs for palladium complexes 139. Complexes based on σ-donor ligands (●), charged palladium complexes based on σ-donor ligands (), complexes containing P=O groups ( are the P=O phosphorus, are other phosphorus atoms in these types of complexes), and complexes based on π-donor ligands ().
Figure 1. Correlation of calculated (PBE0/{6-311G(2d,2p); Pd(SDD)}//PBE0/{6-31+G(d); Pd(SDD)}) vs. experimental 31P CSs for palladium complexes 139. Complexes based on σ-donor ligands (●), charged palladium complexes based on σ-donor ligands (), complexes containing P=O groups ( are the P=O phosphorus, are other phosphorus atoms in these types of complexes), and complexes based on π-donor ligands ().
Molecules 27 02668 g001
Figure 2. “Training” set of palladium complexes.
Figure 2. “Training” set of palladium complexes.
Molecules 27 02668 g002
Figure 3. R2 dependence for the “training” set of Pd-complexes on combinations used in calculations. In all cases, the functional was fixed to the PBE0.
Figure 3. R2 dependence for the “training” set of Pd-complexes on combinations used in calculations. In all cases, the functional was fixed to the PBE0.
Molecules 27 02668 g003
Figure 4. Relative time spent for 22 using different basis sets at the geometry optimization stage (a) and calculation of 31P shift (b).
Figure 4. Relative time spent for 22 using different basis sets at the geometry optimization stage (a) and calculation of 31P shift (b).
Molecules 27 02668 g004
Figure 5. Correlation of calculated (PBE0/{6-311G(2d,2p); Pd(SDD)}//PBE0/{6-311+G(2d); Pd(SDD)}) vs. experimental 31P CSs for palladium complexes 139. Notation of the complexes is the same as that in Figure 1.
Figure 5. Correlation of calculated (PBE0/{6-311G(2d,2p); Pd(SDD)}//PBE0/{6-311+G(2d); Pd(SDD)}) vs. experimental 31P CSs for palladium complexes 139. Notation of the complexes is the same as that in Figure 1.
Molecules 27 02668 g005
Figure 6. Structures of complexes 35 and 35a with experimental (green) and calculated (black) 31P CSs (in ppm) included.
Figure 6. Structures of complexes 35 and 35a with experimental (green) and calculated (black) 31P CSs (in ppm) included.
Molecules 27 02668 g006
Table 1. R2 dependence on combination used in the calculations for the “training” set of Pd complexes (2, 9, 17, 2223, 26, 2830, 32, 38).
Table 1. R2 dependence on combination used in the calculations for the “training” set of Pd complexes (2, 9, 17, 2223, 26, 2830, 32, 38).
EntryCombinationBasis Set aR2Δδ b, ppm
ElementsPd
1comb_16-311G(2d,2p)//6-31+G(d)SDD//SDD0.93624
2comb_26-311G(2d,2p)//6-31+G(2d)SDD//SDD0.95119
3comb_36-311G(2d,2p)//6-311+G(d)SDD//SDD0.94422
4comb_46-311G(2d,2p)//6-311+G(2d)SDD//SDD0.95518
5comb_56-311+G(2d,2p)//6-311+G(2d)SDD//SDD0.95918
6comb_66-311G(2d,2p)//6-311+G(3df)SDD//SDD0.95914
7comb_76-311G(2d,2p)//6-311+G(2d)def2-TZVPD//def2-TZVPD0.95319
8comb_86-311G(2d,2p)//6-311+G(2d)SDD//DKH3-DZP0.90420
9comb_96-311G(2d,2p)//6-311+G(2d)DKH3-DZP//DKH3-DZP0.82928 (65) c
10comb_10cc-pVTZ//6-311+G(3df)SDD//SDD0.95015
11comb_11pc-2//6-311+G(3df)SDD//SDD0.94319
12comb_12def2-TZVP//6-311+G(3df)SDD//SDD0.9659
13comb_13def2-TZVP//6-311+G(2d)SDD//SDD0.96413
14comb_14TZV//6-311+G(3df)SDD//SDD0.83368
15comb_15def2-TZVP//def2-TZVPSDD//SDD0.96411
16comb_166-311G(2d,2p)//cc-pVDZSDD//SDD0.91136
17comb_176-311G(2d,2p)//cc-pVTZSDD//SDD0.95317
18comb_186-311G(2d,2p)//pc-2SDD//SDD0.95811
19comb_196-311G(2d,2p)//def2-TZVPSDD//SDD0.95615
20comb_20TZV//TZVSDD//SDD0.689118 (101) c
21comb_21cc-pVTZ//6-31+G(d)SDD//SDD0.92722
22comb_22cc-pVTZ//cc-pVDZSDD//SDD0.90137
a Basis sets used for shielding calculation//geometry optimization steps; b average deviation from the general correlation line for P=O phosphorus atoms; c in brackets average deviation from the general correlation line for P=P phosphorus atoms.
Table 2. Empirical scaling factors obtained by the linear regression analysis of calculated and experimental δ31P for the title Pd complexes.
Table 2. Empirical scaling factors obtained by the linear regression analysis of calculated and experimental δ31P for the title Pd complexes.
Level of TheoryR2SlopeInterceptRMSE
PBE0/6-311G(2d,2p)//PBE0/6-31+G(d)0.9541.1766−12.2978.9
PBE0/6-311G(2d,2p)//PBE0/6-311+G(2d)0.9631.1346−13.89628.0
PBE0/def2-TZVP//PBE0/6-311+G(2d)0.9720.92971.12356.9
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Kondrashova, S.A.; Polyancev, F.M.; Latypov, S.K. DFT Calculations of 31P NMR Chemical Shifts in Palladium Complexes. Molecules 2022, 27, 2668. https://doi.org/10.3390/molecules27092668

AMA Style

Kondrashova SA, Polyancev FM, Latypov SK. DFT Calculations of 31P NMR Chemical Shifts in Palladium Complexes. Molecules. 2022; 27(9):2668. https://doi.org/10.3390/molecules27092668

Chicago/Turabian Style

Kondrashova, Svetlana A., Fedor M. Polyancev, and Shamil K. Latypov. 2022. "DFT Calculations of 31P NMR Chemical Shifts in Palladium Complexes" Molecules 27, no. 9: 2668. https://doi.org/10.3390/molecules27092668

Article Metrics

Back to TopTop