Stacking (chemistry)

Last updated

In chemistry, stacking refers to superposition of molecules or atomic sheets owing to attractive interactions between these molecules or sheets.

Contents

Metal dichalcogenide compounds

MoS2, the most common metal dichalcogenide, adopts a layered structure. Molybdenite-3D-balls.png
MoS2, the most common metal dichalcogenide, adopts a layered structure.

Metal dichalcogenides have the formula ME2, where M = a transition metal and E = S, Se, Te. [1] In terms of their electronic structures, these compounds are usually viewed as derivatives of M4+. They adopt stacked structures, which is relevant to their ability to undergo intercalation, e.g. by lithium, and their lubricating properties. The corresponding diselenides and even ditellurides are known, e.g., TiSe2, MoSe2, and WSe2.

Charge transfer salts

Edge-on view of portion of crystal structure of hexamethyleneTTF/TCNQ charge transfer salt, highlighting the segregated stacking. SegStackEdgeOnHMTFCQ.jpg
Edge-on view of portion of crystal structure of hexamethyleneTTF/TCNQ charge transfer salt, highlighting the segregated stacking.

A combination of tetracyanoquinodimethane (TCNQ) and tetrathiafulvalene (TTF) forms a strong charge-transfer complex referred to as TTF-TCNQ. [3] The solid shows almost metallic electrical conductance. In a TTF-TCNQ crystal, TTF and TCNQ molecules are arranged independently in separate parallel-aligned stacks, and an electron transfer occurs from donor (TTF) to acceptor (TCNQ) stacks. [4]

Graphite

Side view of ABA layer stacking in graphite Graphite-layers-side-3D-balls.png
Side view of ABA layer stacking in graphite

Graphite consists of stacked sheets of covalently bonded carbon. [5] [6] The individual layers are called graphene. In each layer, each carbon atom is bonded to three other atoms forming a continuous layer of sp2 bonded carbon hexagons, like a honeycomb lattice with a bond length of 0.142 nm, and the distance between planes is 0.335 nm. [7] Bonding between layers is relatively weak van der Waals bonds, which allows the graphene-like layers to be easily separated and to glide past each other. [8] Electrical conductivity perpendicular to the layers is consequently about 1000 times lower. [9]

Linear chain compounds

Linear chain compounds are materials composed of stacked arrays of metal-metal bonded molecules or ions. Such materials exhibit anisotropic electrical conductivity. [10] One example is Rh(acac)(CO)2 (acac = acetylacetonate, which stack with Rh···Rh distances of about 326  pm. [11] Classic examples include Krogmann's salt and Magnus's green salt.

Portion of the lattice of Dicarbonyl(acetylacetonato)rhodium(I) (
Rh(acac)(CO)2) showing the "stacking" of the individual planar units through
Rh***Rh interactions. ACABRH02.png
Portion of the lattice of Dicarbonyl(acetylacetonato)rhodium(I) (Rh(acac)(CO)2) showing the "stacking" of the individual planar units through Rh···Rh interactions.

π–π stacking is a noncovalent interaction between the pi bonds of aromatic rings. [12] Such "sandwich interactions" are however generally electrostatically repulsive. What is more commonly observed areeither a staggered stacking (parallel displaced) or pi-teeing (perpendicular T-shaped) interaction both of which are electrostatic attractive. [13] For example, the most commonly observed interactions between aromatic rings of amino acid residues in proteins is a staggered stacked followed by a perpendicular orientation. Sandwiched orientations are relatively rare. [14] Pi stacking is repulsive as it places carbon atoms with partial negative charges from one ring on top of other partial negatively charged carbon atoms from the second ring and hydrogen atoms with partial positive charges on top of other hydrogen atoms that likewise carry partial positive charges. [15]

The synthesis of catenane exploiting attractive pi-stacking interactions between electron-rich and electron-poor arenes. 24 fig. 2.png
The synthesis of catenane exploiting attractive pi-stacking interactions between electron-rich and electron-poor arenes.

π–π interactions play a role in supramolecular chemistry, specifically the synthesis of catenane. The major challenge for the synthesis of catenane is to interlock molecules in a controlled fashion. Attractive π–π interactions exist between electron-rich benzene derivatives and electron-poor pyridinium rings. [16] [2]Catanene was synthesized by treating 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.

See also

Related Research Articles

<span class="mw-page-title-main">Conjugated system</span> System of connected p-orbitals with delocalized electrons in a molecule

In theoretical chemistry, a conjugated system is a system of connected p-orbitals with delocalized electrons in a molecule, which in general lowers the overall energy of the molecule and increases stability. It is conventionally represented as having alternating single and multiple bonds. Lone pairs, radicals or carbenium ions may be part of the system, which may be cyclic, acyclic, linear or mixed. The term "conjugated" was coined in 1899 by the German chemist Johannes Thiele.

<span class="mw-page-title-main">Aromaticity</span> Chemical property

In organic chemistry, aromaticity is a chemical property describing the way in which a conjugated ring of unsaturated bonds, lone pairs, or empty orbitals exhibits a stabilization stronger than would be expected by the stabilization of conjugation alone. The earliest use of the term was in an article by August Wilhelm Hofmann in 1855. There is no general relationship between aromaticity as a chemical property and the olfactory properties of such compounds.

<span class="mw-page-title-main">Pi bond</span> Type of chemical bond

In chemistry, pi bonds are covalent chemical bonds, in each of which two lobes of an orbital on one atom overlap with two lobes of an orbital on another atom, and in which this overlap occurs laterally. Each of these atomic orbitals has an electron density of zero at a shared nodal plane that passes through the two bonded nuclei. This plane also is a nodal plane for the molecular orbital of the pi bond. Pi bonds can form in double and triple bonds but do not form in single bonds in most cases.

In chemistry, molecular orbital theory (MO theory or MOT) is a method for describing the electronic structure of molecules using quantum mechanics. It was proposed early in the 20th century. The MOT explains the paramagnetic nature of O2, which VSEPR theory cannot explain.

<span class="mw-page-title-main">Delocalized electron</span> Electrons that are not associated with a single atom or covalent bond

In chemistry, delocalized electrons are electrons in a molecule, ion or solid metal that are not associated with a single atom or a covalent bond.

<span class="mw-page-title-main">Sigma bond</span> Covalent chemical bond

In chemistry, sigma bonds or sigma overlap are the strongest type of covalent chemical bond. They are formed by head-on overlapping between atomic orbitals along the internuclear axis. Sigma bonding is most simply defined for diatomic molecules using the language and tools of symmetry groups. In this formal approach, a σ-bond is symmetrical with respect to rotation about the bond axis. By this definition, common forms of sigma bonds are s+s, pz+pz, s+pz and dz2+dz2 . Quantum theory also indicates that molecular orbitals (MO) of identical symmetry actually mix or hybridize. As a practical consequence of this mixing of diatomic molecules, the wavefunctions s+s and pz+pz molecular orbitals become blended. The extent of this mixing depends on the relative energies of the MOs of like symmetry.

<span class="mw-page-title-main">Charge-transfer complex</span> Association of molecules in which a fraction of electronic charge is transferred between them

In chemistry, charge-transfer (CT) complex, or electron donor-acceptor complex, describes a type of supramolecular assembly of two or more molecules or ions. The assembly consists of two molecules that self-attract through electrostatic forces, i.e., one has at least partial negative charge and the partner has partial positive charge, referred to respectively as the electron acceptor and electron donor. In some cases, the degree of charge transfer is "complete", such that the CT complex can be classified as a salt. In other cases, the charge-transfer association is weak, and the interaction can be disrupted easily by polar solvents.

<span class="mw-page-title-main">Intercalation (chemistry)</span> Reversible insertion of an ion into a material with layered structure

Intercalation is the reversible inclusion or insertion of a molecule into layered materials with layered structures. Examples are found in graphite and transition metal dichalcogenides.

In chemistry, a non-covalent interaction differs from a covalent bond in that it does not involve the sharing of electrons, but rather involves more dispersed variations of electromagnetic interactions between molecules or within a molecule. The chemical energy released in the formation of non-covalent interactions is typically on the order of 1–5 kcal/mol. Non-covalent interactions can be classified into different categories, such as electrostatic, π-effects, van der Waals forces, and hydrophobic effects.

<span class="mw-page-title-main">Graphite intercalation compound</span> Class of chemical compounds

In the area of solid state chemistry, graphite intercalation compounds are a family of materials prepared from graphite. In particular, the sheets of carbon that comprise graphite can be pried apart by the insertion (intercalation) of ions. The graphite is viewed as a host and the inserted ions as guests. The materials have the formula (guest)Cn where n ≥ 6. The insertion of the guests increases the distance between the carbon sheets. Common guests are reducing agents such as alkali metals. Strong oxidants also intercalate into graphite. Intercalation involves electron transfer into or out of the carbon sheets. So, in some sense, graphite intercalation compounds are salts. Intercalation is often reversible: the inserted ions can be removed and the sheets of carbon collapse to a graphite-like structure.

<span class="mw-page-title-main">Tetracyanoquinodimethane</span> Organic compound with formula C12H4N4

Tetracyanoquinodimethane (TCNQ) is an organic compound with the chemical formula (N≡C−)2C=C6H4=C(−C≡N)2. It is an orange crystalline solid. This cyanocarbon, a relative of para-quinone, is an electron acceptor that is used to prepare charge transfer salts, which are of interest in molecular electronics.

<span class="mw-page-title-main">Hapticity</span> Number of contiguous atoms in a ligand that bond to the central atom in a coordination complex

In coordination chemistry, hapticity is the coordination of a ligand to a metal center via an uninterrupted and contiguous series of atoms. The hapticity of a ligand is described with the Greek letter η ('eta'). For example, η2 describes a ligand that coordinates through 2 contiguous atoms. In general the η-notation only applies when multiple atoms are coordinated. In addition, if the ligand coordinates through multiple atoms that are not contiguous then this is considered denticity, and the κ-notation is used once again. When naming complexes care should be taken not to confuse η with μ ('mu'), which relates to bridging ligands.

<span class="mw-page-title-main">Cation–π interaction</span> Noncovalent molecular 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.

<span class="mw-page-title-main">Tetrathiafulvalene</span> Organosulfuric compound with formula C6H4S4

Tetrathiafulvalene (TTF) is an organosulfur compound with the formula 2. Studies on this heterocyclic compound contributed to the development of molecular electronics. TTF is related to the hydrocarbon fulvalene, (C5H4)2, by replacement of four CH groups with sulfur atoms. Over 10,000 scientific publications discuss TTF and its derivatives.

<span class="mw-page-title-main">Molecular solid</span> Solid consisting of discrete molecules

A molecular solid is a solid consisting of discrete molecules. The cohesive forces that bind the molecules together are van der Waals forces, dipole–dipole interactions, quadrupole interactions, π–π interactions, hydrogen bonding, halogen bonding, London dispersion forces, and in some molecular solids, coulombic interactions. Van der Waals, dipole interactions, quadrupole interactions, π–π interactions, hydrogen bonding, and halogen bonding are typically much weaker than the forces holding together other solids: metallic, ionic, and network solids.

A network solid or covalent network solid is a chemical compound in which the atoms are bonded by covalent bonds in a continuous network extending throughout the material. In a network solid there are no individual molecules, and the entire crystal or amorphous solid may be considered a macromolecule. Formulas for network solids, like those for ionic compounds, are simple ratios of the component atoms represented by a formula unit.

C<sub>70</sub> fullerene Chemical compound

C70 fullerene is the fullerene molecule consisting of 70 carbon atoms. It is a cage-like fused-ring structure which resembles a rugby ball, made of 25 hexagons and 12 pentagons, with a carbon atom at the vertices of each polygon and a bond along each polygon edge. A related fullerene molecule, named buckminsterfullerene (or C60 fullerene) consists of 60 carbon atoms.

<span class="mw-page-title-main">Titanium diselenide</span> Chemical compound

Titanium diselenide (TiSe2) also known as titanium(IV) selenide, is an inorganic compound of titanium and selenium. In this material selenium is viewed as selenide (Se2−) which requires that titanium exists as Ti4+. Titanium diselenide is a member of metal dichalcogenides, compounds that consist of a metal and an element of the chalcogen column within the periodic table. Many exhibit properties of potential value in battery technology, such as intercalation and electrical conductivity, although most applications focus on the less toxic and lighter disulfides, e.g. TiS2.

Cyclobis(paraquat-<i>p</i>-phenylene) Chemical compound

Cyclobis(paraquat-p-phenylene) belongs to the class of cyclophanes, and consists of aromatic units connected by methylene bridges. It is able to incorporate small guest molecule and has played an important role in host–guest chemistry and supramolecular chemistry.

<span class="mw-page-title-main">Pi-stacking</span> Attractive interactions between aromatic rings

In chemistry, pi stacking refers to the presumptive attractive, noncovalent pi interactions between the pi bonds of aromatic rings. According to some authors direct stacking of aromatic rings is electrostatically repulsive.

References

  1. Wells, A.F. (1984) Structural Inorganic Chemistry, Oxford: Clarendon Press. ISBN   0-19-855370-6.
  2. D. Chasseau; G. Comberton; J. Gaultier; C. Hauw (1978). "Réexamen de la structure du complexe hexaméthylène-tétrathiafulvalène-tétracyanoquinodiméthane". Acta Crystallographica Section B. 34 (2): 689. Bibcode:1978AcCrB..34..689C. doi:10.1107/S0567740878003830.
  3. P. W. Anderson; P. A. Lee; M. Saitoh (1973). "Remarks on giant conductivity in TTF-TCNQ". Solid State Communications . 13 (5): 595–598. Bibcode:1973SSCom..13..595A. doi:10.1016/S0038-1098(73)80020-1.
  4. Van De Wouw, Heidi L.; Chamorro, Juan; Quintero, Michael; Klausen, Rebekka S. (2015). "Opposites Attract: Organic Charge Transfer Salts". Journal of Chemical Education. 92 (12): 2134–2139. Bibcode:2015JChEd..92.2134V. doi:10.1021/acs.jchemed.5b00340.
  5. Delhaes, Pierre (2000). "Polymorphism of carbon". In Delhaes, Pierre (ed.). Graphite and precursors. Gordon & Breach. pp. 1–24. ISBN   9789056992286.
  6. Pierson, Hugh O. (2012). Handbook of carbon, graphite, diamond, and fullerenes : properties, processing, and applications. Noyes Publications. pp. 40–41. ISBN   9780815517399.
  7. Delhaes, P. (2001). Graphite and Precursors. CRC Press. ISBN   978-90-5699-228-6.
  8. Chung, D. D. L. (2002). "Review Graphite". Journal of Materials Science. 37 (8): 1475–1489. doi:10.1023/A:1014915307738. S2CID   189839788.
  9. Pierson, Hugh O. (1993). Handbook of carbon, graphite, diamond, and fullerenes : properties, processing, and applications. Park Ridge, N.J.: Noyes Publications. ISBN   0-8155-1739-4. OCLC   49708274.
  10. Bera, J. K.; Dunbar, K. R. (2002). "Chain Compounds Based on Transition Metal Backbones: New Life for an Old Topic". Angew. Chem. Int. Ed. 41 (23): 4453–4457. doi:10.1002/1521-3773(20021202)41:23<4453::AID-ANIE4453>3.0.CO;2-1. PMID   12458505.
  11. Huq, Fazlul; Skapski, Andrzej C. (1974). "Refinement of the crystal structure of acetylacetonatodicarbonylrhodium(I)". J. Cryst. Mol. Struct. 4 (6): 411–418. doi:10.1007/BF01220097. S2CID   96977904.
  12. Smith, Michael B.; March, Jerry (2007), Advanced Organic Chemistry: Reactions, Mechanisms, and Structure (6th ed.), New York: Wiley-Interscience, p. 114, ISBN   978-0-471-72091-1
  13. Lewis M, Bagwill C, Hardebeck L, Wireduaah S (2016). "Modern Computational Approaches to Understanding Interactions of Aromatics". In Johnson DW, Hof F (eds.). Aromatic Interactions: Frontiers in Knowledge and Application. England: Royal Society of Chemistry. pp. 1–17. ISBN   978-1-78262-662-6.
  14. McGaughey GB, Gagné M, Rappé AK (June 1998). "pi-Stacking interactions. Alive and well in proteins". The Journal of Biological Chemistry. 273 (25): 15458–63. doi: 10.1074/jbc.273.25.15458 . PMID   9624131.
  15. Martinez CR, Iverson BL (2012). "Rethinking the term "pi-stacking"". Chemical Science. 3 (7): 2191. doi:10.1039/c2sc20045g. hdl: 2152/41033 . ISSN   2041-6520. S2CID   95789541.
  16. Ashton PR, Goodnow TT, Kaifer AE, Reddington MV, Slawin AM, Spencer N, et al. (1989). "A [2] Catenane Made to Order". J. Angew. Chem. Int. Ed. 28 (10): 1396–1399. doi:10.1002/anie.198913961.