Next Article in Journal
Methyl 2-((3-(3-methoxyphenyl)-4-oxo-3,4-dihydroquinazolin-2-yl)thio)acetate
Previous Article in Journal
3-Phenyl-3H-naphtho[1,2-e][1,2,3]oxadiazine
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Short Note

[N,N′-Di-tert-butyl-P,P-diphenylphosphinimidic Amidato-κNN′]chlorosilicon-κSi-tetracarbonyliron

Department of Chemistry, Graduate School of Science and Engineering, Saitama University, 255 Shimo-okubo, Sakura-ku, Saitama 338-8570, Japan
*
Author to whom correspondence should be addressed.
Molbank 2022, 2022(3), M1433; https://doi.org/10.3390/M1433
Submission received: 1 August 2022 / Revised: 14 August 2022 / Accepted: 20 August 2022 / Published: 25 August 2022
(This article belongs to the Section Structure Determination)

Abstract

:
The title complex {[Ph2P(tBuN)2](Cl)Si:->Fe(CO)4} (2) was synthesized via the reaction of chlorosilylene [Ph2P(tBuN)2]SiCl (1), supported by an iminophosphonamide ligand with Fe(CO)5 in THF. The molecular structure of 2 was fully characterized by NMR (1H, 13C, 29Si, and 31P) and IR spectroscopies, as well as single-crystal X-ray diffraction (SCXRD) analysis. In the SCXRD analysis of 2, the silylene ligand was located in the axial positions of the coordination sphere of the central iron atom and other sites were occupied by carbonyl ligands.

1. Introduction

The chemistry of stable divalent silicon(II) species, silylenes, is currently one of the most intensively researched areas of basic silicon chemistry [1,2,3]. In this context, a series of Lewis-base-stabilized silylenes can serve as an ancillary ligand in the coordination chemistry because they can behave as a strong nucleophile [4,5,6]. In particular, the coordination chemistry of silylenes, supported by N,N′-chelating amidinate ligands, has developed dramatically over the past few years. Indeed, numerous transition metal complexes utilizing amidinato–silylene ligands have been investigated [7,8]. The resulting complexes were found to be capable of acting as catalysts for transforming organic molecules [5,6] or activating inactive small molecules [9,10]. Among them, amidinato–silylene–iron complexes are particularly attractive candidates for catalysis due to the low cost and high natural abundance of iron, as well as their non-toxic and biocompatible properties [11,12,13,14,15,16,17,18,19,20].
Our group has been engaged in the chemistry of p-block elements supported by N,N′-chelating iminophosphonamides [R2P(NR′)2] [21,22,23,24,25,26]. In particular, chlorosilylenes [Ph2P(NR)2SiCl] (R = tBu and 2,6-iPr2C6H3) show the unprecedented complexation ability of transition metals and unique nucleophile behavior due to their strong σ-donating properties [21,22,23,26]. In this contribution, we present the synthesis and structural characterization of silylene–tetracarbonyliron(0) complex {[Ph2P(tBuN)2](Cl)Si:->Fe(CO)4}, wherein the silylene(II) center is stabilized by an N,N′-di-tert-butyliminophosphonamide ligand.

2. Results and Discussion

The complexation of chlorosilylene 1 with Fe(CO)5 proceeded readily at room temperature in THF to afford the corresponding silylene–tetracarbonyliron(0) complex 2 as yellow crystals in 69% yield with the release of a molecule of CO (Scheme 1).
Single crystals of 2 were obtained from a saturated THF solution at −10 °C. The molecular structure of 2 was unequivocally determined using single-crystal X-ray diffraction (SCXRD) analysis. The ORTEP is depicted in Figure 1, and selected bond lengths and angles are summarized in Table 1. Complex 2 crystallized in the monoclinic space group P21/c with a molecule of THF per unit cell. The solvate THF molecules in 2 were disordered in two positions with the relative occupancy of 0.722(8) and 0.278(8). The SCXRD analysis of 2 reveals that the silylene ligand was located in the axial positions of the coordination sphere of the central iron atom and other sites were occupied by carbonyl ligands. The chlorine atom on the silicon atom and an equatorial carbonyl ligand were trans to one another with respect to the Si–Fe bond. The geometry of five-coordinated compounds can be evaluated using the angular structure parameter τ [τ = (β − α)/60] [27]. This parameter can be readily calculated using the two largest basal angles (α and β) around the central atom in the five-coordinated compound (Figure 2). Thus, τ = 1 had a trigonal bipyramidal structure with D3 h symmetry, while the square pyramidal structure with C4v symmetry had τ = 0. The calculated τ value for 2 was 0.78, indicating that the iron atom is defined as the somewhat distorted trigonal bipyramidal geometry. The silicon atom exhibited a distorted tetrahedral geometry; the sum of the bond angles around the Si1 atom (362.36°) was within the range of those of the corresponding silylene-ligated tetracarbonyliron(0) complexes [353.92–362.64°] [17,18,19]. Despite the strong σ-donor ability of 1, the Si1–Fe1 bond length [2.2675(6) Å] was slightly longer than that of the related amidinato–silylene–iron(0) complexes [2.237(7)–2.2777(5) Å] [17,18,19], probably due to the steric repulsion between the phenyl group on the phosphorus atom and the carbonyl ligand. Two Si–N bond lengths [1.8001(15) and 1.8078(15) Å] in the four-membered ring fell in the same range as those of the reported three-coordinated silylenes [1.8291(19)–1.8750(16) Å] [17,18,19]. The Si–Cl bond length in 2 was 2.1161(7) Å, which is slightly shorter than the corresponding bond length in the starting 1 [2.2291(11) Å] [21].
The structure of 2 was also confirmed following spectroscopic analyses. The 1H NMR spectrum of 2 in C6D6 displayed a singlet signal assigned to the tBu groups at 1.22 ppm. The aromatic protons were observed as broad signals at 7.07 and 8.03 ppm. In the 13C{1H} NMR spectrum of 2, the carbonyl ligands binding to the iron center were resonated at 217.0 ppm, which is comparable to those of the related amidinato–silylene–Fe(CO)4 complexes (213.1–217.2 ppm) [17,18,19,20]. In the 31P{1H} NMR of 2, a sharp singlet signal appeared at 46.8 ppm, which is quite similar to that of the starting 1 (46.6 ppm) [21]. The 29Si{1H} NMR spectrum of 2 showed a doublet signal via the 31P coupling of 11.9 Hz at 74.6 ppm, and shifts to a lower field compared to that of the starting 1 (59.8 ppm) [21] and falls in the range of those in the related amidinato–silylene–Fe(CO)4 complexes (40.3–112.8 ppm) [17,18,19,20]. The absorptions due to carbonyl stretching vibrations in the IR spectrum were observed at 2029, 1944, and 1913 cm1, consistent with the results of SCXRD analysis exhibiting local C3 V symmetry on the iron(0) site. All NMR (Figures S1–S4) and IR (Figure S5) spectra are in the Supplementary Materials.

3. Materials and Methods

3.1. General Considerations

Unless otherwise noted, all experiments were carried out under an argon atmosphere using standard Schlenk-line techniques or a UNICO glovebox. 1H, 13C, 29Si, and 31P NMR spectra were recorded on Bruker Avance-500 (500 MHz for 1H, 126 MHz for 13C, 99 MHz for 29Si, and 202 MHz for 31P) and Bruker Avance-400 (400 MHz for 1H, 101 MHz for 13C, and 162 MHz for 31P) spectrometers, with C6D6 as the solvent at room temperature. The IR spectrum was recorded on a TENSOR II. All melting points were determined on a Mel-Temp capillary tube apparatus and were uncorrected. Elemental analyses were carried out at the Molecular Analysis and Life Science Center of Saitama University. All solvents were dried over 4A molecular sieves or potassium mirror before use. All materials were obtained from commercial suppliers and were used without further purification, except 1 [21], which was prepared according to the corresponding literature procedure.

3.2. Synthesis of Complex 2

In a Schlenk tube, to a THF (2 mL) solution of chlorosilylene 1 (140 mg, 0.36 mmol), Fe(CO)5 (0.05 mL, 0.37 mmol) was added at ambient temperature. After stirring for 1 h, the reaction mixture was concentrated under reduced pressure, and the crude product was crystallized from a saturated THF solution at −10 °C to give 2 (138 mg, 69%) as yellow crystals. Mp. 141–142 °C (decomp.). 1H NMR (400 MHz, 25 °C C6D6): δ = 1.22 (s, 18H, CH3tBu), 7.07 (br, 6H, CHAr), 8.03 (br, 4H, CHAr). 13C{1H} (100 MHz, 25 °C, C6D6): δ = 32.0 (d, JCP = 5.6 Hz, CH3tBu), 55.3 (CtBu), 126.2 (d, JCP = 99.3 Hz, CAr), 129.1 (d, JCP = 11.3 Hz, CHAr), 129.1 (d, JCP = 11.0 Hz, CHAr), 134.3 (d, JCP = 2.9 Hz, CHAr), 134.5 (d, JCP = 2.9 Hz, CHAr), 134.7 (d, JCP = 12.1 Hz, CHAr), 134.8 (d, JCP = 12.2 Hz, CHAr), 217.0 (CO). 31P{1H} NMR (162 MHz, 25 °C, C6D6): δ = 46.6. 29Si{1H} NMR (99 MHz, 25 °C, C6D6): δ = 74.6 (d, JSiP = 11.9 Hz). IR(KBr): νCO = 2029, 1944, 1913 cm–1. Anal. Calculated for C24H28ClFeN2O4PSi: C, 51.58; H, 5.05; N, 5.01. Found: C, 51.84; H, 5.07; N, 4.90.

3.3. SCXRD Analysis of 2

A yellow single crystal of 2 was grown from a saturated THF solution at −10 °C. The intensity data were collected at 100 K on a Bruker SMART APEX II diffractometer employing graphite-monochromated MoKα radiation (λ = 0.71073 Å). The structure was solved using direct methods (SHELXT) [29] and refined with full-matrix least-squares procedures on F2 for all reflections (SHELXL) [30]. Hydrogen atoms were located by assuming the ideal geometry and were included in the structure calculations without any further refinement of the parameters. The structures of overlapped and disordered THF molecules were restrained to be identical to each other using DFIX (σ = 0.01), SAME (σ1 = 0.01, σ2 = 0.02,), and DELU (σ1 = 0.01, σ2 = 0.02) instructions.
Crystal data for C28H36ClFeN2O5PSi (2): M = 630.95 g mol−1, monoclinic, P21/c, a = 9.3227(7), b = 27.786(2), c = 11.5041(9) Å, β = 94.5030(10)°, V = 2970.9(4) Å3, Z = 4, Dx = 1.411 g cm−3, F(000) = 1320, and μ = 0.732 mm−1. CCDC deposition number: 2192007.

4. Conclusions

We demonstrated the synthesis and spectroscopic characterization of the novel silylene–tetracarbonyliron(0) complex {[Ph2P(tBuN)2](Cl)Si:->Fe(CO)4}. Further applications of N,N′-chelating silylenes are currently under investigation in our laboratory.

Supplementary Materials

The following are available online. All NMR (Figures S1–S4) and IR (Figure S5) spectra, and crystallographic data for 2 in Crystallographic Information File (CIF) format. CCDC 2192007 also contains the supplementary crystallographic data for this paper.

Author Contributions

Conceptualization, S.T. and N.N.; methodology, N.N.; formal analysis, S.T. and K.N.; investigation, S.T. and K.N.; resources, N.N.; data curation, N.N.; writing—original draft preparation, S.T. and K.N.; writing—review and editing, A.I. and N.N.; visualization, N.N.; supervision, N.N. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by JSPS KAKENHI (grant number: JP22K05138 to N.N.).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

CCDC 2192007 (2) contains the supplementary crystallographic data for this paper. These data can be obtained free of charge via www.ccdc.cam.ac.uk/data_request/cif (accessed on 31 July 2022), or by emailing data_request@ccdc.cam.ac.uk or contacting The Cambridge Crystallographic Data Centre, 12 Union Road, Cambridge CB2 1EZ, UK (fax: +44-1223-336033; e-mail: deposit@ccdc.cam.ac.uk).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Nagendran, S.; Roesky, H.W. The chemistry of aluminum(I), silicon(II), and germanium(II). Organometallics 2008, 27, 457–492. [Google Scholar] [CrossRef]
  2. Mizuhata, Y.; Sasamori, T.; Tokitoh, N. Stable heavier carbene analogues. Chem. Rev. 2009, 109, 3479–3511. [Google Scholar] [CrossRef] [PubMed]
  3. Krahfuss, M.J.; Radius, U. N-Heterocyclic silylenes as ambiphilic activators and ligands. Dalton Trans. 2021, 50, 6752–6765. [Google Scholar] [CrossRef]
  4. Blom, B.; Stoelzel, M.; Driess, M. New vistas in N-heterocyclic silylene (NHSi) transition-metal coordination chemistry: Syntheses, structures and reactivity towards activation of small molecules. Chem. Eur. J. 2013, 19, 40–62. [Google Scholar] [CrossRef] [PubMed]
  5. Raoufmoghaddam, S.; Zhou, Y.-P.; Wang, Y.; Driess, M. N-heterocyclic silylene as powerful steering ligands in catalyst. J. Organomet. Chem. 2017, 829, 2–10. [Google Scholar] [CrossRef]
  6. Zhou, Y.-P.; Driess, M. Isolable silylene ligands can boost efficiencies and selectivities in metal-mediated catalysis. Angew. Chem. Int. Ed. 2019, 58, 3715–3728. [Google Scholar] [CrossRef]
  7. Álvarez-Rodríguez, L.; Cabeza, J.A.; García-Álvarez, P.; Polo, D. The transition-metal chemistry of amidinatosilylenes, -germylenes and -stannylenes. Coord. Chem. Rev. 2015, 300, 1–28. [Google Scholar] [CrossRef]
  8. Yang, W.; Dong, Y.; Sun, H.; Li, X. Progress in the preparation and characterization of silylene iron, cobalt and nickel complexes. Dalton. Trans. 2021, 50, 6766–6772. [Google Scholar] [CrossRef]
  9. Wang, Y.; Koestenko, A.; Yao, S.; Driess, M. Divalent silicon-assisted activation of dihydrogen in a bis(N-heterocyclic silylene)xantene nickel(0) complex for efficient catalytic hydrogenation of olefins. J. Am. Chem. Soc. 2017, 139, 13499–13506. [Google Scholar] [CrossRef]
  10. Bai, Y.; Zhang, J.; Cui, C. An arene-tethered silylene ligand enabling reversible dinitrogen binding to iron and catalytic silylation. Chem. Commun. 2018, 54, 8124–8127. [Google Scholar] [CrossRef]
  11. Blom, B.; Enthaler, S.; Inoue, S.; Irran, E.; Driess, M. Electron-rich N-heterocyclic silylene (NHS)-iron complexes: Synthesis, structures, and catalytic ability of an isolable hydridosilylene-iron complex. J. Am. Chem. Soc. 2013, 135, 6703–6713. [Google Scholar] [CrossRef]
  12. Gallego, D.; Inoue, S.; Blom, B.; Driess, M. Highly electron-rich pincer-type iron complexes bearing innocent bis(metallylene)pyridine ligands: Syntheses, structures, and catalytic activity. Organometallics 2014, 33, 6885–6897. [Google Scholar] [CrossRef]
  13. Luecke, M.-P.; Porwal, D.; Kostenko, A.; Zhou, Y.-P.; Yao, S.; Keck, M.; Limberg, C.; Oestreich, M.; Driess, M. Bis(silylenyl)-substituted ferrocene-stabilized h6-arene iron(0) complexes: Synthesis, structure and catalytic application. Dalton. Trans. 2017, 46, 16412–16418. [Google Scholar] [CrossRef] [PubMed]
  14. He, Z.; Xue, X.; Liu, Y.; Yu, N.; Krogman, J.P. Aminolysis of bis[bis(trimethylsilyl)amido]-manganese, -iron, and -cobalt for the synthesis of mono- and bis- silylene complexes. Dalton. Trans. 2020, 49, 12586–12591. [Google Scholar] [CrossRef] [PubMed]
  15. Du, X.; Qi, X.; Li, K.; Li, X.; Sun, H.; Fuhr, O.; Fenske, D. (NHSi) iron (II) hydride for hydrosilylation of aldehydes and ketones. Appl. Organomet. Chem. 2021, 35, e6286. [Google Scholar] [CrossRef]
  16. Fan, Q.; Li, Q.; Qi, X.; Du, X.; Ren, S.; Li, X.; Fuhr, O.; Sun, H. Synthesis and structure of silylene iron complex. Z. Anorg. Allg. Chem. 2022, 648, e202200084. [Google Scholar] [CrossRef]
  17. Yang, W.; Fu, H.; Wang, H.; Chen, M.; Ding, Y.; Roesky, H.W.; Jana, A. A base-stabilized silylene with tricoordinate silicon atom as a ligand for a metal complex. Inorg. Chem. 2009, 48, 5058–5060. [Google Scholar] [CrossRef]
  18. Tacke, R.; Kobelt, C.; Baus, J.A.; Bertermann, R.; Burschka, C. Synthesis, structure and reactivity of a donor-stabilized silylene with a bulky bidentate benzamidinato ligand. Dalton Trans. 2015, 44, 14959–14974. [Google Scholar] [CrossRef]
  19. Breit, N.C.; Eisenhut, C.; Inoue, S. Phosphinosilylene as a novel ligand system for heterobimetallic complexes. Chem. Commun. 2016, 52, 5523–5526. [Google Scholar] [CrossRef]
  20. Blom, B.; Pohl, M.; Tan, G.; Gallego, D.; Driess, M. From unsymmetrically substituted benzamidinato and guanidinato dichlorohydridosilanes to novel hydrido N-heterocyclic silylene iron complexes. Organometallics 2014, 33, 5272–5282. [Google Scholar] [CrossRef]
  21. Takahashi, S.; Sekiguchi, J.; Ishii, A.; Nakata, N. An iminophosphonamido-chlorosilylene as a strong σ-donating NHSi ligand: Synthesis and coordination chemistry. Angew. Chem. Int. Ed. 2021, 133, 4101–4105. [Google Scholar] [CrossRef]
  22. Takahashi, S.; Ishii, A.; Nakata, N. Interconversion between a silaimine and an aminosilylene supported by an iminophosphonamide ligand. Chem. Commun. 2021, 57, 3203–3206. [Google Scholar] [CrossRef] [PubMed]
  23. Takahashi, S.; Ishii, A.; Nakata, N. Formation of silaimines from a sterically demanding iminophosphonamido chlorosilylene via intramolecular N–P bond cleavage. Chem. Commun. 2021, 57, 6728–6731. [Google Scholar] [CrossRef]
  24. Nakaya, K.; Takahashi, S.; Ishii, A.; Boonpalit, K.; Surawatanawong, P.; Nakata, N. Hydroboration of carbonyls and imines by an iminophosphonamido Tin(II) precatalyst. Dalton Trans. 2021, 50, 14810–14819. [Google Scholar] [CrossRef] [PubMed]
  25. Nakaya, K.; Ishii, A.; Nakata, N. Aluminum(III) di- and monochlorides incorporating an N,N’-chelating iminophosphonamide ligand: Synthesis and structures. Mendeleev Commun. 2022, 32, 71–73. [Google Scholar] [CrossRef]
  26. Takahashi, S.; Sekiguchi, J.; Nakaya, K.; Ishii, A.; Nakata, N. Halogen-exchange reactions of iminophosphonamido-chlorosilylenes with alkali halides: Convenient synthesis of heavier halosilylenes. Inorg. Chem. 2022, 61, 7266–7273. [Google Scholar] [CrossRef]
  27. Klein, A.; Neugebauer, M.; Krest, A.; Lüning, A.; Garbe, S.; Arefyeva, N.; Schlörer, N. Five coordinate platinum(II) in [Pt(Bpy)(Cod)(Me)][SbF6]: A structural and spectroscopic study. Inorganics 2015, 3, 118–138. [Google Scholar] [CrossRef]
  28. Burnett, M.N.; Johnson, C.K. ORTEPIII; Report ORNL-6895; Oak Ridge National Laboratory: Oak Ridge, TN, USA, 1996. [Google Scholar]
  29. Sheldrick, G.M. SHELXT—Integrated space-group and crystal-structure determination. Acta Crystallogr. Sect. A Found. Adv. 2015, A71, 3–8. [Google Scholar] [CrossRef]
  30. Sheldrick, G.M. Crystal structure refinement with SHELXL. Acta Crystallogr. Sect. C Struct. Chem. 2015, C71, 3–8. [Google Scholar] [CrossRef]
Scheme 1. Synthesis of silylene–tetracarbonyliron(0) complex 2.
Scheme 1. Synthesis of silylene–tetracarbonyliron(0) complex 2.
Molbank 2022 m1433 sch001
Figure 1. ORTEP [28] of 2 with thermal ellipsoids at a 50% probability. All hydrogen atoms and a molecule of THF in the unit cell were omitted for clarity.
Figure 1. ORTEP [28] of 2 with thermal ellipsoids at a 50% probability. All hydrogen atoms and a molecule of THF in the unit cell were omitted for clarity.
Molbank 2022 m1433 g001
Figure 2. The trigonal bipyramidal geometry with basal angles α and β, and angular structure parameter τ.
Figure 2. The trigonal bipyramidal geometry with basal angles α and β, and angular structure parameter τ.
Molbank 2022 m1433 g002
Table 1. Selected bond lengths [Å] and bond angles [°].
Table 1. Selected bond lengths [Å] and bond angles [°].
Bond Lengths[Å]Bond Angles[°]
Si1–Fe12.2675(6)Si1–Fe1–C21174.28(6)
Si1–Cl12.1161(7)Si1–Fe1–C2286.99(6)
Si1–N11.8001(15)Si1–Fe1–C2386.12(6)
Si1–N21.8078(15)Si1–Fe1–C2491.12(6)
Fe1–C211.791(2)C22–Fe1–C23127.39(9)
Fe1–C221.784(2)C23–Fe1–C24111.64(9)
Fe1–C231.774(2)C22–Fe1–C24120.59(9)
Fe1–C241.788(2)Cl1–Si1–N1104.37(6)
Cl1–Si1–N2104.82(5)
N1–Si1–N280.31(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

Takahashi, S.; Nakaya, K.; Ishii, A.; Nakata, N. [N,N′-Di-tert-butyl-P,P-diphenylphosphinimidic Amidato-κNN′]chlorosilicon-κSi-tetracarbonyliron. Molbank 2022, 2022, M1433. https://doi.org/10.3390/M1433

AMA Style

Takahashi S, Nakaya K, Ishii A, Nakata N. [N,N′-Di-tert-butyl-P,P-diphenylphosphinimidic Amidato-κNN′]chlorosilicon-κSi-tetracarbonyliron. Molbank. 2022; 2022(3):M1433. https://doi.org/10.3390/M1433

Chicago/Turabian Style

Takahashi, Shintaro, Kazuki Nakaya, Akihiko Ishii, and Norio Nakata. 2022. "[N,N′-Di-tert-butyl-P,P-diphenylphosphinimidic Amidato-κNN′]chlorosilicon-κSi-tetracarbonyliron" Molbank 2022, no. 3: M1433. https://doi.org/10.3390/M1433

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