Next Article in Journal
A Commonly Used Chinese Herbal Formula, Shu-Jing-Hwo-Shiee-Tang, Potentiates Anticoagulant Activity of Warfarin in a Rabbit Model
Previous Article in Journal
Characterization of Amide Bond Conformers for a Novel Heterocyclic Template of N-acylhydrazone Derivatives
Previous Article in Special Issue
Synthesis and Biological Activity of 6-Selenocaffeine: Potential Modulator of Chemotherapeutic Drugs in Breast Cancer Cells
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Modes of Neighbouring Group Participation by the Methyl Selenyl Substituent in β-Methylselenylmethyl-substituted 1-Phenylethyl Carbenium Ions

by
Benjamin L. Harris
and
Jonathan M. White
*
School of Chemistry and BIO-21 Institute, University of Melbourne, Parkville 3010, Victoria, Australia
*
Author to whom correspondence should be addressed.
Molecules 2013, 18(10), 11705-11711; https://doi.org/10.3390/molecules181011705
Submission received: 29 July 2013 / Revised: 7 September 2013 / Accepted: 9 September 2013 / Published: 25 September 2013
(This article belongs to the Special Issue Selenium and Tellurium Chemistry)

Abstract

:
Selenium substituents which are disposed β to an electron deficient centre, such as a carbocation p-orbital, or the π* orbital of an electron deficient p-system, interact in a stabilising way by a combination of C-Se hyperconjugation (σSe-C–π* interaction), and a through-space homoconjugative nSe–π* interaction. The relative importance of these two modes of interaction is dependant on the electron demand of the cation, with hyperconjugation predominating for low electron demand systems, and the nSe–π* interaction predominating for high electron demand cations.

1. Introduction

Unimolecular solvolyses of the conformationally biased β-phenylselenyl trifluoroacetate 1 (Figure 1) occurs at a rate which is 107 times faster than the corresponding unsubstituted derivative 2 (Figure 1) suggesting that the selenium substituent provides strong assistance in the departure of the trifluoroacetate leaving group [1]. The mechanism of participation by the selenium substituent might reasonably be described by conventional neighbouring group participation [2]. In this case the selenium lone pair electrons act as an internal nucleophile, displacing the leaving group to give the seleniranium ion intermediate 3 (Figure 1). This is an example of non-vertical participation [3].
Figure 1. Modes of participation by β-Selenium substituents.
Figure 1. Modes of participation by β-Selenium substituents.
Molecules 18 11705 g001
However, consideration was given to the possibility, that participation by the selenium substituent might occur by σC-Se-p hyperconjugation (vertical participation), and involve the open carbenium ion 4 (Figure 1) as an intermediate [3,4,5,6]. This mode of participation is analogous to that provided by the trimethylsilyl substituent in the carbenium ion 5 (Figure 1) which is the basis of the silicon β-effect [7,8,9,10]. Application of the variable oxygen probe to ether and ester derivatives of the antiperiplanar β-phenylselenyl alcohol 6 (Figure 1) provided crystallographic evidence that the C-Se bond is a strong σ-donor and can therefore effectively stabilise a neighbouring carbenium ion by hyperconjugation alone [1]. More recently NMR, crystallographic and computational studies on phenylselenylmethyl-substituted pyridinium ions 7 and 8 (Figure 2) revealed that a number of orbital interactions involving the selenium substituent were responsible for stabilisation of the charge on the adjacent carbon (Figure 3) [11].
Figure 2. Selenium substituted pyridinium ions.
Figure 2. Selenium substituted pyridinium ions.
Molecules 18 11705 g002
Figure 3. Orbital interactions involving the selenium substituent in 2- and 4-substituted pyridinium ions.
Figure 3. Orbital interactions involving the selenium substituent in 2- and 4-substituted pyridinium ions.
Molecules 18 11705 g003
The orbital interactions include σSe-C-π* hyperconjugation (Figure 3A), an anomeric effect (Figure 3B) and a through-space interaction between the selenium p-type lone pair orbital and the electron deficient pyridinium ion π system (Figure 3C) which represents the early stages of the bridging interaction as represented by structure 3 above. The latter two interactions explain the preferred gauche dihedral angle about the Se-C bond in these structures, a conformation, which is also preferred in α-phenylselenyl ketones, where similar orbital interactions are plausible [12,13].
Calculations showed that C-Se hyperconjugation (σC-Se–π*) is the predominant mode of stabilisation in the weakly electron demanding pyridinium ions 7 and 8, where the σC-Se–π* hyperconjugative interaction provides 34.8 and 34.2 kJ mol−1 stabilisation respectively, while the through-space nSe–π* interaction provides 9.0 and 8.0 kJ mol−1 of stabilisation. However the through-space (nSe–π*) interaction becomes more important as the electron demand of the β-cation increases. For example, in the selenylmethyl-substituted cyclopropenium ion 9 the NBO interaction energies for the σC-Se–π* interaction is 104.9 kJ mol−1 while the nSe–π* through-space interaction is 73.7 kJ mol-1. Also consistent with the increasing importance of the through space interaction as the electron demand increases is the closing of the Se-C-C(+) bond angle, which decreases from 110.9° for the pyridinium ion 8 to 101.1° in the cyclopropenium ion 9. The anomeric interaction (nSe–σ*CC) was found to be relatively unimportant in all ions. In this paper we investigate computationally the relative importance of the stabilising orbital interactions in the more highly electron demanding ions 1013 (Figure 4).
Figure 4. β-Selenium-substituted ions with higher electron demand.
Figure 4. β-Selenium-substituted ions with higher electron demand.
Molecules 18 11705 g004

2. Methods

Calculations were performed at the B3LYP/6-311++G** level of theory [14,15,16,17,18], a level of theory which has been previously employed to investigate stereoelectronic effects of chalcogen substituents [11,19]. Natural Bond Orbitals (NBOs) were calculated using the NBO 3.1 program [20] as implemented in the Gaussian 03 package [21].

3. Results and Discussion

The parent benzylically-stabilised β-selenium substituted 1-phenylethyl cation 12 has two low energy conformations, both of which allow vertical and non-vertical modes of participation to occur. In both conformations the C-Se bond is aligned with the direction of the carbenium ion p-orbital, allowing for σC-Se–π* hyperconjugation to occur effectively, in addition the CH3-Se-CH2-C(+) dihedral angle is close to orthogonal, which allows the through-space nSe–π* interaction between the selenium p-type lone pair orbital and the carbenium ion p-orbital to occur. This gives rise to the exo conformation 12a and the endo conformation 12b these conformations are very similar energetically, with the exo conformer being slightly favoured (1.1 kJ mol−1) (Figure 5). For practical purposes the comparisons made below apply to the exo conformer.
Figure 5. Low energy conformations of the seleniranium ion 12.
Figure 5. Low energy conformations of the seleniranium ion 12.
Molecules 18 11705 g005
The computed structures of the β-selenyl-carbenium ions 1013 are presented in Figure 6, while selected geometrical parameters and NBO orbital interaction energies are presented in Table 1 [20]. A convenient measure of electron demand of a cation is the pKR+ value, those, which are available from the literature have been included in this table.
Figure 6. Calculated structures for the β-selenium substituted 1-phenylethyl cations 1013.
Figure 6. Calculated structures for the β-selenium substituted 1-phenylethyl cations 1013.
Molecules 18 11705 g006
Decreasing stabilisation of the carbenium ions by delocalization into the aromatic ring is demonstrated by the C(Ar)-C+ distance which increases from 1.389 Å in 10 to 1.452 Å in 13 where there is little resonance interaction. The general trend apparent from Table 1 is that as the magnitude of both σC-Se–π hyperconjugation and the through-space nSe–p interaction increases with increasing electron demand of the 1-phenylethyl cation. Increasing strength of σC-Se–π hyperconjugation is evident from the decreasing population of the σC-Se orbital with increasing electron demand, and the increasing magnitude of the orbital overlap term [F(i,j)], while the increasing strength of the through-space nSe-p interaction is evident from the decreasing population of the nSe p-type lone pair orbital with increasing electron demand, and an increasing orbital overlap term [F(I,j)]. However the relative importance of the through-space stabilising interaction increases with increasing electron demand. For example in the relatively stable 4-amino-1-phenylethyl cation 10 σC-Se–π hyperconjugation is the most important stabilising interaction (55.7 kJ mol−1 vs. 17.2 kJ mol−1 involving the selenium substituent, however while this stabilising interaction increases with increasing electron demand, the nSe–p through space interaction increases more profoundly, and in the parent cation 12 the through space through-space nSe–p interaction is the most important stabilising interaction (418.4 vs. 211.6 kJ mol−1). The increasing importance of the through-space interaction is consistent with the steady closing of the Se-CH2-C(+) bond angle from 1012, while in the most electron deficient cation, the p-nitrophenylethyl cation the ion is bridged, and the individual contributions from hyperconjugation and the through-space interaction can no-longer be deconvoluted.
Table 1. Structural and orbital properties, and NBO interaction energies of β-selenium substituted 1-phenylethyl cations 1013.
Table 1. Structural and orbital properties, and NBO interaction energies of β-selenium substituted 1-phenylethyl cations 1013.
10111213
Se-CH2 (Å)2.0122.0112.0102.010
Se-CH2-C+ (°)98.8493.4184.7878.96
Se…C(+) Å2.6642.5542.3722.245
C(Ar)-C+1.3891.4031.4231.452
pKR+ a −12.4 [ 22]<−20 [ 23]
Vertical interaction E(2) (kJ mol−1)55.773.3211.6-
σC-Seenergy (a.u.)−0.641−0.653−0.660
σC-Sepopulation1.8911.8661.820
Overlap, F(i,j) (a.u.)0.0760.0870.139
Nonvertical interaction E(2) (kJ mol−1)17.249.0418.4-
nSeenergy (a.u.)−0.363−0.379−0.397
nSepopulation1.7991.7271.606
F(i,j) (a.u.)0.0370.0460.083
a pKR+ values for the corresponding non-substituted carbenium ions.

4. Conclusions

Selenium substituents interact with electron deficient orbitals at the β-position by a combination of C-Se hyperconjugation, in which the electrons in the σC-Se bonding orbital mix with the electron deficient orbital, and a through space interaction between the selenium p-type lone pair orbital and the electron deficient centre, this latter interaction is also referred to as homo-conjugation. For cations with low electron demand, there is very little distortion of the Se-C-C(+) bond angle, and the most important mode of stabilisation is by σC-Se–π hyperconjugation. However as the electron demand of the cation increases, then closing of the Se-C-C(+) bond angle occurs, this increases the orbital overlap between the selenium p-type lone pair orbital and the cabocation p-orbital and this becomes the predominant mode of stabilisation.

Supplementary Materials

Supplementary materials can be accessed at: https://www.mdpi.com/1420-3049/18/10/11705/s1.

Acknowledgments

We thank the Australian Research Council for financial support for financial support (DP0770565) and an award of an APA to B.H. We would also like to thank the Victorian Partnership for Advanced Computing and the Victorian Institute for Chemical Sciences High Performance Computing Facility for the computational time.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. White, J.M.; Lambert, J.B.; Spiniello, M.; Jones, S.A.; Gable, R.W. Vertical and Nonvertival Participation by Sulfur, Selenium, and Tellurium. Chem. Eur. J. 2002, 8, 2799–2811. [Google Scholar] [CrossRef]
  2. Capon, B.; McManus, S.P. Neighbouring Group Participation; Plenum Press: New York, NY, USA, 1976; Volume 1. [Google Scholar]
  3. Traylor, T.G.; Berwin, H.J.; Jerkunica, M.L. σ–π Conjugation: Occurrence and magnitude. Pure Appl. Chem. 1972, 30, 599–606. [Google Scholar]
  4. Hanstein, W.; Berwin, H.J.; Traylor, T.G. Modes of Carbonium Ion Stabilization. Evidence from Charge-Transfer Spectra. J. Am. Chem. Soc. 1970, 92, 829–835. [Google Scholar] [CrossRef]
  5. Hanstein, W.; Berwin, H.J.; Traylor, T.G. σ–π onjugation of Carbon-Metal Bonds. Stereoelectronic and Inductive Effects. J. Am. Chem. Soc. 1970, 92, 7476–7477. [Google Scholar]
  6. Traylor, T.G.; Hanstein, W.; Berwin, H.J.; Clinton, N.A.; Brown, R.S. Vertical Stabilization of Cations by Neighbouring Sigma-Bonds- General Considerations. J. Am. Chem. Soc. 1971, 93, 5715–5725. [Google Scholar] [CrossRef]
  7. Lambert, J.B. The Interaction of Silicon with Positively Charged Carbon. Tetrahedron 1990, 46, 2677–2899. [Google Scholar] [CrossRef]
  8. Lambert, J.B.; Zhou, Y.; Emblidge, R.W.; Salvador, L.A.; Liu, X.Y.; So, J.H.; Chelius, E.C. The β effect of silicon and related manifestations of σ conjugation. Acc. Chem. Res. 1999, 32, 183–190. [Google Scholar] [CrossRef]
  9. White, J.M.; Clark, C. Stereoelectronic Effects of Group 4 Metal substituents in Organic Chemistry. In Topics in Stereochemistry; Denmark, S., Ed.; John Wiley and Sons: New York, NY, USA, 1999; Volume 22, Chapter 3. [Google Scholar]
  10. White, J.M. Reactivity and Ground State Effects of Silicon. Aust. J. Chem. 1995, 48, 1227–1251. [Google Scholar] [CrossRef]
  11. Lim, S.F.; Harris, B.L.; Blanc, P.; White, J.M. Orbital interactions in selenylmethyl substituted pyridinium ions and carbenium ions with higher electron demand. J. Org. Chem. 2011, 76, 1673–1682. [Google Scholar] [CrossRef]
  12. McLeod, R.G.; Johnson, B.D.; Pinto, B.M. A Generalized exo-Anomeric Effect. Substituent and Solvent Effects on the Conformational Equilibria of 2-(Arylseleno)cyclohexanones. Israel. J. Chem. 2000, 40, 307–316. [Google Scholar] [CrossRef]
  13. Szabo, K.J.; Frisell, H.; Engman, L.; Piatek, M.; Oleksyn, B.; Sliwinski, J. α-(Phenylselenyl) ketones—Structure, molecular modeling and rationalization of their glutathione peroxidase-like activity. J. Molec. Struct. 1998, 448, 21–28. [Google Scholar] [CrossRef]
  14. Becke, A.D. Density Functional Thermochemistry 3. The Role of Exact Exchange. J. Chem. Phys. 1993, 98, 5648–5652. [Google Scholar] [CrossRef]
  15. McLean, A.D.; Chandler, G.S. Contracted Gaussian-Basis Sets for Molecular Calculations. J. Chem. Phys. 1980, 72, 5639–5648. [Google Scholar] [CrossRef]
  16. Krishnan, R.; Schlegel, H.B.; Pople, J.A. Derivative Studies in Configuration-Interaction. J. Chem. Phys. 1980, 72, 4654–4655. [Google Scholar] [CrossRef]
  17. 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. Comp. Chem. 1983, 4, 294–301. [Google Scholar] [CrossRef]
  18. 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]
  19. Alabugin, I.V.; Manoharan, M. Zeidan, Stereoelectronic effects and general trends in hyperconjugative acceptor ability of σ bonds, T.A. J. Am. Chem. Soc. 2002, 124, 3175–3185. [Google Scholar] [CrossRef]
  20. Glendening, E.D.; Reed, A.E.; Carpenter, J.E.; Weinhold, F. NBO; Version 3.1; Theoretical Chemistry Institute, University of Wisconsin: Madison, WI, USA, 1990. [Google Scholar]
  21. 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 C.02; Gaussian, Inc.: Wallingford, CT, USA, 2004. [Google Scholar]
  22. Toteva, M.M.; Moran, M.; Amyes, T.L.; Richard, J.P. Substituent Effects of Carbocation Stability: The pK(R) for p-quinonemethide. J. Am. Chem. Soc. 2003, 125, 8814–8819. [Google Scholar] [CrossRef]
  23. Amyes, T.L.; Richard, J.P.; Novak, M. Experiments and Calculations for Determinations of the Stabilities of Benzyl, Benzhydryl, and Fluorenyl Carbocations—Antiaromaticity Revisited. J. Am. Chem. Soc. 1992, 114, 8032–8041. [Google Scholar] [CrossRef]
  • Sample Availability: Not available.

Share and Cite

MDPI and ACS Style

Harris, B.L.; White, J.M. Modes of Neighbouring Group Participation by the Methyl Selenyl Substituent in β-Methylselenylmethyl-substituted 1-Phenylethyl Carbenium Ions. Molecules 2013, 18, 11705-11711. https://doi.org/10.3390/molecules181011705

AMA Style

Harris BL, White JM. Modes of Neighbouring Group Participation by the Methyl Selenyl Substituent in β-Methylselenylmethyl-substituted 1-Phenylethyl Carbenium Ions. Molecules. 2013; 18(10):11705-11711. https://doi.org/10.3390/molecules181011705

Chicago/Turabian Style

Harris, Benjamin L., and Jonathan M. White. 2013. "Modes of Neighbouring Group Participation by the Methyl Selenyl Substituent in β-Methylselenylmethyl-substituted 1-Phenylethyl Carbenium Ions" Molecules 18, no. 10: 11705-11711. https://doi.org/10.3390/molecules181011705

Article Metrics

Back to TopTop