Spin-forbidden reactions

Last updated

In chemistry, the selection rule (also known as the transition rule) formally restricts certain reactions, known as spin-forbidden reactions, from occurring due to a required change between two differing quantum states. When a reactant exists in one spin state and the product exists in a different spin state, the corresponding reaction will have an increased activation energy when compared to a similar reaction in which the spin states of the reactant and product are isomorphic. As a result of this increased activation energy, a decreased rate of reaction is observed.

Contents

Case of some cobalt carbonyls

Singlet and Triplet Cobalt Complexes. The ground state of Tp Co(CO) has two unpaired electrons. CobaltSingletTriplet.gif
Singlet and Triplet Cobalt Complexes. The ground state of Tp Co(CO) has two unpaired electrons.

Singlet and triplet states can occur within organometallic complexes as well, such as Tpi-Pr,MeCo(CO)2 and Tpi-Pr,MeCo(CO), respectively.

Singlet and Triplet Molecular Orbitals. SomoLumoHomo2.gif
Singlet and Triplet Molecular Orbitals.

Changing spin states

When a reaction converts a metal from a singlet to triplet state (or vice versa):

  1. The energy of the two spin states must be nearly equal, as dictated by temperature,
  2. A mechanism is required to change spin states.

Strong spin-orbital coupling can satisfy the 2nd condition. Parameter 1, however, can lead to very slow reactions due to large disparities between the metal complex's potential energy surfaces, which only cross at high energy leading to a substantial activation barrier. [2]

Spin-forbidden reactions formally fall into the category of electronically non-adiabatic reactions. [3] In general, potential energy surfaces fall into either the adiabatic and diabatic classification. Potential Energy Surfaces that are adiabatic rely on the use of the full electronic Hamiltonian, which includes the spin-orbit term. Those that are diabatic are likewise derived by solving the eigenvalues of the Schrödinger equation, but in this case one or more terms are omitted. [4]

Non-adiabatic transition

Potential energy surfaces for spin-forbidden reactions of both diabatic and adiabatic variety. The difference between the two adiabatic surfaces is 2H12, where H12 = <Ps1|Hsoc|Ps2>. Adiabatic and Diabatic Curves.gif
Potential energy surfaces for spin-forbidden reactions of both diabatic and adiabatic variety. The difference between the two adiabatic surfaces is 2H12, where H12 = <Ψ1|Hsoc2>.

Once a minimum energy crossing point is reached and parameter 1 above is satisfied, the system needs to hop from one diabatic surface to the other, as stated above by parameter 2. At a given energy (E), the rate coefficient [k(E)] of a spin-forbidden reaction can be calculated using the density of rovibrational states of the reactant [ρ(E)] and the effective integrated density of states in the crossing seam between the two surfaces [Ner(E)].

where

The probability of hopping (psh) is calculated from Landau-Zener theory giving

where

in which the spin-orbit coupling derived off the diagonal Hamiltonian matrix element between two electronic states (H12), the relative slope of the two surfaces at the crossing seam [F(Δ)], the reduced mass of the system through its movement along the hopping coordinate (μ), and the kinetic energy of the system passing through the crossing point (E) are used.

It is useful to note that when Eh < Ec (when below the minimum energy crossing point) the probability of hopping between spin states is null. [5]

An example of a spin-forbidden reaction

Energy diagrams of Fe(CO)n - Fe(CO)n+1 The above image shows the energy diagrams through transitions as CO is added to and Fe-centered organometallic molecule. The first three complexes have S=2 while the final one has S=0. FeCoSpin.gif
Energy diagrams of Fe(CO)n → Fe(CO)n+1 The above image shows the energy diagrams through transitions as CO is added to and Fe-centered organometallic molecule. The first three complexes have S=2 while the final one has S=0.

One example showing the slowing effect of spin-forbidden reaction takes place when Fe(CO)x is placed under CO pressure. Transitions from x = 2,3,4 to x = 3,4,5 demonstrate a slowing rate when the reactant is in a triplet ground state but the product is in a singlet ground state. In the case of Fe(CO)x, when x = 2, 3, 4 the iron exists as a triplet in ground state; when x = 5, the iron exists as a singlet in ground state. The kinetics of said system can be represented by:

where:

,
, and
.

The rate constants above are notably temperature independent, tested at 55 °C, 21 °C, and 10 °C, indicating, according to the authors, that the observed 500 fold reduction in rate occurs due to the spin-forbidden nature of the latter equation and not due to the kinetics of adding an additional ligand. [7]

Application to catalysis

Ligand association and dissociation from metal centers involve a fundamental change in the coordination sphere of that metal. This change in certain reactions also requires a change in spin state to occur, which can retard the rate of ligand association. Oxidative addition and reductive elimination can be thought of in a similar manner to ligand association and dissociation, respectively. [8] Mathematically, the rate ligand dissociation can increase when the reaction proceeds from one spin state to another, although it must be noted that this effect is often small in comparison with other factors such as sterics around the metal center. [9] [10]

C-H activation

Insertion into C-H bonds, known as C-H activation, is an integral first step in C-H functionalization. [11] For some metal complexes with identical ligands, C-H activation is rapid when one metal is used and slow when other metals are used, often first row transition metals, due to the spin allowed nature of the former case and the spin-forbidden nature of the latter case. The difference in rates of C-H activation of methane for CoCp(CO), RhCp(CO), and IrCp(CO) readily demonstrate this property. CoCp(CO), the starting material in a C-H activation, exists in a triplet spin state while RhCp(CO) exists in a singlet state, with the triplet state only 5.9 kcal/mol away. IrCp(CO) is unique among these complexes in that its starting state is essentially degenerate between the triplet and singlet states. The given product of C-H insertion, CpMH(CO)(CH3), where M = Co, Rh, Ir, is in a singlet state meaning that the C-H activation with CoCp(CO) must reach the minimum energy crossing point for the reactant and product's potential energy surfaces, thus requiring relatively high energies to proceed. [12]

CH Activation Singlet-Triplet.gif

Oxidative addition into silicon-hydrogen bonds

Through the use of photolysis, the rate of oxidative addition into silicon-hydrogen bonds has been shown to increase when the starting material is excited to the correct spin-state. CpRe(CO)2 and CpMn(CO)2 were subjected to photolysis to change their spin states from triplet to singlet and vice versa, respectively, such that an oxidative addition to Et3Si-H could occur at a greatly accelerated rate. [13]

Photolysis to achieve Si-H Insertion.gif

Oxidation chemistry

Reactions of manganese-oxo complexes with alkenes. Depending on the spin state of the starting material, triplet or quintet in this case, route A, B, or C can be followed, yielding potentially different products. The reactions also proceed differently for different R substitutes, where one is always alkyl and the second one is alkyl in the case of route A, aryl, alkenyl, or alkynyl in the case of route B, or alkyl, aryl, alkenyl, or alkynyl for route C. For routes B and C, the alkyl R is internal for the radical case. Manganese Oxo Reactivity.gif
Reactions of manganese-oxo complexes with alkenes. Depending on the spin state of the starting material, triplet or quintet in this case, route A, B, or C can be followed, yielding potentially different products. The reactions also proceed differently for different R substitutes, where one is always alkyl and the second one is alkyl in the case of route A, aryl, alkenyl, or alkynyl in the case of route B, or alkyl, aryl, alkenyl, or alkynyl for route C. For routes B and C, the alkyl R is internal for the radical case.

Metal-oxo species, due to their small spatial extent of metal-centered d orbitals leading to weak bonding, often have similar energies for both the low spin () and high spin configuration (). [14] This similarity in energy between the low- and high spin configurations of oxo-species lends itself to the study of spin-forbidden reactions, such as Mn(salen)-catalyzed epoxidation. The Mn(salen)-oxo species can exist in either a triplet or quintet state. While the product of the quintet lies at a lower energy, both the triplet and quintet products can be observed. [15]

Energy of triplet and quintet states of Mn-Oxo species.gif

Related Research Articles

In a chemical reaction, chemical equilibrium is the state in which both the reactants and products are present in concentrations which have no further tendency to change with time, so that there is no observable change in the properties of the system. This state results when the forward reaction proceeds at the same rate as the reverse reaction. The reaction rates of the forward and backward reactions are generally not zero, but they are equal. Thus, there are no net changes in the concentrations of the reactants and products. Such a state is known as dynamic equilibrium.

<span class="mw-page-title-main">Chemical reaction</span> Process that results in the interconversion of chemical species

A chemical reaction is a process that leads to the chemical transformation of one set of chemical substances to another. Classically, chemical reactions encompass changes that only involve the positions of electrons in the forming and breaking of chemical bonds between atoms, with no change to the nuclei, and can often be described by a chemical equation. Nuclear chemistry is a sub-discipline of chemistry that involves the chemical reactions of unstable and radioactive elements where both electronic and nuclear changes can occur.

<span class="mw-page-title-main">Inorganic chemistry</span> Field of chemistry

Inorganic chemistry deals with synthesis and behavior of inorganic and organometallic compounds. This field covers chemical compounds that are not carbon-based, which are the subjects of organic chemistry. The distinction between the two disciplines is far from absolute, as there is much overlap in the subdiscipline of organometallic chemistry. It has applications in every aspect of the chemical industry, including catalysis, materials science, pigments, surfactants, coatings, medications, fuels, and agriculture.

<span class="mw-page-title-main">Metallocene</span>

A metallocene is a compound typically consisting of two cyclopentadienyl anions (C
5
H
5
, abbreviated Cp) bound to a metal center (M) in the oxidation state II, with the resulting general formula (C5H5)2M. Closely related to the metallocenes are the metallocene derivatives, e.g. titanocene dichloride, vanadocene dichloride. Certain metallocenes and their derivatives exhibit catalytic properties, although metallocenes are rarely used industrially. Cationic group 4 metallocene derivatives related to [Cp2ZrCH3]+ catalyze olefin polymerization.

<span class="mw-page-title-main">Photochemistry</span> Sub-discipline of chemistry

Photochemistry is the branch of chemistry concerned with the chemical effects of light. Generally, this term is used to describe a chemical reaction caused by absorption of ultraviolet, visible light (400–750 nm) or infrared radiation (750–2500 nm).

<span class="mw-page-title-main">Intersystem crossing</span>

Intersystem crossing (ISC) is an isoenergetic radiationless process involving a transition between the two electronic states with different spin multiplicity.

In organic chemistry, a carbene is a molecule containing a neutral carbon atom with a valence of two and two unshared valence electrons. The general formula is R−:C−R' or R=C: where the R represents substituents or hydrogen atoms.

The equilibrium constant of a chemical reaction is the value of its reaction quotient at chemical equilibrium, a state approached by a dynamic chemical system after sufficient time has elapsed at which its composition has no measurable tendency towards further change. For a given set of reaction conditions, the equilibrium constant is independent of the initial analytical concentrations of the reactant and product species in the mixture. Thus, given the initial composition of a system, known equilibrium constant values can be used to determine the composition of the system at equilibrium. However, reaction parameters like temperature, solvent, and ionic strength may all influence the value of the equilibrium constant.

<span class="mw-page-title-main">Singlet oxygen</span> Oxygen with all of its electrons spin paired

Singlet oxygen, systematically named dioxygen(singlet) and dioxidene, is a gaseous inorganic chemical with the formula O=O (also written as 1
[O
2
]
or 1
O
2
), which is in a quantum state where all electrons are spin paired. It is kinetically unstable at ambient temperature, but the rate of decay is slow.

<span class="mw-page-title-main">Photosensitizer</span> Type of molecule reacting to light

Photosensitizers are light absorbers that alters the course of a photochemical reaction. They usually are catalysts. They can function by many mechanisms, sometimes they donate an electron to the substrate, sometimes they abstract a hydrogen atom from the substrate. At the end of this process, the photosensitizer returns to its ground state, where it remains chemically intact, poised to absorb more light. One branch of chemistry which frequently utilizes photosensitizers is polymer chemistry, using photosensitizers in reactions such as photopolymerization, photocrosslinking, and photodegradation. Photosensitizers are also used to generate prolonged excited electronic states in organic molecules with uses in photocatalysis, photon upconversion and photodynamic therapy. Generally, photosensitizers absorb electromagnetic radiation consisting of infrared radiation, visible light radiation, and ultraviolet radiation and transfer absorbed energy into neighboring molecules. This absorption of light is made possible by photosensitizers' large de-localized π-systems, which lowers the energy of HOMO and LUMO orbitals to promote photoexcitation. While many photosensitizers are organic or organometallic compounds, there are also examples of using semiconductor quantum dots as photosensitizers.

In theoretical chemistry, Marcus theory is a theory originally developed by Rudolph A. Marcus, starting in 1956, to explain the rates of electron transfer reactions – the rate at which an electron can move or jump from one chemical species (called the electron donor) to another (called the electron acceptor). It was originally formulated to address outer sphere electron transfer reactions, in which the two chemical species only change in their charge with an electron jumping (e.g. the oxidation of an ion like Fe2+/Fe3+), but do not undergo large structural changes. It was extended to include inner sphere electron transfer contributions, in which a change of distances or geometry in the solvation or coordination shells of the two chemical species is taken into account (the Fe-O distances in Fe(H2O)2+ and Fe(H2O)3+ are different).

Spin chemistry is a sub-field of chemistry and physics, positioned at the intersection of chemical kinetics, photochemistry, magnetic resonance and free radical chemistry, that deals with magnetic and spin effects in chemical reactions. Spin chemistry concerns phenomena such as chemically induced dynamic nuclear polarization (CIDNP), chemically induced electron polarization (CIDEP), magnetic isotope effects in chemical reactions, and it is hypothesized to be key in the underlying mechanism for avian magnetoreception and consciousness.

The 18-electron rule is a chemical rule of thumb used primarily for predicting and rationalizing formulas for stable transition metal complexes, especially organometallic compounds. The rule is based on the fact that the valence orbitals in the electron configuration of transition metals consist of five (n−1)d orbitals, one ns orbital, and three np orbitals, where n is the principal quantum number. These orbitals can collectively accommodate 18 electrons as either bonding or non-bonding electron pairs. This means that the combination of these nine atomic orbitals with ligand orbitals creates nine molecular orbitals that are either metal-ligand bonding or non-bonding. When a metal complex has 18 valence electrons, it is said to have achieved the same electron configuration as the noble gas in the period, lending stability to the complex. Transition metal complexes that deviate from the rule are often interesting or useful because they tend to be more reactive. The rule is not helpful for complexes of metals that are not transition metals. The rule was first proposed by American chemist Irving Langmuir in 1921.

In chemistry and molecular physics, fluxionalmolecules are molecules that undergo dynamics such that some or all of their atoms interchange between symmetry-equivalent positions. Because virtually all molecules are fluxional in some respects, e.g. bond rotations in most organic compounds, the term fluxional depends on the context and the method used to assess the dynamics. Often, a molecule is considered fluxional if its spectroscopic signature exhibits line-broadening due to chemical exchange. In some cases, where the rates are slow, fluxionality is not detected spectroscopically, but by isotopic labeling and other methods.

<span class="mw-page-title-main">Transition state theory</span> Theory describing the reaction rates of elementary chemical reactions

In chemistry, transition state theory (TST) explains the reaction rates of elementary chemical reactions. The theory assumes a special type of chemical equilibrium (quasi-equilibrium) between reactants and activated transition state complexes.

<span class="mw-page-title-main">Radical (chemistry)</span> Atom, molecule, or ion that has an unpaired valence electron; typically highly reactive

In chemistry, a radical, also known as a free radical, is an atom, molecule, or ion that has at least one unpaired valence electron. With some exceptions, these unpaired electrons make radicals highly chemically reactive. Many radicals spontaneously dimerize. Most organic radicals have short lifetimes.

Equilibrium chemistry is concerned with systems in chemical equilibrium. The unifying principle is that the free energy of a system at equilibrium is the minimum possible, so that the slope of the free energy with respect to the reaction coordinate is zero. This principle, applied to mixtures at equilibrium provides a definition of an equilibrium constant. Applications include acid–base, host–guest, metal–complex, solubility, partition, chromatography and redox equilibria.

Methylene is an organic compound with the chemical formula CH
2
. It is a colourless gas that fluoresces in the mid-infrared range, and only persists in dilution, or as an adduct.

<span class="mw-page-title-main">Cyclopentadienyliron dicarbonyl dimer</span> Chemical compound

Cyclopentadienyliron dicarbonyl dimer is an organometallic compound with the formula [(η5-C5H5)Fe(CO)2]2, often abbreviated to Cp2Fe2(CO)4, [CpFe(CO)2]2 or even Fp2, with the colloquial name "fip dimer". It is a dark reddish-purple crystalline solid, which is readily soluble in moderately polar organic solvents such as chloroform and pyridine, but less soluble in carbon tetrachloride and carbon disulfide. Cp2Fe2(CO)4 is insoluble in but stable toward water. Cp2Fe2(CO)4 is reasonably stable to storage under air and serves as a convenient starting material for accessing other Fp (CpFe(CO)2) derivatives (described below).

A metal-centered cycloaddition is a subtype of the more general class of cycloaddition reactions. In such reactions "two or more unsaturated molecules unite directly to form a ring", incorporating a metal bonded to one or more of the molecules. Cycloadditions involving metal centers are a staple of organic and organometallic chemistry, and are involved in many industrially-valuable synthetic processes.

References

  1. Theopold, Klaus H (1995). "Can Spin State Change Slow Organometallic Reactions". Journal of the American Chemical Society. 117 (47): 11745–8. doi:10.1021/ja00152a015.
  2. Cundari, Thomas (2001). Computational Organometallic Chemistry . Marcel Dekker, Inc. pp.  293. ISBN   9780824704780.
  3. Cundari, Thomas (2001). Computational Organometallic Chemistry . Marcel Dekker, Inc. pp.  294. ISBN   9780824704780.
  4. Harvey, Jeremy (2006). "Understanding the Kinetics of Spin-Forbidden Chemical Reactions". Physical Chemistry Chemical Physics. 9 (3): 331–2. doi:10.1039/b614390c. PMID   17199148.
  5. Harvey, Jeremy (2006). "Understanding the Kinetics of Spin-Forbidden Chemical Reactions". Physical Chemistry Chemical Physics. 9 (3): 332–3. doi:10.1039/b614390c. PMID   17199148.
  6. Kirchner, Karl (2010). "Reactivity of coordinatively unsaturated iron complexes towards carbon monoxide: to bind or not to bind?". Dalton Transactions. 40 (18): 4778–4792. doi:10.1039/c0dt01636e. PMID   21380474.
  7. Weitz, Eric (1986). "The wavelength dependence of excimer laser photolysis of Fe(CO)5 in the gas phase. Transient infrared spectroscopy and kinetics of the FeCOx (x=4,3,2) photofragments". The Journal of Chemical Physics. 84 (4): 1977–1986. Bibcode:1986JChPh..85.1977S. doi:10.1063/1.451141.
  8. Cundari, Thomas (2001). Computational Organometallic Chemistry . Marcel Dekker, Inc. pp.  299–303. ISBN   9780824704780.
  9. Poli, Rinaldo (1995). "Dissociative phosphine exchange for cyclopentadienylmolybdenum(III) systems. Bridging the gap between Werner-like coordination chemistry and low-valent organometallic chemistry" (PDF). Inorganica Chimica Acta. 240 (1–2): 355–66. doi:10.1016/0020-1693(95)04554-6.
  10. Poli, Rinaldo (1996). "Molybdenum Open-Shell Organometallics. Spin State Changes in Pairing Energy Effects" (PDF). Journal of the American Chemical Society. 30 (12): 494–501. doi:10.1021/ar960280g.
  11. Organometallic C–H Bond Activation: An Introduction Alan S. Goldman and Karen I. Goldberg ACS Symposium Series 885, Activation and Functionalization of C–H Bonds, 2004, 1–43
  12. Siegbahn, Per (1996). "Comparison of the C-H Activation of Methane by M(C5H5)(CO) for M=Cobalt, Rhodium, and Iridium". Journal of the American Chemical Society. 118 (6): 1487–96. doi:10.1021/ja952338c.
  13. Harris, Charles (1999). "Ultrafast Infrared Studies of Bond Activation in Organometallic Complexes". Acc. Chem. Res. 32 (7): 551–60. doi:10.1021/ar970133y.
  14. Cudari, Thomas (2001). Computational Organometallic Chemistry. Marcel Dekker Inc. pp. 301–2.
  15. Linde, C.; Åkermark, B.; Norrby, P.-O.; Svensson, M. (1999). "Timing Is Critical: Effect of Spin Changes on the Diastereoslectivity in Mn(salen)-Catalyzed Epoxidation". Journal of the American Chemical Society. 121 (21): 5083–4. doi:10.1021/ja9809915.