Next Article in Journal
Role of Disordered Precursor in L10 Phase Formation in FePt-Based Nanocomposite Magnet
Next Article in Special Issue
The Microscopic Mechanisms Involved in Superexchange
Previous Article in Journal
Revisiting the Potential Functionality of the MagR Protein
Previous Article in Special Issue
Simulating Static and Dynamic Properties of Magnetic Molecules with Prototype Quantum Computers
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Predicting Pt-195 NMR Chemical Shift and 1J(195Pt-31P) Coupling Constant for Pt(0) Complexes Using the NMR-DKH Basis Sets

by
Joyce H. C. e Silva
1,
Hélio F. Dos Santos
2 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 27973-545, RJ, Brazil
2
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
*
Author to whom correspondence should be addressed.
Magnetochemistry 2021, 7(11), 148; https://doi.org/10.3390/magnetochemistry7110148
Submission received: 20 September 2021 / Revised: 28 October 2021 / Accepted: 2 November 2021 / Published: 12 November 2021

Abstract

:
Pt(0) complexes have been widely used as catalysts for important reactions, such as the hydrosilylation of olefins. In this context, nuclear magnetic resonance (NMR) spectroscopy plays an important role in characterising of new structures and elucidating reaction mechanisms. In particular, the Pt-195 NMR is fundamental, as it is very sensitive to the ligand type and the oxidation state of the metal. In the present study, quantum mechanics computational schemes are proposed for the theoretical prediction of the Pt-195 NMR chemical shift and 1J(195Pt–31P) in Pt(0) complexes. The protocols were constructed using the B3LYP/LANL2DZ/def2-SVP/IEF-PCM(UFF) level for geometry optimization and the GIAO-PBE/NMR-DKH/IEF-PCM(UFF) level for NMR calculation. The NMR fundamental quantities were then scaled by empirical procedures using linear correlations. For a set of 30 Pt(0) complexes, the results showed a mean absolute deviation (MAD) and mean relative deviation (MRD) of only 107 ppm and 2.3%, respectively, for the Pt-195 NMR chemical shift. When the coupling constant is taken into account, the MAD and MRD for a set of 33 coupling constants in 26 Pt(0) complexes were of 127 Hz and 3.3%, respectively. In addition, the models were validated for a group of 17 Pt(0) complexes not included in the original group that had MAD/MRD of 92 ppm/1.7% for the Pt-195 NMR chemical shift and 146 Hz/3.6% for the 1J(195Pt–31P).

Graphical Abstract

1. Introduction

Currently, there is a great interest in the study of Pt compounds due to their application as anticancer drugs [1,2] and due to their use as heterogeneous or homogeneous catalysts [3] in modern organic chemistry [4,5]. In particular, Pt(0) complexes are widely used as catalysts in hydrosilylation processes [6,7,8,9,10].
The Pt(0) complexes have coordination numbers (C.N.) 2, 3, and 4 and exhibit linear, trigonal, and tetrahedral geometries, respectively [11,12]. Considering the importance of Pt(0) complexes and their structural diversity, the nuclear magnetic resonance (NMR) spectroscopy for the Pt-195 nucleus is an important tool used for mechanistic studies [13]. Pt-195 NMR can help in structural characterization of the complexes, cis/trans discrimination, stoichiometry, elucidation of reaction mechanisms, etc. [11,12,13].
The studies on NMR of Pt-195 date back to the 1960s, when the effect of structure on the Pt chemical shift was first described [14]. The Pt-195 nucleus is the only active isotope of Pt with spin quantum number I = ½, a relative sensitivity of 9.94 × 10−3 (1H: 1.00), a gyromagnetic ratio γ = 5.768 × 107 rad s−1 T−1, and a natural abundance of 33.8% [13,15]. The Pt-195 NMR chemical shift (δ195Pt) includes values in the range of about 15,000 ppm, from +8000 to −7000 ppm relative to the reference compound [PtCl6]2− in D2O, which are sensitive to the oxidation state of Pt (0, 2, or 4), temperature, solvent, type of ligands in the coordination sphere, and their spatial arrangement [11,13,14,15]. In addition, the one-bond spin–spin coupling constant (SSCC) 1J195Pt–L (L = Ligand) provides important information about the spatial position of the ligands [15].
Theoretical prediction of Pt-195 NMR parameters is a difficult task, as several factors must be taken into account, such as the geometry of the complexes, electronic correlation, basis sets, solvents, and relativistic effects [16,17,18,19,20]. Despite this difficulty, some works on the theoretical prediction of δ195Pt and 1J195Pt–L for Pt(II) and Pt(IV) complexes can be found in the literature [18,19,20,21,22,23,24,25,26]. However, computational studies on NMR prediction of Pt-195 in Pt(0) complexes are still very scarce [27].
In a previous paper [18], we developed an all-electron relativistic triple-zeta doubly polarized basis set (NMR-DKH) for Pt and the main ligand atoms. In addition, an empirical model for predicting Pt-195 NMR chemical shift in Pt(II) complexes was proposed. For a set of 258 Pt(II) complexes, a mean absolute deviation (MAD) and mean relative deviation (MRD) of 168 ppm and 5%, respectively, were obtained at GIAO-PBE/NMR-DKH/IEF-PCM(UFF)//B3LYP/LANL2DZ/def2-SVP/IEF-PCM(UFF) level (double slashes mean that the NMR properties were calculated at GIAO-PBE/NMR-DKH/IEF-PCM(UFF) and the structure at B3LYP/LANL2DZ/def2-SVP/IEF-PCM(UFF)), followed by an empirical scaling procedure. Then, the same computational protocol as in Ref. [18] was used to build an empirical model for predicting 1J(195Pt–15N) for a set of 71 Pt(II) complexes, which yielded a MAD of 36 Hz and an MRD of 10.4% [19]. In the present work, we extend our previous studies by proposing quantum mechanical computational schemes to predict Pt-195 NMR chemical shift and 1J(195Pt–31P) for Pt(0) complexes.

2. Theoretical Methodology

The geometries of all Pt(0) complexes (see Supplementary Materials) studied in the present paper were optimized and characterized as a local minimum on the potential energy surface (PES) by the positive (real) values of all harmonic mode frequencies at density functional theory (DFT) level with the hybrid functional B3LYP [28,29,30], the effective core potential (ECP) LANL2DZ [31,32] for Pt and the def2-SVP [33] basis set for ligands atoms (B3LYP/LANL2DZ/def2-SVP/IEF-PCM(UFF)). The solvent effect was considered in both geometry optimization and NMR calculations using the integral equation formalism for polarized continuum method (IEF-PCM) and the radii were set as for UFF force field [34]. For each Pt(0) complex, the solvent considered in the calculations is the same as used in the experimental measurements. All Pt(0) complexes studied in the present paper have a singlet electronic ground state [12].
The Pt-195 NMR shielding constant (σ195Pt) and the 1J(195Pt–31P) were calculated from the gauge-independent atomic orbitals (GIAO) [35,36] method at DFT level using the GGA PBE [37] functional with the NMR-DKH [18] basis sets (GIAO-PBE/NMR-DKH/IEF-PCM(UFF)). The computational protocol (NMR/geometry) is represented as in our previous papers [18,19], namely, GIAO-PBE/NMR-DKH/IEF-PCM(UFF)//B3LYP/LANL2DZ/def2-SVP/IEF-PCM(UFF). All calculations were carried out in the GAUSSIAN 09 program Revision D.01 [38].
The NMR fundamental quantities, σ195Pt and the 1J(195Pt–31P), were used as independent variables to fit the scaling models. The Equation (1) was used to predict δ195Pt, in which the “a” and “b” parameters were fitted using a standard linear correlation for a set of 30 Pt(0) complexes for which experimental data are found in the literature.
δ Pt calc . 195 = a × σ Pt calc . 195 + b
For 1J(195Pt–31P), the Equation (2) was adjusted considering a set of 33 coupling constants in 26 Pt(0) complexes.
J scal . 1 ( Pt 195 P 31 ) = a × J calc . 1 ( Pt 195 P 31 ) + b
In the last part, we validated the proposed scaling models using a set of 17 Pt(0) complexes not included in the original set.

3. Results and Discussion

As shown in our previous papers [18,19,20], the inclusion of relativistic effects is fundamental for the theoretical prediction of NMR parameters of the heavy Pt-195 nucleus. Relativistic Hamiltonians are not always available in computational packages, so the construction of nonrelativistic computational protocols must be of great interest and useful if properly scaled. In the present work, we present computational protocols using a nonrelativistic Hamiltonian for predicting δ195Pt and 1J(195Pt–31P) in Pt(0) complexes.
First, we discuss the Pt-195 NMR chemical shift for the Pt(0) complexes. To partially recover the relativistic effects and account for the intrinsic errors of the computational methods, an empirical scaling model (Equation (1)) was constructed from the linear correlation of the calculated σ195Pt and the experimental δ195Pt for a set of 30 Pt(0) complexes showing coordination numbers 2 and 3. The linear regression is shown in Figure 1a and the parameters a and b are given in Table 1 (Model 1). It can be observed that the obtained angular coefficient (a = −0.8277) is close to −1, indicating that the proposed empirical Model 1 is physically consistent [18]. Moreover, the coefficient of determination (R2) was 0.95, which confirms the quality of the proposed empirical statistical model.
Table 2 shows the predicted Pt-195 NMR chemical shift using Model 1 for the 30 Pt(0) complexes. Considering all the complexes, the MAD and MRD were 107 ppm and 2.3%, respectively, showing that Model 1 has satisfactory predictive power. When the complexes with C.N. 2 and 3 are evaluated separately, the MAD and the MRD are 48 ppm and 0.7% (four linear complexes with C.N. 2), and 120 ppm and 2.6% (25 trigonal complexes with C.N. 3).
A comparison between the calculated and experimental Pt-195 NMR chemical shift is shown in Figure 2a. Although the ranges of chemical shift within each group (C.N. 2 or 3) are small, Model 1 correctly predicted the experimental trends and proved to be very sensitive to small structural changes. Moreover, considering that one of the main applications of Pt-195 NMR in Pt(0) complexes is the determination of the stoichiometry (C.N.) of the synthesized complexes, the proposed empirical Model 1 can assist the experimentalist in this prediction.
As for the prediction of SSCC, of the 30 selected Pt(0) complexes, experimental data are available for 26 structures. Within these 26 Pt(0) complexes, a total of 33 coupling constants have been calculated. All calculated values of 1J(195Pt–31P) were underestimated compared to the experimental data (Table 2), showing MAD and MRD of 1250 Hz and 34.0%, respectively (labeled as Model 0). Therefore, to correct the calculated coupling constants, a scaling procedure similar to the one used for the 1J(195Pt–15N) calculation in a previous work [19] was used, Equation (2).
A linear correlation between the calculated and experimental 1J(195Pt–31P) was performed for the set of 33 available coupling constants. The parameters obtained from the linear regression (Figure 1b) are listed in Table 1 (Models 2 and 3). For Model 2, which includes structures with C.N. 2 and 3, MAD and MRD were 127 Hz and 3.3%, respectively. For Model 3, which includes only C.N. 3 (29 coupling constants of the 22 complexes of Pt(0)), MAD and MRD were 58 Hz and 1.6%, respectively, indicating excellent predictive ability of the proposed model. Comparing Models 2 and 3 with Model 0, the improvement of the results with scaling is evident. The overall MRD decreases from 34% to ~3%. Figure 2b shows a comparison between the calculated and experimental 1J(195Pt–31P). From the results, it can be seen that both Model 2 and Model 3 were able to adequately predict the small variations observed experimentally in the values of the coupling constants.
In the final part of the study, the scaling Models were validated for a set of 17 Pt(0) complexes that were not included in the original set (Table 3). For this set of molecules, experimental chemical shifts are available for all 17 complexes and experimental 1J(195Pt–31P) for nine complexes (18 coupling constants). For the Pt-195 NMR chemical shift, the MAD and MRD were 92 ppm and 1.7%, respectively, when Model 1 was used. For 1J(195Pt –31P), MAD and MRD were 146 Hz and 3.6%, respectively, with Model 2 and 98 Hz and 2.6%, respectively, with Model 3. Both Model 2 and Model 3 were able to distinguish the coupling constants in cis- and trans-positions to the COOR group in the nine Pt(0) complexes. Despite the slightly larger error obtained for the validation group (Figure 3), the predictive capacity of the Models 1 (δ195Pt), 2, and 3 (1J(195Pt–31P)) is still satisfactory and could be used to assist the experimentalists in the structural characterization.
Among the compounds used for validation, we highlight the Pt(0)–carbene complex ([Pt(ICy)(dvtms)], dvtms = divinyltetramethyldisiloxane (Figure 4) which is a selective and efficient hydrosilylation catalyst [10]. The calculated δ195Pt was −5270 ppm, showing an absolute deviation (AD) of 73 ppm and a relative deviation (RD) of only 1.4% compared to the experimental data (δ195Pt = −5343 ppm).

4. Conclusions

In this study, quantum mechanics computational schemes have been proposed for predicting Pt-195 NMR chemical shift and spin–spin coupling constant 1J(195Pt–31P) for a series of linear and trigonal-planar in Pt(0) complexes. The NMR fundamental quantities, namely the shielding constant (σ195Pt) and the coupling constant (1J(195Pt–31P)), were calculated at the PBE/NMR-DKH/IEF-PCM(UFF)//B3LYP/LANL2DZ/def2-SVP/IEF-PCM(UFF) level and scaled using linear models. The Model 1, δ Pt calc 195 = 0.8277 × σ Pt calc 195 2566 , predicted the Pt-195 NMR chemical shift with MAD and MRD of 107 ppm and 2.3%, respectively, for a set of 30 Pt(0) complexes with C.N. 2 and 3. For the coupling constant, Model 2, J scal . 1 ( Pt 195 P 31 ) = 0.5508 × J calc . 1 ( Pt 195 P 31 ) + 2373.3 , predicted 1J(195Pt–31P) with MAD and MRD of 127 Hz and 3.3%, respectively, for a set of 33 coupling constants in 26 Pt(0) complexes. When only C.N. 3 is considered, Model 3, J scal . 1 ( Pt 195 P 31 ) = 0.9482 × J calc . 1 ( Pt 195 P 31 ) + 1478.7 , was better than Model 2, with MAD and MRD of 58 Hz and 1.6% for a set of 29 coupling constants in 22 trigonal Pt(0) complexes. We validated all Models for a set of 17 Pt(0) complexes that were not included in the original set. The results were (MAD/MRD): 92 ppm/1.7% (Model 1), 146 Hz/3.6% (Model 2) and 98 Hz/2.6% (Model 3).
In summary, the scaling models proposed here could be useful methods for predicting Pt-195 NMR parameters in Pt(0) complexes. This extends our previous protocols for predicting similar properties for Pt(II) complexes using the same quantum mechanical theory to calculate the fundamental quantities. In addition, the present results also support the use of the NMR-DKH basis set, specifically constructed for the calculation of NMR spectra of heavy metal complexes, in particular Pt complexes.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/magnetochemistry7110148/s1, The xyz coordinates of the optimized geometries of 47 Pt(0) complexes studied in the present paper are in the Supplementary Materials.

Author Contributions

Conceptualization, J.H.C.e.S., H.F.D.S. and D.F.S.P.; Methodology, J.C.H.e.S., H.F.D.S. and D.F.S.P.; Software, J.C.H.e.S., H.F.D.S. and D.F.S.P.; Validation, D.F.S.P.; Formal Analysis, J.H.C.e.S., H.F.D.S. and D.F.S.P.; Investigation, J.H.C.e.S., H.F.D.S. and D.F.S.P.; Resources, H.F.D.S. and D.F.S.P.; Writing, J.H.C.e.S., H.F.D.S. and D.F.S.P.; Supervision, H.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

The authors would like to thank the Brazilian agency FAPERJ (E-26/201.737/2017—BOLSA and E-26/010.002261/2019—EMERGENTES) 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.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Acknowledgments

DFSP thanks the FAPERJ for supporting to the NQTCM laboratory. HFDS thanks the FAPEMIG and CNPq for continuing support to the NEQC laboratory.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Dilruba, S.; Kalayda, G.V. Platinum-Based Drugs: Past, Present and Future. Cancer Chemother. Pharmacol. 2016, 77, 1103–1124. [Google Scholar] [CrossRef]
  2. Deo, K.M.; Ang, D.L.; McGhie, B.; Rajamanickam, A.; Dhiman, A.; Khoury, A.; Holland, J.; Bjelosevic, A.; Pages, B.; Gordon, C.; et al. Platinum Coordination Compounds with Potent Anticancer Activity. Coord. Chem. Rev. 2018, 375, 148–163. [Google Scholar] [CrossRef]
  3. Dedieu, A. Theoretical Studies in Palladium and Platinum Molecular Chemistry. Chem. Rev. 2000, 100, 543–600. [Google Scholar] [CrossRef]
  4. Li, C.-J.; Chen, L. Organic Chemistry in Water. Chem. Soc. Rev. 2006, 35, 68–82. [Google Scholar] [CrossRef]
  5. Li, C.-J. Organic Reactions in Aqueous Media with a Focus on Carbon−Carbon Bond Formations: A Decade Update. Chem. Rev. 2005, 105, 3095–3166. [Google Scholar] [CrossRef] [PubMed]
  6. Silbestri, G.F.; Flores, J.C.; de Jesús, E. Water-Soluble N-Heterocyclic Carbene Platinum(0) Complexes: Recyclable Catalysts for the Hydrosilylation of Alkynes in Water at Room Temperature. Organometallics 2012, 31, 3355–3360. [Google Scholar] [CrossRef]
  7. Nakajima, Y.; Shimada, S. Hydrosilylation Reaction of Olefins: Recent Advances and Perspectives. RSC Adv. 2015, 5, 20603–20616. [Google Scholar] [CrossRef]
  8. Sprengers, J.W.; de Greef, M.; Duin, M.A.; Elsevier, C.J. Stable Platinum(0) Catalysts for Catalytic Hydrosilylation of Styrene and Synthesis of [Pt(Ar-Bian)(H2-Alkene)] Complexes. Eur. J. Inorg. Chem. 2003, 2003, 3811–3819. [Google Scholar] [CrossRef]
  9. Troegel, D.; Stohrer, J. Recent Advances and Actual Challenges in Late Transition Metal Catalyzed Hydrosilylation of Olefins from an Industrial Point of View. Coord. Chem. Rev. 2011, 255, 1440–1459. [Google Scholar] [CrossRef]
  10. Markó, I.E.; Stérin, S.; Buisine, O.; Mignani, G.; Branlard, P.; Tinant, B.; Declercq, J.-P. Selective and Efficient Platinum(0)-Carbene Complexes As Hydrosilylation Catalysts. Science 2002, 298, 204–206. [Google Scholar] [CrossRef]
  11. Pregosin, P.S. Platinum NMR Spectroscopy. Annu. Rep. NMR Spectrosc. 1986, 17, 285–349. [Google Scholar] [CrossRef]
  12. Georgii, I.; Mann, B.E.; Taylor, B.F.; Musco, A. 195Pt n.m.r. Spectra of Some [Pt(PR3)n] Complexes. Inorg. Chim. Acta 1984, 86, L81–L82. [Google Scholar] [CrossRef]
  13. Still, B.M.; Kumar, P.G.A.; Aldrich-Wright, J.R.; Price, W.S. 195Pt NMR—Theory and Application. Chem. Soc. Rev. 2007, 36, 665–686. [Google Scholar] [CrossRef]
  14. Pregosin, P.S. Platinum-195 Nuclear Magnetic Resonance. Coord. Chem. Rev. 1982, 44, 247–291. [Google Scholar] [CrossRef]
  15. Priqueler, J.R.L.; Butler, I.S.; Rochon, F.D. An Overview of 195Pt Nuclear Magnetic Resonance Spectroscopy. Appl. Spectrosc. Rev. 2006, 41, 185–226. [Google Scholar] [CrossRef]
  16. Vaara, J. Theory and Computation of Nuclear Magnetic Resonance Parameters. Phys. Chem. Chem. Phys. 2007, 9, 5399. [Google Scholar] [CrossRef] [PubMed]
  17. 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]
  18. 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]
  19. 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]
  20. da Silva Paschoal, D.F.; da Silva Gomes, M.; Machado, L.P.N.; dos Santos, H.F. Basis Sets for Heavy Atoms. In Basis Sets in Computational Chemistry. Lecture Notes in Chemistry; Perlt, E., Ed.; Springer: Cham, Switzerland, 2021; Volume 107, pp. 183–214. [Google Scholar]
  21. Semenov, V.A.; Samultsev, D.O.; Rusakova, I.L.; Krivdin, L.B. Computational Multinuclear NMR of Platinum Complexes: A Relativistic Four-Component Study. J. Phys. Chem. A 2019, 123, 4908–4920. [Google Scholar] [CrossRef]
  22. Sterzel, M.; Autschbach, J. Toward an Accurate Determination of 195Pt Chemical Shifts by Density Functional Computations: The Importance of Unspecific Solvent Effects and the Dependence of Pt Magnetic Shielding Constants on Structural Parameters. Inorg. Chem. 2006, 45, 3316–3324. [Google Scholar] [CrossRef] [PubMed]
  23. Semenov, V.A.; Rusakov, Y.Y.; Samultsev, D.O.; Krivdin, L.B. Geometries and NMR Properties of Cisplatin and Transplatin Revisited at the Four-Component Relativistic Level. Mendeleev Commun. 2019, 29, 315–317. [Google Scholar] [CrossRef]
  24. Sutter, K.; Truflandier, L.A.; Autschbach, J. NMR J-Coupling Constants in Cisplatin Derivatives Studied by Molecular Dynamics and Relativistic DFT. ChemPhysChem 2011, 12, 1448–1455. [Google Scholar] [CrossRef] [PubMed]
  25. Tsipis, A.C.; Karapetsas, I.N. Accurate Prediction of 195Pt NMR Chemical Shifts for a Series of Pt(II) and Pt(IV) Antitumor Agents by a Non-Relativistic DFT Computational Protocol. Dalton Trans. 2014, 43, 5409–5426. [Google Scholar] [CrossRef] [PubMed]
  26. Gilbert, T.M.; Ziegler, T. Prediction of 195 Pt NMR Chemical Shifts by Density Functional Theory Computations: The Importance of Magnetic Coupling and Relativistic Effects in Explaining Trends. J. Phys. Chem. A 1999, 103, 7535–7543. [Google Scholar] [CrossRef]
  27. Berkefeld, A.; Reimann, M.; Hörner, G.; Kaupp, M.; Schubert, H. C–P vs C–H Bond Cleavage of Triphenylphosphine at Platinum(0): Mechanism of Formation, Reactivity, Redox Chemistry, and NMR Chemical Shift Calculations of a μ-Phosphanido Diplatinum(II) Platform. Organometallics 2020, 39, 443–452. [Google Scholar] [CrossRef]
  28. Becke, A.D. Density-functional Thermochemistry. III. The Role of Exact Exchange. J. Chem. Phys. 1993, 98, 5648–5652. [Google Scholar] [CrossRef] [Green Version]
  29. 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] [Green Version]
  30. 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]
  31. Hay, P.J.; Wadt, W.R. Ab Initio Effective Core Potentials for Molecular Calculations. Potentials for the Transition Metal Atoms Sc to Hg. J. Chem. Phys. 1985, 82, 270–283. [Google Scholar] [CrossRef]
  32. Hay, P.J.; Wadt, W.R. Ab Initio Effective Core Potentials for Molecular Calculations. Potentials for K to Au Including the Outermost Core Orbitals. J. Chem. Phys. 1985, 82, 299–310. [Google Scholar] [CrossRef]
  33. 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. [Google Scholar] [CrossRef] [PubMed]
  34. 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] [PubMed]
  35. 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]
  36. 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]
  37. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett 1997, 77, 3865. [Google Scholar] [CrossRef] [Green Version]
  38. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G.A.; et al. Gaussian ∼09 Revision D.01. Available online: https://www.scienceopen.com/document?vid=839f33cc-9114-4a55-8f1a-3f1520324ef5 (accessed on 20 September 2021).
  39. Mann, B.E.; Musco, A. A 31P Nuclear Magnetic Resonance Investigation of the Structure, Equilibria, and Kinetics of [Pt(PR3)n] in Solution. J. Chem. Soc. Dalton Trans. 1980, 776–785. [Google Scholar] [CrossRef]
  40. Benn, R.; Michael Büch, H.; Reinhardt, R.-D. Heavy Metal Spin- ½ Nuclei. Platinum-195 Relaxation Times in Phosphorus-Platinum(0) Compounds and Their Dependence on the Geometry of the Complex. Magn. Reson. Chem. 1985, 23, 559–564. [Google Scholar] [CrossRef]
  41. Koie, Y.; Shinoda, S.; Saito, Y. Platinum-195 Nuclear Magnetic Resonance Study of Platinum(0) Complexes Containing a Series of Acetylenes. J. Chem. Soc. Dalton Trans. 1981, 1082–1088. [Google Scholar] [CrossRef]
  42. Kennedy, J.D.; McFarlane, W.; Puddephatt, R.J.; Thompson, P.J. Magnetic Double-Resonance Studies of Platinum-195 Chemical Shifts in Organoplatinum Compounds. J. Chem. Soc. Dalton Trans. 1976, 874–879. [Google Scholar] [CrossRef]
  43. Wrackmeyer, B.; Klimkina, E.V.; Schmalz, T.; Milius, W. Synthesis, NMR Spectroscopic Characterization and Structure of a Divinyldisilazane-(Triphenylphosphine)Platinum(0) Complex: Observation of Isotope-Induced Chemical Shifts 1Δ12/13C(195Pt). Magn. Reson. Chem. 2013, 51, 283–291. [Google Scholar] [CrossRef] [PubMed]
  44. Ruiz-Varilla, A.M.; Baquero, E.A.; Silbestri, G.F.; Gonzalez-Arellano, C.; de Jesús, E.; Flores, J.C. Synthesis and Behavior of Novel Sulfonated Water-Soluble N-Heterocyclic Carbene (η4-Diene) Platinum(0) Complexes. Dalton Trans. 2015, 44, 18360–18369. [Google Scholar] [CrossRef] [PubMed]
  45. Buchner, M.R.; Bechlars, B.; Wahl, B.; Ruhland, K. Influence of Electronically and Sterically Tunable Cinnamate Ligands on the Spectroscopic Properties and Reactivity of Bis(Triphenylphosphine)Platinum(0) Olefin Complexes. Organometallics 2013, 32, 1643–1653. [Google Scholar] [CrossRef]
Figure 1. Linear regression between: (a) calculated σ195Pt and experimental δ195Pt for a set of 30 Pt(0) complexes (Model 1); (b) calculated and experimental 1J(195Pt–31P) for a set of 33 coupling constants in 26 linear and trigonal-planar Pt(0) complexes (Model 2); and (c) calculated and experimental 1J(195Pt–31P) for a set of 29 coupling constants in 22 trigonal-planar Pt(0) complexes (Model 3).
Figure 1. Linear regression between: (a) calculated σ195Pt and experimental δ195Pt for a set of 30 Pt(0) complexes (Model 1); (b) calculated and experimental 1J(195Pt–31P) for a set of 33 coupling constants in 26 linear and trigonal-planar Pt(0) complexes (Model 2); and (c) calculated and experimental 1J(195Pt–31P) for a set of 29 coupling constants in 22 trigonal-planar Pt(0) complexes (Model 3).
Magnetochemistry 07 00148 g001
Figure 2. Relation between calculated and experimental: (a) δ195Pt (ppm)—Model 1; (b) 1J(195Pt–31P)—Models 2 and 3.
Figure 2. Relation between calculated and experimental: (a) δ195Pt (ppm)—Model 1; (b) 1J(195Pt–31P)—Models 2 and 3.
Magnetochemistry 07 00148 g002
Figure 3. Correlation between the calculated and experimental values for the validation group of the models. (a) δ195Pt for the set of 17 Pt(0) complexes (Model 1); (b) 1J(195Pt–31P) for a set of 18 coupling constants in 9 trigonal-planar Pt(0) complexes (Model 2); and (c) 1J(195Pt–31P) for a set of 18 coupling constants in 9 trigonal-planar Pt(0) complexes (Model 3).
Figure 3. Correlation between the calculated and experimental values for the validation group of the models. (a) δ195Pt for the set of 17 Pt(0) complexes (Model 1); (b) 1J(195Pt–31P) for a set of 18 coupling constants in 9 trigonal-planar Pt(0) complexes (Model 2); and (c) 1J(195Pt–31P) for a set of 18 coupling constants in 9 trigonal-planar Pt(0) complexes (Model 3).
Magnetochemistry 07 00148 g003
Figure 4. Pt(0)-carbene complex ([Pt(Icy)(dvtms)].
Figure 4. Pt(0)-carbene complex ([Pt(Icy)(dvtms)].
Magnetochemistry 07 00148 g004
Table 1. Empirical scaling models obtained to predict δ195Pt and 1J(195Pt–31P) in Pt(0) complexes.
Table 1. Empirical scaling models obtained to predict δ195Pt and 1J(195Pt–31P) in Pt(0) complexes.
ModelsLinear Regression ModelsR2
Model 1 a δ Pt calc 195 = 0.8277 × σ Pt calc 195 2566 0.9522
Model 2 b J scal . 1 ( Pt 195 P 31 ) = 0.5508 × J calc . 1 ( Pt 195 P 31 ) + 2373.3 0.7976
Model 3 c J scal . 1 ( Pt 195 P 31 ) = 0.9482 × J calc . 1 ( Pt 195 P 31 ) + 1478.7 0.9385
a Model 1 constructed for linear (C.N. = 2) and trigonal (C.N. = 3) Pt(0) complexes using a set of 30 complexes. b Model 2 constructed for linear (C.N. = 2) and trigonal (C.N. = 3) Pt(0) complexes using a set of 33 coupling constant in 26 complexes. c Model 3 constructed for trigonal (C.N. = 3) Pt(0) complexes using a set of 29 coupling constant in 22 complexes. The fundamental quantities, δ195Pt, and 1J(195Pt–31P), were calculated at GIAO-PBE/NMR-DKH/IEF-PCM(UFF)//B3LYP/LANL2DZ/def2-SVP/IEF-PCM(UFF) level.
Table 2. Calculated δ195Pt, in ppm, and 1J(195Pt–31P), in Hz, for the Pt(0) complexes used in the construction of the proposed empirical models.
Table 2. Calculated δ195Pt, in ppm, and 1J(195Pt–31P), in Hz, for the Pt(0) complexes used in the construction of the proposed empirical models.
δ195Pt (ppm)1J(195Pt–31P) (Hz)
Pt(0) ComplexesSolventModel 1Expt.Model 0Model 2Model 3Expt.
Linear Geometry (C.N. = 2)
1[Pt(PPri3)2]Toluene−6557−6607 a36654392-4104 b
2[Pt(Pcy3)2], cy = cyclohexilToluene−6583−6555 a37424435-4120 b
3[Pt(PBut2Ph)2Toluene−6434−6526 a38984520-4592 b
4[Pt(PBut3)2THF-d8−6457−6479 c40654612-4399 c
MAD/MRD (C.N. = 2) 48/0.7% 461/11%222/5.3%-
Trigonal Geometry (C.N. = 3)
5[Pt(PEt3)3]Toluene−4768−4526 a2897396942264188 b
6[Pt(PBun3)3]Toluene−4761−4511 a2916398042444211 b
7[Pt(P(CH2Ph)3)3]Toluene−4790−4439 a3086407344054377 b
8[Pt(Pcy3)3], cy = cyclohexilToluene−4878−4567 a----
9[Pt(PPh3)2(PhCy≡CPh)]CD2Cl2−4642−4741 d2186357735513452 d
10[Pt(PPh3)2(PhC≡CMe)]CD2Cl2−4630−4727 d2083352034543377 d
--2159356235263454 d
11[Pt(PPh3)2(PhC≡CCO2Me)]CD2Cl2−4602−4710 d2130354734993403 d
--2196358335613741 d
12[Pt(PPh3)2(PhC≡CH)]CD2Cl2−4668−4690 d2121354134903464 d
--2158356235253547 d
13[Pt(PPh3)2(EtC≡CEt)]CD2Cl2−4669−4689 d2189357935543425 d
14[Pt(PPh3)2(MeC≡CCO2Me)]CD2Cl2−4719−4682 d1957345133343366 d
--2312364736713803 d
15[Pt(PPh3)2(MeC≡CMe)]CD2Cl2−4702−4674 d2086352234573420 d
16[Pt(PPh3)2(HC≡CH)]CD2Cl2−4607−4658 d2242360836043626 d
17[Pt(PPh3)2(EtO2CC≡CCO2Et)]CD2Cl2−4586−4655 d2203358735673722 d
18[Pt(PPh3)2(MeO2CC≡CCO2Me)]CD2Cl2−4537−4653 d2342366437003722 d
19[Pt(PPh3)2(F3CC≡CCF3)]CD2Cl2−4576−4645 d2202358635663595 d
20[Pt(PPh3)2(PhC≡CCN)]CD2Cl2−4549−4640 d1919343032983336 d
--2506375438553772 d
21[Pt(PPh3)2(F3CH2CO2CC≡CCO2CH2CF3)]CD2Cl2−4488−4626 d2346366537033726 d
22[Pt(PPh3)2(MeC≡CCN)]CD2Cl2−4590−4598 d1923343333023303 d
--2550377838963864 d
23[Pt(PPh3)2(HC≡CCN)]CD2Cl2−4539−4573 d2026348934003434 d
--2581379539263887 d
24[Pt(PPh3)2(NCC≡CCN)]CD2Cl2−4395−4586 d2320365136783696 d
25[Pt(F3CC≡CCF3)(PPh3)2]CDCl3−4570−4645 e2222359735853590 e
26[Pt(F2C=CF2)(PPh3)2]CDCl3−4728−4791 e----
27[Pt(H2C=CH2)(PPh3)2]CDCl3−5033−5065e----
28[Pt(P(O-o-tolyl)3)]CD2Cl2−5049−4858 c----
29[Pt(PPh3)3]THF-d8−4804−4583 c2999402543224455 c
30[Pt(dvtms)(PPh3)]CDCl3−5358−5572 f2304364336643609 f
MAD/MRD (C.N. = 3)-120/2.6%-1359/37%114/3.0%58/1.6%-
MAD/MRD (All Pt(0) complexes studied)-107/2.3%-1250/34%127/3.3%58/1.6%-
Experimental values obtained from: a Georgii et al. [12]; b Mann et al. [39]; c Benn et al. [40]; d Koie et al. [41]; e Kennedy et al. [42]; f Wrackmeyer et al. [43]. MAD = mean absolute deviation— MAD = 1 n k k = 1 n k | δ P 195 t calc . δ Pt expt . 195 | MRD = mean relative deviation— MRD = ( 1 n k k = 1 n k | δ Pt calc . 195 δ P 195 t expt . δ Pt expt . 195 | ) × 100 % . Model 0. Unscaled model. Model 1. δ Pt calc . 195 = 0.8277 × σ Pt calc . 195 2566 . Model 2. J scal . 1 ( Pt 195 P 31 ) = 0.5508 × J calc . 1 ( Pt 195 P 31 ) + 2373.3 . Model 3. J scal . 1 ( Pt 195 P 31 ) = 0.9482 × J calc . 1 ( Pt 195 P 3.1 ) + 1478.7 .
Table 3. Calculated δ195Pt in ppm, and 1J(195Pt–31P) in Hz, for the Pt(0) complexes used in the validation of the proposed empirical models.
Table 3. Calculated δ195Pt in ppm, and 1J(195Pt–31P) in Hz, for the Pt(0) complexes used in the validation of the proposed empirical models.
δ195Pt (ppm)1J(195Pt–31P) (Hz)
Pt(0) ComplexesSolventModel 1Expt.Model 0Model 2Model 3Expt.
Trigonal Geometry (C.N. = 3)
31[Pt(ICy)(dvtms)]CDCl3−5270−5343 a----
32[Pt(Mes-NHC-Prn-SO3Na)(dvtms)]DMSO−5212−5352 b----
33[Pt(IPr-4-SO3Na)(dvtms)]DMSO−5141−5332 b----
34[Pt(IXy-4-SO3Na)(dvtms)] D2O−5160−5336 c----
35anti-[Pt(IMes-4-SO3Na)(dvtms)]D2O−5102−5342 c----
36syn-[Pt(SIMes-4-SO3Na)(dvtms)]D2O−5246−5372 c----
37[Pt(Mes-NHC-Prn-SO3Na)(AE)]DMSO−5411−5597 c----
38[Pt(IXy-4,4-SO3Na)(AE)]DMSO−5374−5562 c----
39[Pt(PPh3)2(MeOPhHC=CHCOOPhOMe)]Benzene−5033−5044 d2412370237663625 d
trans/cis COOR --2797391441314218 d
40[Pt(PPh3)2(PhHC=CHCOOPhMe)] Benzene−4985−5053 d2502375138513645 d
trans/cis COOR --2741388340774170 d
41[Pt(PPh3)2(NO2PhHC=CHCOOPhNO2)] Benzene−5057−5047 d2408370037623682 d
trans/cis COOR --2731387840684095 d
42[Pt(PPh3)2(NO2PhHC=CHCOOPh)] Benzene−5026−5047 d2505375338543737 d
trans/cis COOR --2656383639974039 d
43[Pt(PPh3)2(MePhHC=CHCOOPh)] Benzene−5080−5052 d2453372438043644 d
trans/cis COOR --2736388040734207 d
44[Pt(PPh3)2(PhHC=CHCOOPh)] Benzene−4970−5053 d2499375038483642 d
trans/cis COOR --2757389240934176 d
45[Pt(PPh3)2(NO2PhHC=CHCOOPhOMe)] Benzene−5040−5049 d2495374838453742 d
trans/cis COOR --2639382739814032 d
46[Pt(PPh3)2(NO2PhHC=CHCOOPhMe)] Benzene−5041−5048 d2483374138333742 d
trans/cis COOR --2644383039864033 d
47[Pt(PPh3)2(NO2PhHC=CHCOOPri)] Benzene−5045−5051 d2513375738613774 d
trans/cis COOR --2633382439753993 d
MAD/MRD (Pt(0) complexes studied)-92/1.7%-1311/34%146/3.6%98/2.6%-
Experimental values obtained from: a Markó et al. [10]; b Silbestri et al. [6]; c Ruiz-Varilla et al. [44]; d Buchner et al. [45]. MAD = mean absolute deviation— MAD = 1 n k k = 1 n k | δ P 195 t calc . δ P 195 t expt . | MRD = mean relative deviation— MRD = ( 1 n k k = 1 n k | δ P 195 t calc . δ P 195 t expt . δ P 195 t expt . | ) × 100 % . Model 0. Unscaled model. Model 1. δ Pt calc . 195 = 0.8277 × σ Pt calc . 195 2566 . Model 2. J scal . 1 ( Pt 195 P 31 ) = 0.5508 × J calc . 1 ( Pt 195 P 31 ) + 2373.3 . Model 3. J scal . 1 ( Pt 195 P 31 ) = 0.9482 × J calc . 1 ( Pt 195 P 31 ) + 1478.7 .
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

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. https://doi.org/10.3390/magnetochemistry7110148

AMA Style

e Silva JHC, Dos Santos HF, Paschoal DFS. 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(11):148. https://doi.org/10.3390/magnetochemistry7110148

Chicago/Turabian Style

e Silva, Joyce H. C., Hélio F. Dos Santos, and Diego F. S. Paschoal. 2021. "Predicting Pt-195 NMR Chemical Shift and 1J(195Pt-31P) Coupling Constant for Pt(0) Complexes Using the NMR-DKH Basis Sets" Magnetochemistry 7, no. 11: 148. https://doi.org/10.3390/magnetochemistry7110148

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