fbpx
Wikipedia

Cation–π interaction

Cation–π interaction is a noncovalent molecular interaction between the face of an electron-rich π system (e.g. benzene, ethylene, acetylene) and an adjacent cation (e.g. Li+, Na+). This interaction is an example of noncovalent bonding between a monopole (cation) and a quadrupole (π system). Bonding energies are significant, with solution-phase values falling within the same order of magnitude as hydrogen bonds and salt bridges. Similar to these other non-covalent bonds, cation–π interactions play an important role in nature, particularly in protein structure, molecular recognition and enzyme catalysis. The effect has also been observed and put to use in synthetic systems.[1][2]

Cation–π interaction between benzene and a sodium cation.
The π system above and below the benzene ring leads to a quadrupole charge distribution.

Origin of the effect edit

Benzene, the model π system, has no permanent dipole moment, as the contributions of the weakly polar carbon–hydrogen bonds cancel due to molecular symmetry. However, the electron-rich π system above and below the benzene ring hosts a partial negative charge. A counterbalancing positive charge is associated with the plane of the benzene atoms, resulting in an electric quadrupole (a pair of dipoles, aligned like a parallelogram so there is no net molecular dipole moment). The negatively charged region of the quadrupole can then interact favorably with positively charged species; a particularly strong effect is observed with cations of high charge density.[2]

Nature of the cation–π interaction edit

The most studied cation–π interactions involve binding between an aromatic π system and an alkali metal or nitrogenous cation. The optimal interaction geometry places the cation in van der Waals contact with the aromatic ring, centered on top of the π face along the 6-fold axis.[3] Studies have shown that electrostatics dominate interactions in simple systems, and relative binding energies correlate well with electrostatic potential energy.[4][5]

The Electrostatic Model developed by Dougherty and coworkers describes trends in binding energy based on differences in electrostatic attraction. It was found that interaction energies of cation–π pairs correlate well with electrostatic potential above the π face of arenes: for eleven Na+-aromatic adducts, the variation in binding energy between the different adducts could be completely rationalized by electrostatic differences. Practically, this allows trends to be predicted qualitatively based on visual representations of electrostatic potential maps for a series of arenes. Electrostatic attraction is not the only component of cation–π bonding. For example, 1,3,5-trifluorobenzene interacts with cations despite having a negligible quadrupole moment. While non-electrostatic forces are present, these components remain similar over a wide variety of arenes, making the electrostatic model a useful tool in predicting relative binding energies. The other "effects" contributing to binding are not well understood. Polarization, donor-acceptor[permanent dead link] and charge-transfer interactions have been implicated; however, energetic trends do not track well with the ability of arenes and cations to take advantage of these effects. For example, if induced dipole was a controlling effect, aliphatic compounds such as cyclohexane should be good cation–π partners (but are not).[4]

The cation–π interaction is noncovalent and is therefore fundamentally different than bonding between transition metals and π systems. Transition metals have the ability to share electron density with π-systems through d-orbitals, creating bonds that are highly covalent in character and cannot be modeled as a cation–π interaction.

Factors influencing the cation–π bond strength edit

Several criteria influence the strength of the bonding: the nature of the cation, solvation effects, the nature of the π system, and the geometry of the interaction.

Nature of the cation edit

From electrostatics (Coulomb's law), smaller and more positively charged cations lead to larger electrostatic attraction. Since cation–π interactions are predicted by electrostatics, it follows that cations with larger charge density interact more strongly with π systems.

The following table shows a series of Gibbs free energy of binding between benzene and several cations in the gas phase.[2][6] For a singly charged species, the gas-phase interaction energy correlates with the ionic radius,   (non-spherical ionic radii are approximate).[7][8]

M+ Li+ Na+ K+ NH4+ Rb+ NMe4+
–ΔG [kcal/mol] 38 27 19 19 16 9
  [Å] 0.76 1.02 1.38 1.43 1.52 2.45

This trend supports the idea that coulombic forces play a central role in interaction strength, since for other types of bonding one would expect the larger and more polarizable ions to have greater binding energies.

Solvation effects edit

The nature of the solvent also determines the absolute and relative strength of the bonding. Most data on cation–π interaction is acquired in the gas phase, as the attraction is most pronounced in that case. Any intermediating solvent molecule will attenuate the effect, because the energy gained by the cation–π interaction is partially offset by the loss of solvation energy.

For a given cation–π adduct, the interaction energy decreases with increasing solvent polarity. This can be seen by the following calculated interaction energies of methylammonium and benzene in a variety of solvents.[9]

 
Calculated interaction energies of methylamonium and benzene in a variety of solvents

Additionally, the trade-off between solvation and the cation–π effect results in a rearrangement of the order of interaction strength for a series of cations. While in the gas phase the most densely charged cations have the strongest cation–π interaction, these ions also have a high desolvation penalty. This is demonstrated by the relative cation–π bond strengths in water for alkali metals:[10]

 

Nature of the π system edit

Quadrupole moment edit

Comparing the quadrupole moment of different arenes is a useful qualitative tool to predict trends in cation–π binding, since it roughly correlates with interaction strength. Arenes with larger quadrupole moments are generally better at binding cations.

However, a quadrupole-ion model system cannot be used to quantitatively model cation–π interactions. Such models assume point charges, and are therefore not valid given the short cation–π bond distance. In order to use electrostatics to predict energies, the full electrostatic potential surface must be considered, rather than just the quadrupole moment as a point charge.[2]

Substituents on the aromatic ring edit

 
Binding energy (in kcal/mol) for Na+ to benzene with prototypical substituents.[4]

The electronic properties of the substituents also influence the strength of the attraction.[11] Electron withdrawing groups (for example, cyano −CN) weaken the interaction, while electron donating substituents (for example, amino −NH2) strengthen the cation–π binding. This relationship is illustrated quantitatively in the margin for several substituents.

The electronic trends in cation–π binding energy are not quite analogous to trends in aryl reactivity. Indeed, the effect of resonance participation by a substituent does not contribute substantively to cation–π binding, despite being very important in many chemical reactions with arenes. This was shown by the observation that cation–π interaction strength for a variety of substituted arenes correlates with the σmeta Hammett parameter. This parameter is meant to illustrate the inductive effects of functional groups on an aryl ring.[4]

The origin of substituent effects in cation–π interactions has often been attributed to polarization from electron donation or withdrawal into or out of the π system.[12] This explanation makes intuitive sense, but subsequent studies have indicated that it is flawed. Recent computational work by Wheeler and Houk strongly indicate that the effect is primarily due to direct through-space interaction between the cation and the substituent dipole. In this study, calculations that modeled unsubstituted benzene plus interaction with a molecule of "H-X" situated where a substituent would be (corrected for extra hydrogen atoms) accounted for almost all of the cation–π binding trend. For very strong pi donors or acceptors, this model was not quite able account for the whole interaction; in these cases polarization may be a more significant factor.[5]

Binding with heteroaromatic systems edit

Heterocycles are often activated towards cation–π binding when the lone pair on the heteroatom is in incorporated into the aromatic system (e.g. indole, pyrrole). Conversely, when the lone pair does not contribute to aromaticity (e.g. pyridine), the electronegativity of the heteroatom wins out and weakens the cation–π binding ability.

Since several classically "electron rich" heterocycles are poor donors when it comes to cation–π binding, one cannot predict cation–π trends based on heterocycle reactivity trends. Fortunately, the aforementioned subtleties are manifested in the electrostatic potential surfaces of relevant heterocycles.[2]

 

cation–heterocycle interaction is not always a cation–π interaction; in some cases it is more favorable for the ion to be bound directly to a lone pair. For example, this is thought to be the case in pyridine-Na+ complexes.

Geometry edit

cation–π interactions have an approximate distance dependence of 1/rn where n<2. The interaction is less sensitive to distance than a simple ion-quadrupole interaction which has 1/r3 dependence.[13]

A study by Sherrill and coworkers probed the geometry of the interaction further, confirming that cation–π interactions are strongest when the cation is situated perpendicular to the plane of atoms (θ = 0 degrees in the image below). Variations from this geometry still exhibit a significant interaction which weakens as θ angle approaches 90 degrees. For off-axis interactions the preferred ϕ places the cation between two H atoms. Equilibrium bond distances also increase with off-axis angle. Energies where the cation is coplanar with the carbon ring are saddle points on the potential energy surface, which is consistent with the idea that interaction between a cation and the positive region of the quadrupole is not ideal.[14]

 

Relative interaction strength edit

In aqueous media, the cation–π interaction is comparable to (and potentially stronger than) ammonium-carboxylate salt bridges. Computed values below show that as solvent polarity increases, the strength of the cation–π complex decreases less dramatically. This trend can be rationalized by desolvation effects: salt bridge formation has a high desolvation penalty for both charged species whereas the cation–π complex would only pay a significant penalty for the cation.[9]

 

In nature edit

Nature's building blocks contain aromatic moieties in high abundance. Recently, it has become clear that many structural features that were once thought to be purely hydrophobic in nature are in fact engaging in cation–π interactions. The amino acid side chains of phenylalanine, tryptophan, tyrosine, histidine, are capable of binding to cationic species such as charged amino acid side chains, metal ions, small-molecule neurotransmitters and pharmaceutical agents. In fact, macromolecular binding sites that were hypothesized to include anionic groups (based on affinity for cations) have been found to consist of aromatic residues instead in multiple cases. Cation–π interactions can tune the pKa of nitrogenous side-chains, increasing the abundance of the protonated form; this has implications for protein structure and function.[15] While less studied in this context, the DNA bases are also able to participate in cation–π interactions.[16][17]

Role in protein structure edit

Early evidence that cation–π interactions played a role in protein structure was the observation that in crystallographic data, aromatic side chains appear in close contact with nitrogen-containing side chains (which can exist as protonated, cationic species) with disproportionate frequency.

A study published in 1986 by Burley and Petsko looked at a diverse set of proteins and found that ~ 50% of aromatic residues Phe, Tyr, and Trp were within 6Å of amino groups. Furthermore, approximately 25% of nitrogen containing side chains Lys, Asn, Gln, and His were within van der Waals contact with aromatics and 50% of Arg in contact with multiple aromatic residues (2 on average).[18]

Studies on larger data sets found similar trends, including some dramatic arrays of alternating stacks of cationic and aromatic side chains. In some cases the N-H hydrogens were aligned toward aromatic residues, and in others the cationic moiety was stacked above the π system. A particularly strong trend was found for close contacts between Arg and Trp. The guanidinium moiety of Arg in particular has a high propensity to be stacked on top of aromatic residues while also hydrogen-bonding with nearby oxygen atoms.[19][20][21]

Molecular recognition and signaling edit

 
Cationic Acetylcholine binding to a tryptophan residue of the nicotinamide acetylcholine receptor via a cation–π effect.

An example of cation–π interactions in molecular recognition is seen in the nicotinic acetylcholine receptor (nAChR) which binds its endogenous ligand, acetylcholine (a positively charged molecule), via a cation–π interaction to the quaternary ammonium. The nAChR neuroreceptor is a well-studied ligand-gated ion channel that opens upon acetylcholine binding. Acetylcholine receptors are therapeutic targets for a large host of neurological disorders, including Parkinson's disease, Alzheimer's disease, schizophrenia, depression and autism. Studies by Dougherty and coworkers confirmed that cation–π interactions are important for binding and activating nAChR by making specific structural variations to a key tryptophan residue and correlating activity results with cation–π binding ability.[22]

The nAChR is especially important in binding nicotine in the brain, and plays a key role in nicotine addiction. Nicotine has a similar pharmacophore to acetylcholine, especially when protonated. Strong evidence supports cation–π interactions being central to the ability of nicotine to selectively activate brain receptors without affecting muscle activity.[23][24]

 

A further example is seen in the plant UV-B sensing protein UVR8. Several tryptophan residues interact via cation–π interactions with arginine residues which in turn form salt bridges with acidic residues on a second copy of the protein. It has been proposed[25] that absorption of a photon by the tryptophan residues disrupts this interaction and leads to dissociation of the protein dimer.

Cation–π binding is also thought to be important in cell-surface recognition[2][26]

Enzyme catalysis edit

Cation–π interactions can catalyze chemical reactions by stabilizing buildup of positive charge in transition states. This kind of effect is observed in enzymatic systems. For example, acetylcholine esterase contains important aromatic groups that bind quaternary ammonium in its active site.[2]

Polycyclization enzymes also rely on cation–π interactions. Since proton-triggered polycyclizations of squalene proceed through a (potentially concerted) cationic cascade, cation–π interactions are ideal for stabilizing this dispersed positive charge. The crystal structure of squalene-hopene cyclase shows that the active site is lined with aromatic residues.[27]

 
Cyclization of squalene to form hopene

In synthetic systems edit

Solid state structures edit

Cation–π interactions have been observed in the crystals of synthetic molecules as well. For example, Aoki and coworkers compared the solid state structures of Indole-3-acetic acid choline ester and an uncharged analogue. In the charged species, an intramolecular cation–π interaction with the indole is observed, as well as an interaction with the indole moiety of the neighboring molecule in the lattice. In the crystal of the isosteric neutral compound the same folding is not observed and there are no interactions between the tert-butyl group and neighboring indoles.[28]

 
Cation–π interaction in indole-3-acetic acid choline ester compared to neutral analog

Supramolecular receptors edit

 
Cyclophane host–guest complex

Some of the first studies on the cation–π interaction involved looking at the interactions of charged, nitrogenous molecules in cyclophane host–guest chemistry. It was found that even when anionic solubilizing groups were appended to aromatic host capsules, cationic guests preferred to associate with the π-system in many cases. The type of host shown to the right was also able to catalyze N-alkylation reactions to form cationic products.[29]

More recently, cation–π centered substrate binding and catalysis has been implicated in supramolecular metal-ligand cluster catalyst systems developed by Raymond and Bergman.[30]

Use of π-π, CH-π, and π-cation interactions in supramolecular assembly edit

π-systems are important building blocks in supramolecular assembly because of their versatile noncovalent interactions with various functional groups. Particularly, π-π, CH-π, and π-cation interactions are widely used in supramolecular assembly and recognition.

 
Fig. 1: Examples of π-π. CH-π, and π-cation interactions

π-π interaction concerns the direct interactions between two &pi-systems; and cation–π interaction arises from the electrostatic interaction of a cation with the face of the π-system. Unlike these two interactions, the CH-π interaction arises mainly from charge transfer between the C-H orbital and the π-system.

A notable example of applying π-π interactions in supramolecular assembly is the synthesis of catenane. The major challenge for the synthesis of catenane is to interlock molecules in a controlled fashion. Stoddart and co-workers developed a series of systems utilizing the strong π-π interactions between electron-rich benzene derivatives and electron-poor pyridinium rings.[31] [2]Catanene was synthesized by reacting bis(pyridinium) (A), bisparaphenylene-34-crown-10 (B), and 1, 4-bis(bromomethyl)benzene C (Fig. 2). The π-π interaction between A and B directed the formation of an interlocked template intermediate that was further cyclized by substitution reaction with compound C to generate the [2]catenane product.

 
Fig. 2: The Stoddart synthesis of [2]catenane

Organic synthesis and catalysis edit

Cation–π interactions have likely been important, though unnoticed, in a multitude of organic reactions historically. Recently, however, attention has been drawn to potential applications in catalyst design. In particular, noncovalent organocatalysts have been found to sometimes exhibit reactivity and selectivity trends that correlate with cation–π binding properties. A polycyclization developed by Jacobsen and coworkers shows a particularly strong cation–π effect using the catalyst shown below.[32]

 

Anion–π interaction edit

 
Quadrupole moments of benzene and hexafluorobenzene. The polarity is inverted due to differences in electronegativity for hydrogen and fluorine relative to carbon; the inverted quadrupole moment of hexafluorobenzene is necessary for anion-pi interactions.

In many respects, anion–π interaction is the opposite of cation–π interaction, although the underlying principles are identical. Significantly fewer examples are known to date. In order to attract a negative charge, the charge distribution of the π system has to be reversed. This is achieved by placing several strong electron withdrawing substituents along the π system (e. g. hexafluorobenzene).[33] The anion–π effect is advantageously exploited in chemical sensors for specific anions.[34]

See also edit

References edit

  1. ^ Eric V. Anslyn; Dennis A. Dougherty (2004). Modern Physical Organic Chemistry. University Science Books. ISBN 978-1-891389-31-3.
  2. ^ a b c d e f g Dougherty, D. A.; J.C. Ma (1997). "The Cation–π Interaction". Chemical Reviews. 97 (5): 1303–1324. doi:10.1021/cr9603744. PMID 11851453.
  3. ^ Tsuzuki, Seiji; Yoshida, Masaru; Uchimaru, Tadafumi; Mikami, Masuhiro (2001). "The Origin of the Cation/π Interaction: The Significant Importance of the Induction in Li+and Na+Complexes". The Journal of Physical Chemistry A. 105 (4): 769–773. Bibcode:2001JPCA..105..769T. doi:10.1021/jp003287v.
  4. ^ a b c d S. Mecozzi; A. P. West; D. A. Dougherty (1996). "Cation–π Interactions in Simple Aromatics: Electrostatics Provide a Predictive Tool". Journal of the American Chemical Society. 118 (9): 2307–2308. doi:10.1021/ja9539608.
  5. ^ a b S. E. Wheeler; K. N. Houk (2009). "Substituent Effects in Cation/π Interactions and Electrostatic Potentials above the Center of Substituted Benzenes Are Due Primarily to through-Space Effects of the Substituents". J. Am. Chem. Soc. 131 (9): 3126–7. doi:10.1021/ja809097r. PMC 2787874. PMID 19219986.
  6. ^ J. C. Amicangelo; P. B. Armentrout (2000). "Absolute Binding Energies of Alkali-Metal Cation Complexes with Benzene Determined by Threshold Collision-Induced Dissociation Experiments and ab Initio Theory". J. Phys. Chem. A. 104 (48): 11420–11432. Bibcode:2000JPCA..10411420A. doi:10.1021/jp002652f.
  7. ^ Robinson RA, Stokes RH. Electrolyte solutions. UK: Butterworth Publications, Pitman, 1959.
  8. ^ Clays and Clay Minerals, Vol.45, No. 6, 859-866, 1997.
  9. ^ a b Gallivan, Justin P.; Dougherty, Dennis A. (2000). "A Computational Study of Cation−π Interactions vs Salt Bridges in Aqueous Media: Implications for Protein Engineering" (PDF). Journal of the American Chemical Society. 122 (5): 870–874. doi:10.1021/ja991755c.
  10. ^ Kumpf, R.; Dougherty, D. (1993). "A mechanism for ion selectivity in potassium channels: Computational studies of cation–pi interactions". Science. 261 (5129): 1708–10. Bibcode:1993Sci...261.1708K. doi:10.1126/science.8378771. PMID 8378771.
  11. ^ Raju, Rajesh K.; Bloom, Jacob W. G.; An, Yi; Wheeler, Steven E. (2011). "Substituent Effects on Non-Covalent Interactions with Aromatic Rings: Insights from Computational Chemistry". ChemPhysChem. 12 (17): 3116–30. doi:10.1002/cphc.201100542. PMID 21928437.
  12. ^ Hunter, C. A.; Low, C. M. R.; Rotger, C.; Vinter, J. G.; Zonta, C. (2002). "Supramolecular Chemistry and Self-assembly Special Feature: Substituent effects on cation–pi interactions: A quantitative study". Proceedings of the National Academy of Sciences. 99 (8): 4873–4876. Bibcode:2002PNAS...99.4873H. doi:10.1073/pnas.072647899. PMC 122686. PMID 11959939.
  13. ^ Dougherty, Dennis A. (1996). "cation–pi interactions in chemistry and biology: A new view of benzene, Phe, Tyr, and Trp". Science. 271 (5246): 163–168. Bibcode:1996Sci...271..163D. doi:10.1126/science.271.5246.163. PMID 8539615. S2CID 9436105.
  14. ^ Marshall, Michael S.; Steele, Ryan P.; Thanthiriwatte, Kanchana S.; Sherrill, C. David (2009). "Potential Energy Curves for Cation−π Interactions: Off-Axis Configurations Are Also Attractive". The Journal of Physical Chemistry A. 113 (48): 13628–32. Bibcode:2009JPCA..11313628M. doi:10.1021/jp906086x. PMID 19886621.
  15. ^ Lund-Katz, S; Phillips, MC; Mishra, VK; Segrest, JP; Anantharamaiah, GM (1995). "Microenvironments of basic amino acids in amphipathic alpha-helices bound to phospholipid: 13C NMR studies using selectively labeled peptides". Biochemistry. 34 (28): 9219–9226. doi:10.1021/bi00028a035. PMID 7619823.
  16. ^ M. M. Gromiha; C. Santhosh; S. Ahmad (2004). "Structural analysis of cation–π interactions in DNA binding proteins". Int. J. Biol. Macromol. 34 (3): 203–11. doi:10.1016/j.ijbiomac.2004.04.003. PMID 15225993.
  17. ^ J. P. Gallivan; D. A. Dougherty (1999). "Cation–π interactions in structural biology". PNAS. 96 (17): 9459–9464. Bibcode:1999PNAS...96.9459G. doi:10.1073/pnas.96.17.9459. PMC 22230. PMID 10449714.
  18. ^ Burley, SK; Petsko, GA (1986). "Amino-aromatic interactions in proteins". FEBS Lett. 203 (2): 139–143. doi:10.1016/0014-5793(86)80730-X. PMID 3089835.
  19. ^ Brocchieri, L; Karlin, S (1994). "Geometry of interplanar residue contacts in protein structures". Proc. Natl. Acad. Sci. USA. 91 (20): 9297–9301. Bibcode:1994PNAS...91.9297B. doi:10.1073/pnas.91.20.9297. PMC 44799. PMID 7937759.
  20. ^ Karlin, S; Zuker, M; Brocchieri, L (1994). "Measuring residue associations in protein structures. Possible implications for protein folding". J. Mol. Biol. 239 (2): 227–248. doi:10.1006/jmbi.1994.1365. PMID 8196056.
  21. ^ Nandi, CL; Singh, J; Thornton, JM (1993). "Atomic environments of arginine side chains in proteins". Protein Eng. 6 (3): 247–259. doi:10.1093/protein/6.3.247. PMID 8506259.
  22. ^ Zhong, W; Gallivan, JP; Zhang, Y; Li, L; Lester, HA; Dougherty, DA (1998). "From ab initio quantum mechanics to molecular neurobiology: A cation–π binding site in the nicotinic receptor". Proc. Natl. Acad. Sci. USA. 95 (21): 12088–12093. Bibcode:1998PNAS...9512088Z. doi:10.1073/pnas.95.21.12088. PMC 22789. PMID 9770444.
  23. ^ D. L. Beene; G. S. Brandt; W. Zhong; N. M. Zacharias; H. A. Lester; D. A. Dougherty (2002). "Cation–π Interactions in Ligand Recognition by Serotonergic (5-HT3A) and Nicotinic Acetylcholine Receptors: The Anomalous Binding Properties of Nicotine". Biochemistry. 41 (32): 10262–9. doi:10.1021/bi020266d. PMID 12162741.
  24. ^ Xiu, Xinan; Puskar, Nyssa L.; Shanata, Jai A. P.; Lester, Henry A.; Dougherty, Dennis A. (2009). "Nicotine Binding to Brain Receptors Requires a Strong Cation–π Interaction". Nature. 458 (7237): 534–7. Bibcode:2009Natur.458..534X. doi:10.1038/nature07768. PMC 2755585. PMID 19252481.
  25. ^ Di Wu, W.; Hu, Q.; Yan, Z.; Chen, W.; Yan, C.; Huang, X.; Zhang, J.; Yang, P.; Deng, H.; Wang, J.; Deng, X.; Shi, Y. (2012). "Structural basis of ultraviolet-B perception by UVR8". Nature. 484 (7393): 214–219. Bibcode:2012Natur.484..214D. doi:10.1038/nature10931. PMID 22388820. S2CID 2971536.
  26. ^ Waksman, G; Kominos, D; Robertson, SC; Pant, N; Baltimore, D; Birge, RB; Cowburn, D; Hanafusa, H; et al. (1992). "Crystal structure of the phosphotyrosine recognition domain SH2 of v-src complexed with tyrosine-phosphorylated peptides". Nature. 358 (6388): 646–653. Bibcode:1992Natur.358..646W. doi:10.1038/358646a0. PMID 1379696. S2CID 4329216.
  27. ^ Wendt, K. U.; Poralla, K; Schulz, GE (1997). "Structure and Function of a Squalene Cyclase". Science. 277 (5333): 1811–5. doi:10.1126/science.277.5333.1811. PMID 9295270.
  28. ^ Aoki, K; K. Muyayama; H. Nishiyama (1995). "Cation–? interaction between the trimethylammonium moiety and the aromatic ring within lndole-3-acetic acid choline ester, a model compound for molecular recognition between acetylcholine and its esterase: an X-ray study". Journal of the Chemical Society, Chemical Communications (21): 2221–2222. doi:10.1039/c39950002221.
  29. ^ McCurdy, Alison; Jimenez, Leslie; Stauffer, David A.; Dougherty, Dennis A. (1992). "Biomimetic catalysis of SN2 reactions through cation–.pi. Interactions. The role of polarizability in catalysis". Journal of the American Chemical Society. 114 (26): 10314–10321. doi:10.1021/ja00052a031.
  30. ^ Fiedler (2005). "Selective Molecular Recognition, C-H Bond Activation, and Catalysis in Nanoscale Reaction Vessels". Accounts of Chemical Research. 38 (4): 351–360. CiteSeerX 10.1.1.455.402. doi:10.1021/ar040152p. PMID 15835881. S2CID 2954569.
  31. ^ Ashton, P. R., Goodnow, T. T., Kaifer, A. E., Reddington, M. V., Slawin, A. M. Z., Spencer, N., Stoddart, J. F., Vicent, Ch. and Williams, D. J. Angew. Chem. Int. Ed. 1989, 28, 1396–1399.
  32. ^ Knowles, Robert R.; Lin, Song; Jacobsen, Eric N. (2010). "Enantioselective Thiourea-Catalyzed Cationic Polycyclizations". Journal of the American Chemical Society. 132 (14): 5030–2. doi:10.1021/Ja101256v. PMC 2989498. PMID 20369901.
  33. ^ D. Quiñonero; C. Garau; C. Rotger; A. Frontera; P. Ballester; A. Costa; P. M. Deyà (2002). "Anion-π Interactions: Do They Exist?". Angew. Chem. Int. Ed. 41 (18): 3389–3392. doi:10.1002/1521-3773(20020916)41:18<3389::AID-ANIE3389>3.0.CO;2-S. PMID 12298041.
  34. ^ P. de Hoog; P. Gamez; I. Mutikainen; U. Turpeinen; J. Reedijk (2004). "An Aromatic Anion Receptor: Anion-π Interactions do Exist". Angew. Chem. 116 (43): 5939–5941. Bibcode:2004AngCh.116.5939D. doi:10.1002/ange.200460486.

Sources edit

  • J. C. Ma; D. A. Dougherty (1997). "The Cation–π Interaction". Chem. Rev. 97 (5): 1303–1324. doi:10.1021/cr9603744. PMID 11851453..
  • Dougherty, D. A.; Stauffer, D. A. (Dec 1990). "Acetylcholine binding by a synthetic receptor: implications for biological recognition". Science. 250 (4987): 1558–1560. Bibcode:1990Sci...250.1558D. doi:10.1126/science.2274786. ISSN 0036-8075. PMID 2274786. S2CID 20210121.

cation, interaction, noncovalent, molecular, interaction, between, face, electron, rich, system, benzene, ethylene, acetylene, adjacent, cation, this, interaction, example, noncovalent, bonding, between, monopole, cation, quadrupole, system, bonding, energies,. Cation p interaction is a noncovalent molecular interaction between the face of an electron rich p system e g benzene ethylene acetylene and an adjacent cation e g Li Na This interaction is an example of noncovalent bonding between a monopole cation and a quadrupole p system Bonding energies are significant with solution phase values falling within the same order of magnitude as hydrogen bonds and salt bridges Similar to these other non covalent bonds cation p interactions play an important role in nature particularly in protein structure molecular recognition and enzyme catalysis The effect has also been observed and put to use in synthetic systems 1 2 Cation p interaction between benzene and a sodium cation The p system above and below the benzene ring leads to a quadrupole charge distribution Contents 1 Origin of the effect 2 Nature of the cation p interaction 3 Factors influencing the cation p bond strength 3 1 Nature of the cation 3 2 Solvation effects 3 3 Nature of the p system 3 3 1 Quadrupole moment 3 3 2 Substituents on the aromatic ring 3 3 3 Binding with heteroaromatic systems 3 4 Geometry 4 Relative interaction strength 5 In nature 5 1 Role in protein structure 5 2 Molecular recognition and signaling 5 3 Enzyme catalysis 6 In synthetic systems 6 1 Solid state structures 6 2 Supramolecular receptors 6 3 Use of p p CH p and p cation interactions in supramolecular assembly 6 4 Organic synthesis and catalysis 7 Anion p interaction 8 See also 9 References 10 SourcesOrigin of the effect editBenzene the model p system has no permanent dipole moment as the contributions of the weakly polar carbon hydrogen bonds cancel due to molecular symmetry However the electron rich p system above and below the benzene ring hosts a partial negative charge A counterbalancing positive charge is associated with the plane of the benzene atoms resulting in an electric quadrupole a pair of dipoles aligned like a parallelogram so there is no net molecular dipole moment The negatively charged region of the quadrupole can then interact favorably with positively charged species a particularly strong effect is observed with cations of high charge density 2 Nature of the cation p interaction editThe most studied cation p interactions involve binding between an aromatic p system and an alkali metal or nitrogenous cation The optimal interaction geometry places the cation in van der Waals contact with the aromatic ring centered on top of the p face along the 6 fold axis 3 Studies have shown that electrostatics dominate interactions in simple systems and relative binding energies correlate well with electrostatic potential energy 4 5 The Electrostatic Model developed by Dougherty and coworkers describes trends in binding energy based on differences in electrostatic attraction It was found that interaction energies of cation p pairs correlate well with electrostatic potential above the p face of arenes for eleven Na aromatic adducts the variation in binding energy between the different adducts could be completely rationalized by electrostatic differences Practically this allows trends to be predicted qualitatively based on visual representations of electrostatic potential maps for a series of arenes Electrostatic attraction is not the only component of cation p bonding For example 1 3 5 trifluorobenzene interacts with cations despite having a negligible quadrupole moment While non electrostatic forces are present these components remain similar over a wide variety of arenes making the electrostatic model a useful tool in predicting relative binding energies The other effects contributing to binding are not well understood Polarization donor acceptor permanent dead link and charge transfer interactions have been implicated however energetic trends do not track well with the ability of arenes and cations to take advantage of these effects For example if induced dipole was a controlling effect aliphatic compounds such as cyclohexane should be good cation p partners but are not 4 The cation p interaction is noncovalent and is therefore fundamentally different than bonding between transition metals and p systems Transition metals have the ability to share electron density with p systems through d orbitals creating bonds that are highly covalent in character and cannot be modeled as a cation p interaction Factors influencing the cation p bond strength editSeveral criteria influence the strength of the bonding the nature of the cation solvation effects the nature of the p system and the geometry of the interaction Nature of the cation edit From electrostatics Coulomb s law smaller and more positively charged cations lead to larger electrostatic attraction Since cation p interactions are predicted by electrostatics it follows that cations with larger charge density interact more strongly with p systems The following table shows a series of Gibbs free energy of binding between benzene and several cations in the gas phase 2 6 For a singly charged species the gas phase interaction energy correlates with the ionic radius rion displaystyle r mathrm ion nbsp non spherical ionic radii are approximate 7 8 M Li Na K NH4 Rb NMe4 DG kcal mol 38 27 19 19 16 9rion displaystyle r mathrm ion nbsp A 0 76 1 02 1 38 1 43 1 52 2 45 This trend supports the idea that coulombic forces play a central role in interaction strength since for other types of bonding one would expect the larger and more polarizable ions to have greater binding energies Solvation effects edit The nature of the solvent also determines the absolute and relative strength of the bonding Most data on cation p interaction is acquired in the gas phase as the attraction is most pronounced in that case Any intermediating solvent molecule will attenuate the effect because the energy gained by the cation p interaction is partially offset by the loss of solvation energy For a given cation p adduct the interaction energy decreases with increasing solvent polarity This can be seen by the following calculated interaction energies of methylammonium and benzene in a variety of solvents 9 nbsp Calculated interaction energies of methylamonium and benzene in a variety of solventsAdditionally the trade off between solvation and the cation p effect results in a rearrangement of the order of interaction strength for a series of cations While in the gas phase the most densely charged cations have the strongest cation p interaction these ions also have a high desolvation penalty This is demonstrated by the relative cation p bond strengths in water for alkali metals 10 K gt Rb Na Li displaystyle ce K gt Rb gg Na Li nbsp Nature of the p system edit Quadrupole moment edit Comparing the quadrupole moment of different arenes is a useful qualitative tool to predict trends in cation p binding since it roughly correlates with interaction strength Arenes with larger quadrupole moments are generally better at binding cations However a quadrupole ion model system cannot be used to quantitatively model cation p interactions Such models assume point charges and are therefore not valid given the short cation p bond distance In order to use electrostatics to predict energies the full electrostatic potential surface must be considered rather than just the quadrupole moment as a point charge 2 Substituents on the aromatic ring edit nbsp Binding energy in kcal mol for Na to benzene with prototypical substituents 4 The electronic properties of the substituents also influence the strength of the attraction 11 Electron withdrawing groups for example cyano CN weaken the interaction while electron donating substituents for example amino NH2 strengthen the cation p binding This relationship is illustrated quantitatively in the margin for several substituents The electronic trends in cation p binding energy are not quite analogous to trends in aryl reactivity Indeed the effect of resonance participation by a substituent does not contribute substantively to cation p binding despite being very important in many chemical reactions with arenes This was shown by the observation that cation p interaction strength for a variety of substituted arenes correlates with the smeta Hammett parameter This parameter is meant to illustrate the inductive effects of functional groups on an aryl ring 4 The origin of substituent effects in cation p interactions has often been attributed to polarization from electron donation or withdrawal into or out of the p system 12 This explanation makes intuitive sense but subsequent studies have indicated that it is flawed Recent computational work by Wheeler and Houk strongly indicate that the effect is primarily due to direct through space interaction between the cation and the substituent dipole In this study calculations that modeled unsubstituted benzene plus interaction with a molecule of H X situated where a substituent would be corrected for extra hydrogen atoms accounted for almost all of the cation p binding trend For very strong pi donors or acceptors this model was not quite able account for the whole interaction in these cases polarization may be a more significant factor 5 Binding with heteroaromatic systems edit Heterocycles are often activated towards cation p binding when the lone pair on the heteroatom is in incorporated into the aromatic system e g indole pyrrole Conversely when the lone pair does not contribute to aromaticity e g pyridine the electronegativity of the heteroatom wins out and weakens the cation p binding ability Since several classically electron rich heterocycles are poor donors when it comes to cation p binding one cannot predict cation p trends based on heterocycle reactivity trends Fortunately the aforementioned subtleties are manifested in the electrostatic potential surfaces of relevant heterocycles 2 nbsp cation heterocycle interaction is not always a cation p interaction in some cases it is more favorable for the ion to be bound directly to a lone pair For example this is thought to be the case in pyridine Na complexes Geometry edit cation p interactions have an approximate distance dependence of 1 rn where n lt 2 The interaction is less sensitive to distance than a simple ion quadrupole interaction which has 1 r3 dependence 13 A study by Sherrill and coworkers probed the geometry of the interaction further confirming that cation p interactions are strongest when the cation is situated perpendicular to the plane of atoms 8 0 degrees in the image below Variations from this geometry still exhibit a significant interaction which weakens as 8 angle approaches 90 degrees For off axis interactions the preferred ϕ places the cation between two H atoms Equilibrium bond distances also increase with off axis angle Energies where the cation is coplanar with the carbon ring are saddle points on the potential energy surface which is consistent with the idea that interaction between a cation and the positive region of the quadrupole is not ideal 14 nbsp Relative interaction strength editIn aqueous media the cation p interaction is comparable to and potentially stronger than ammonium carboxylate salt bridges Computed values below show that as solvent polarity increases the strength of the cation p complex decreases less dramatically This trend can be rationalized by desolvation effects salt bridge formation has a high desolvation penalty for both charged species whereas the cation p complex would only pay a significant penalty for the cation 9 nbsp In nature editNature s building blocks contain aromatic moieties in high abundance Recently it has become clear that many structural features that were once thought to be purely hydrophobic in nature are in fact engaging in cation p interactions The amino acid side chains of phenylalanine tryptophan tyrosine histidine are capable of binding to cationic species such as charged amino acid side chains metal ions small molecule neurotransmitters and pharmaceutical agents In fact macromolecular binding sites that were hypothesized to include anionic groups based on affinity for cations have been found to consist of aromatic residues instead in multiple cases Cation p interactions can tune the pKa of nitrogenous side chains increasing the abundance of the protonated form this has implications for protein structure and function 15 While less studied in this context the DNA bases are also able to participate in cation p interactions 16 17 Role in protein structure edit Early evidence that cation p interactions played a role in protein structure was the observation that in crystallographic data aromatic side chains appear in close contact with nitrogen containing side chains which can exist as protonated cationic species with disproportionate frequency A study published in 1986 by Burley and Petsko looked at a diverse set of proteins and found that 50 of aromatic residues Phe Tyr and Trp were within 6A of amino groups Furthermore approximately 25 of nitrogen containing side chains Lys Asn Gln and His were within van der Waals contact with aromatics and 50 of Arg in contact with multiple aromatic residues 2 on average 18 Studies on larger data sets found similar trends including some dramatic arrays of alternating stacks of cationic and aromatic side chains In some cases the N H hydrogens were aligned toward aromatic residues and in others the cationic moiety was stacked above the p system A particularly strong trend was found for close contacts between Arg and Trp The guanidinium moiety of Arg in particular has a high propensity to be stacked on top of aromatic residues while also hydrogen bonding with nearby oxygen atoms 19 20 21 Molecular recognition and signaling edit nbsp Cationic Acetylcholine binding to a tryptophan residue of the nicotinamide acetylcholine receptor via a cation p effect An example of cation p interactions in molecular recognition is seen in the nicotinic acetylcholine receptor nAChR which binds its endogenous ligand acetylcholine a positively charged molecule via a cation p interaction to the quaternary ammonium The nAChR neuroreceptor is a well studied ligand gated ion channel that opens upon acetylcholine binding Acetylcholine receptors are therapeutic targets for a large host of neurological disorders including Parkinson s disease Alzheimer s disease schizophrenia depression and autism Studies by Dougherty and coworkers confirmed that cation p interactions are important for binding and activating nAChR by making specific structural variations to a key tryptophan residue and correlating activity results with cation p binding ability 22 The nAChR is especially important in binding nicotine in the brain and plays a key role in nicotine addiction Nicotine has a similar pharmacophore to acetylcholine especially when protonated Strong evidence supports cation p interactions being central to the ability of nicotine to selectively activate brain receptors without affecting muscle activity 23 24 nbsp A further example is seen in the plant UV B sensing protein UVR8 Several tryptophan residues interact via cation p interactions with arginine residues which in turn form salt bridges with acidic residues on a second copy of the protein It has been proposed 25 that absorption of a photon by the tryptophan residues disrupts this interaction and leads to dissociation of the protein dimer Cation p binding is also thought to be important in cell surface recognition 2 26 Enzyme catalysis edit Cation p interactions can catalyze chemical reactions by stabilizing buildup of positive charge in transition states This kind of effect is observed in enzymatic systems For example acetylcholine esterase contains important aromatic groups that bind quaternary ammonium in its active site 2 Polycyclization enzymes also rely on cation p interactions Since proton triggered polycyclizations of squalene proceed through a potentially concerted cationic cascade cation p interactions are ideal for stabilizing this dispersed positive charge The crystal structure of squalene hopene cyclase shows that the active site is lined with aromatic residues 27 nbsp Cyclization of squalene to form hopeneIn synthetic systems editSolid state structures edit Cation p interactions have been observed in the crystals of synthetic molecules as well For example Aoki and coworkers compared the solid state structures of Indole 3 acetic acid choline ester and an uncharged analogue In the charged species an intramolecular cation p interaction with the indole is observed as well as an interaction with the indole moiety of the neighboring molecule in the lattice In the crystal of the isosteric neutral compound the same folding is not observed and there are no interactions between the tert butyl group and neighboring indoles 28 nbsp Cation p interaction in indole 3 acetic acid choline ester compared to neutral analogSupramolecular receptors edit nbsp Cyclophane host guest complexSome of the first studies on the cation p interaction involved looking at the interactions of charged nitrogenous molecules in cyclophane host guest chemistry It was found that even when anionic solubilizing groups were appended to aromatic host capsules cationic guests preferred to associate with the p system in many cases The type of host shown to the right was also able to catalyze N alkylation reactions to form cationic products 29 More recently cation p centered substrate binding and catalysis has been implicated in supramolecular metal ligand cluster catalyst systems developed by Raymond and Bergman 30 Use of p p CH p and p cation interactions in supramolecular assembly edit p systems are important building blocks in supramolecular assembly because of their versatile noncovalent interactions with various functional groups Particularly p p CH p and p cation interactions are widely used in supramolecular assembly and recognition nbsp Fig 1 Examples of p p CH p and p cation interactionsp p interaction concerns the direct interactions between two amp pi systems and cation p interaction arises from the electrostatic interaction of a cation with the face of the p system Unlike these two interactions the CH p interaction arises mainly from charge transfer between the C H orbital and the p system A notable example of applying p p interactions in supramolecular assembly is the synthesis of catenane The major challenge for the synthesis of catenane is to interlock molecules in a controlled fashion Stoddart and co workers developed a series of systems utilizing the strong p p interactions between electron rich benzene derivatives and electron poor pyridinium rings 31 2 Catanene was synthesized by reacting bis pyridinium A bisparaphenylene 34 crown 10 B and 1 4 bis bromomethyl benzene C Fig 2 The p p interaction between A and B directed the formation of an interlocked template intermediate that was further cyclized by substitution reaction with compound C to generate the 2 catenane product nbsp Fig 2 The Stoddart synthesis of 2 catenaneOrganic synthesis and catalysis edit Cation p interactions have likely been important though unnoticed in a multitude of organic reactions historically Recently however attention has been drawn to potential applications in catalyst design In particular noncovalent organocatalysts have been found to sometimes exhibit reactivity and selectivity trends that correlate with cation p binding properties A polycyclization developed by Jacobsen and coworkers shows a particularly strong cation p effect using the catalyst shown below 32 nbsp Anion p interaction edit nbsp Quadrupole moments of benzene and hexafluorobenzene The polarity is inverted due to differences in electronegativity for hydrogen and fluorine relative to carbon the inverted quadrupole moment of hexafluorobenzene is necessary for anion pi interactions In many respects anion p interaction is the opposite of cation p interaction although the underlying principles are identical Significantly fewer examples are known to date In order to attract a negative charge the charge distribution of the p system has to be reversed This is achieved by placing several strong electron withdrawing substituents along the p system e g hexafluorobenzene 33 The anion p effect is advantageously exploited in chemical sensors for specific anions 34 See also editStacking chemistry Salt bridge protein References edit Eric V Anslyn Dennis A Dougherty 2004 Modern Physical Organic Chemistry University Science Books ISBN 978 1 891389 31 3 a b c d e f g Dougherty D A J C Ma 1997 The Cation p Interaction Chemical Reviews 97 5 1303 1324 doi 10 1021 cr9603744 PMID 11851453 Tsuzuki Seiji Yoshida Masaru Uchimaru Tadafumi Mikami Masuhiro 2001 The Origin of the Cation p Interaction The Significant Importance of the Induction in Li and Na Complexes The Journal of Physical Chemistry A 105 4 769 773 Bibcode 2001JPCA 105 769T doi 10 1021 jp003287v a b c d S Mecozzi A P West D A Dougherty 1996 Cation p Interactions in Simple Aromatics Electrostatics Provide a Predictive Tool Journal of the American Chemical Society 118 9 2307 2308 doi 10 1021 ja9539608 a b S E Wheeler K N Houk 2009 Substituent Effects in Cation p Interactions and Electrostatic Potentials above the Center of Substituted Benzenes Are Due Primarily to through Space Effects of the Substituents J Am Chem Soc 131 9 3126 7 doi 10 1021 ja809097r PMC 2787874 PMID 19219986 J C Amicangelo P B Armentrout 2000 Absolute Binding Energies of Alkali Metal Cation Complexes with Benzene Determined by Threshold Collision Induced Dissociation Experiments and ab Initio Theory J Phys Chem A 104 48 11420 11432 Bibcode 2000JPCA 10411420A doi 10 1021 jp002652f Robinson RA Stokes RH Electrolyte solutions UK Butterworth Publications Pitman 1959 Clays and Clay Minerals Vol 45 No 6 859 866 1997 a b Gallivan Justin P Dougherty Dennis A 2000 A Computational Study of Cation p Interactions vs Salt Bridges in Aqueous Media Implications for Protein Engineering PDF Journal of the American Chemical Society 122 5 870 874 doi 10 1021 ja991755c Kumpf R Dougherty D 1993 A mechanism for ion selectivity in potassium channels Computational studies of cation pi interactions Science 261 5129 1708 10 Bibcode 1993Sci 261 1708K doi 10 1126 science 8378771 PMID 8378771 Raju Rajesh K Bloom Jacob W G An Yi Wheeler Steven E 2011 Substituent Effects on Non Covalent Interactions with Aromatic Rings Insights from Computational Chemistry ChemPhysChem 12 17 3116 30 doi 10 1002 cphc 201100542 PMID 21928437 Hunter C A Low C M R Rotger C Vinter J G Zonta C 2002 Supramolecular Chemistry and Self assembly Special Feature Substituent effects on cation pi interactions A quantitative study Proceedings of the National Academy of Sciences 99 8 4873 4876 Bibcode 2002PNAS 99 4873H doi 10 1073 pnas 072647899 PMC 122686 PMID 11959939 Dougherty Dennis A 1996 cation pi interactions in chemistry and biology A new view of benzene Phe Tyr and Trp Science 271 5246 163 168 Bibcode 1996Sci 271 163D doi 10 1126 science 271 5246 163 PMID 8539615 S2CID 9436105 Marshall Michael S Steele Ryan P Thanthiriwatte Kanchana S Sherrill C David 2009 Potential Energy Curves for Cation p Interactions Off Axis Configurations Are Also Attractive The Journal of Physical Chemistry A 113 48 13628 32 Bibcode 2009JPCA 11313628M doi 10 1021 jp906086x PMID 19886621 Lund Katz S Phillips MC Mishra VK Segrest JP Anantharamaiah GM 1995 Microenvironments of basic amino acids in amphipathic alpha helices bound to phospholipid 13C NMR studies using selectively labeled peptides Biochemistry 34 28 9219 9226 doi 10 1021 bi00028a035 PMID 7619823 M M Gromiha C Santhosh S Ahmad 2004 Structural analysis of cation p interactions in DNA binding proteins Int J Biol Macromol 34 3 203 11 doi 10 1016 j ijbiomac 2004 04 003 PMID 15225993 J P Gallivan D A Dougherty 1999 Cation p interactions in structural biology PNAS 96 17 9459 9464 Bibcode 1999PNAS 96 9459G doi 10 1073 pnas 96 17 9459 PMC 22230 PMID 10449714 Burley SK Petsko GA 1986 Amino aromatic interactions in proteins FEBS Lett 203 2 139 143 doi 10 1016 0014 5793 86 80730 X PMID 3089835 Brocchieri L Karlin S 1994 Geometry of interplanar residue contacts in protein structures Proc Natl Acad Sci USA 91 20 9297 9301 Bibcode 1994PNAS 91 9297B doi 10 1073 pnas 91 20 9297 PMC 44799 PMID 7937759 Karlin S Zuker M Brocchieri L 1994 Measuring residue associations in protein structures Possible implications for protein folding J Mol Biol 239 2 227 248 doi 10 1006 jmbi 1994 1365 PMID 8196056 Nandi CL Singh J Thornton JM 1993 Atomic environments of arginine side chains in proteins Protein Eng 6 3 247 259 doi 10 1093 protein 6 3 247 PMID 8506259 Zhong W Gallivan JP Zhang Y Li L Lester HA Dougherty DA 1998 From ab initio quantum mechanics to molecular neurobiology A cation p binding site in the nicotinic receptor Proc Natl Acad Sci USA 95 21 12088 12093 Bibcode 1998PNAS 9512088Z doi 10 1073 pnas 95 21 12088 PMC 22789 PMID 9770444 D L Beene G S Brandt W Zhong N M Zacharias H A Lester D A Dougherty 2002 Cation p Interactions in Ligand Recognition by Serotonergic 5 HT3A and Nicotinic Acetylcholine Receptors The Anomalous Binding Properties of Nicotine Biochemistry 41 32 10262 9 doi 10 1021 bi020266d PMID 12162741 Xiu Xinan Puskar Nyssa L Shanata Jai A P Lester Henry A Dougherty Dennis A 2009 Nicotine Binding to Brain Receptors Requires a Strong Cation p Interaction Nature 458 7237 534 7 Bibcode 2009Natur 458 534X doi 10 1038 nature07768 PMC 2755585 PMID 19252481 Di Wu W Hu Q Yan Z Chen W Yan C Huang X Zhang J Yang P Deng H Wang J Deng X Shi Y 2012 Structural basis of ultraviolet B perception by UVR8 Nature 484 7393 214 219 Bibcode 2012Natur 484 214D doi 10 1038 nature10931 PMID 22388820 S2CID 2971536 Waksman G Kominos D Robertson SC Pant N Baltimore D Birge RB Cowburn D Hanafusa H et al 1992 Crystal structure of the phosphotyrosine recognition domain SH2 of v src complexed with tyrosine phosphorylated peptides Nature 358 6388 646 653 Bibcode 1992Natur 358 646W doi 10 1038 358646a0 PMID 1379696 S2CID 4329216 Wendt K U Poralla K Schulz GE 1997 Structure and Function of a Squalene Cyclase Science 277 5333 1811 5 doi 10 1126 science 277 5333 1811 PMID 9295270 Aoki K K Muyayama H Nishiyama 1995 Cation interaction between the trimethylammonium moiety and the aromatic ring within lndole 3 acetic acid choline ester a model compound for molecular recognition between acetylcholine and its esterase an X ray study Journal of the Chemical Society Chemical Communications 21 2221 2222 doi 10 1039 c39950002221 McCurdy Alison Jimenez Leslie Stauffer David A Dougherty Dennis A 1992 Biomimetic catalysis of SN2 reactions through cation pi Interactions The role of polarizability in catalysis Journal of the American Chemical Society 114 26 10314 10321 doi 10 1021 ja00052a031 Fiedler 2005 Selective Molecular Recognition C H Bond Activation and Catalysis in Nanoscale Reaction Vessels Accounts of Chemical Research 38 4 351 360 CiteSeerX 10 1 1 455 402 doi 10 1021 ar040152p PMID 15835881 S2CID 2954569 Ashton P R Goodnow T T Kaifer A E Reddington M V Slawin A M Z Spencer N Stoddart J F Vicent Ch and Williams D J Angew Chem Int Ed 1989 28 1396 1399 Knowles Robert R Lin Song Jacobsen Eric N 2010 Enantioselective Thiourea Catalyzed Cationic Polycyclizations Journal of the American Chemical Society 132 14 5030 2 doi 10 1021 Ja101256v PMC 2989498 PMID 20369901 D Quinonero C Garau C Rotger A Frontera P Ballester A Costa P M Deya 2002 Anion p Interactions Do They Exist Angew Chem Int Ed 41 18 3389 3392 doi 10 1002 1521 3773 20020916 41 18 lt 3389 AID ANIE3389 gt 3 0 CO 2 S PMID 12298041 P de Hoog P Gamez I Mutikainen U Turpeinen J Reedijk 2004 An Aromatic Anion Receptor Anion p Interactions do Exist Angew Chem 116 43 5939 5941 Bibcode 2004AngCh 116 5939D doi 10 1002 ange 200460486 Sources editJ C Ma D A Dougherty 1997 The Cation p Interaction Chem Rev 97 5 1303 1324 doi 10 1021 cr9603744 PMID 11851453 Dougherty D A Stauffer D A Dec 1990 Acetylcholine binding by a synthetic receptor implications for biological recognition Science 250 4987 1558 1560 Bibcode 1990Sci 250 1558D doi 10 1126 science 2274786 ISSN 0036 8075 PMID 2274786 S2CID 20210121 Retrieved from https en wikipedia org w index php title Cation p interaction amp oldid 1170061202, wikipedia, wiki, book, books, library,

article

, read, download, free, free download, mp3, video, mp4, 3gp, jpg, jpeg, gif, png, picture, music, song, movie, book, game, games.