Singlet oxygen

Singlet oxygen
Names
IUPAC name
Singlet oxygen
Identifiers
3D model (JSmol)
ChEBI
491
Properties
O2
Molar mass 32.00 g·mol−1
Reacts
Except where otherwise noted, data are given for materials in their standard state (at 25 °C [77 °F], 100 kPa).
Infobox references

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, however the rate of decay is slow.

The lowest excited state of the diatomic oxygen molecule is the singlet state. It is a gas with physical properties differing only subtly from those of the more prevalent triplet ground state of O2. In terms of its chemical reactivity, however, singlet oxygen is far more reactive toward organic compounds. It is responsible for the photodegradation of many materials but can be put to constructive use in preparative organic chemistry and photodynamic therapy. Trace amounts of singlet oxygen are found in the upper atmosphere and also in polluted urban atmospheres where it contributes to the formation of lung-damaging nitrogen dioxide.[1]:355–68 It often appears and coexists confounded in environments that also generate ozone, such as pine forests with photodegradation of turpentine.

The terms 'singlet oxygen' and 'triplet oxygen' derive from each form's number of electron spins. The singlet has only one possible arrangement of electron spins with a total quantum spin of 0, while the triplet has three possible arrangements of electron spins with a total quantum spin of 1, corresponding to three degenerate states.

In spectroscopic notation, the singlet and triplet forms of O2 are labeled 1Δg and 3Σ
g
, respectively.[2][3][4]

Electronic structure

Singlet oxygen refers to one of two singlet electronic excited states. The two singlet states are denoted 1Σ+
g
and 1Δg (the preceding superscripted "1" indicates it as a singlet state). The singlet states of oxygen are 158 and 95 kilojoules per mole higher in energy than the triplet ground state of oxygen. Under most common laboratory conditions, the higher energy 1Σ+
g
singlet state rapidly converts to the more stable, lower energy 1Δg singlet state;[2] it is this, the more stable of the two excited states, the one with its electrons remaining in separate degenerate orbital but no longer with like spin, that is referred to by the title term, singlet oxygen, commonly abbreviated 1O2, to distinguish it from the triplet ground state molecule, 3O2.[2][3]

Molecular orbital theory predicts the electronic ground state denoted by the molecular term symbol 3Σ
g
and two low-lying excited singlet states, with molecular term symbols 1Δg and 1Σ+
g
. These three electronic states differ only in the spin and the occupancy of oxygen's two antibonding πg-orbitals, which are degenerate (equal in energy). These two orbitals are classified as antibonding and are of higher energy. Following Hund's first rule, in the ground state, these electrons are unpaired and have like (same) spin. This open-shell triplet ground state of molecular oxygen differs from most stable diatomic molecules, which have singlet (1Σ+
g
) ground states.[5]

Two less stable, higher energy excited states are readily accessible from this ground state, again in accordance with Hund's first rule;[6] the first moves one of the high energy unpaired ground state electrons from one degenerate orbital to the other, where it "flips" and pairs the other, and creates a new state, a singlet state referred to as the 1Δg state (a term symbol, where the preceding superscripted "1" indicates it as a singlet state).[2][3] Alternatively, both electrons can remain in their degenerate ground state orbitals, but the spin of one can "flip" so that it is now opposite to the second (i.e., it is still in a separate degenerate orbital, but no longer of like spin); this also creates a new state, a singlet state referred to as the 1Σ+
g
state.[2][3] The ground and first two singlet excited states of oxygen can be described by the simple scheme in the figure below.[7][8]

MO diagram, triplet ground state and two singlet excited states of molecular dioxygen. Shown are three electronic configurations of the molecular orbitals (MOs) of molecular oxygen, O2. From left to right, the diagrams are for: 1Δg singlet oxygen (first excited state), 1Σ+
g
singlet oxygen (second excited state), and 3Σ
g
triplet oxygen (ground state). The lowest energy 1s molecular orbitals are uniformly filled in all three and are omitted for simplicity. The broad horizontal lines labelled π and π* each represent two molecular orbitals (for filling by up to 4 electrons in total). The three states only differ in the occupancy and spin states of electrons in the two degenerate π* antibonding orbitals.

The 1Δg singlet state is 7882.4 cm−1 above the triplet 3Σ
g
ground state.,[3][9] which in other units corresponds to 94.29 kJ/mol or 0.9773 eV. The 1Σ+
g
singlet is 13 120.9 cm−1[3][9] (157.0 kJ/mol or 1.6268 eV) above the ground state.

Radiative transitions between the three low-lying electronic states of oxygen are formally forbidden as electric dipole processes.[10] The two singlet-triplet transitions are forbidden both because of the spin selection rule ΔS = 0 and because of the parity rule that g-g transitions are forbidden.[11]

The lower, O2(1Δg) state is commonly referred to as singlet oxygen. The energy difference of 94.3 kJ/mol between ground state and singlet oxygen corresponds to a forbidden singlet-triplet transition in the near-infrared at ~1270 nm.[12] As a consequence, singlet oxygen in the gas phase is extremely long lived (72 minutes),[13] although interaction with solvents reduces the lifetime to microseconds or even nanoseconds.[14]

The higher 1Σ+
g
state is very short lived. In the gas phase, it relaxes primarily to the ground state triplet with a mean lifetime of 11.8 s.[10] However in solvents such as CS2 and CCl4, it relaxes to the lower singlet 1Δg in milliseconds due to nonradiative decay channels.[10]

Paramagnetism due to orbital angular momentum

Both singlet oxygen states have no unpaired electrons and therefore no net electron spin. The 1Δg is however paramagnetic as shown by the observation of an electron paramagnetic resonance (EPR) spectrum.[15][16][17] The paramagnetism is due to a net orbital (and not spin) electronic angular momentum. In a magnetic field the degeneracy of the π* level is split into two levels corresponding to molecular orbitals with angular momenta +1ħ and −1ħ around the molecular axis. In the 1Δg state one of these orbitals is doubly occupied and the other empty, so that transitions are possible between the two.

Production

Various methods for the production of singlet oxygen exist. Irradiation of oxygen gas in the presence of an organic dye as a sensitizer, such as rose bengal, methylene blue, or porphyrins—a photochemical method—results in its production.[18][9] Singlet oxygen can also be in non-photochemical, preparative chemical procedures. One chemical method involves the decomposition of triethylsilyl hydrotrioxide generated in situ from triethylsilane and ozone.[19]

(C2H5)3SiH + O3 → (C2H5)3SiOOOH → (C2H5)3SiOH + O2(1Δg)

Another method uses the aqueous reaction of hydrogen peroxide with sodium hypochlorite:[18]

H2O2 + NaOCl → O2(1Δg) + NaCl + H2O

A third method liberates singlet oxygen via phosphite ozonides, which are, in turn, generated in situ.[20] Phosphite ozonides will decompose to give singlet oxygen:[21]

(RO)3P + O3 → (RO)3PO3
(RO)3PO3 → (RO)3PO + O2(1Δg)

An advantage of this method is that it is amenable to non-aqueous conditions.[21]

Reactions

Because of differences in their electron shells, singlet and triplet oxygen differ in their chemical properties; singlet oxygen is highly reactive.[22] The lifetime of singlet oxygen depends on the medium. In normal organic solvents, the lifetime is only a few microseconds whereas in solvents lacking C-H bonds, the lifetime can be as long as seconds.[21]

Organic chemistry

Singlet oxygen-based oxidation of citronellol. This is a net, but not a true ene reaction. Abbreviations, step 1: H2O2, hydrogen peroxide; Na2MoO4 (catalyst), sodium molybdate. Step 2: Na2SO3 (reducing agent), sodium sulfite.

Unlike ground state oxygen, singlet oxygen participates in Diels–Alder [4+2]- and [2+2]-cycloaddition reactions and formal concerted ene reactions.[21] It oxidizes thioethers to sulfoxides and organometallic complexes.[23][24] With some substrates 1,2-dioxetanes are formed; cyclic dienes such as 1,3-cyclohexadiene form [4+2] cycloaddition adducts.[25]

In singlet oxygen reactions with alkenic allyl groups, e.g., citronella, shown, by abstraction of the allylic proton, in an ene-like reaction, yielding the allyl hydroperoxide, R–O–OH (R = alkyl), which can then be reduced to the corresponding allylic alcohol.[21][26][27][28]

In reactions with water trioxidane, an unusual molecule with three consecutive linked oxygen atoms, is formed.

Biochemistry

In photosynthesis, singlet oxygen can be produced from the light-harvesting chlorophyll molecules. One of the roles of carotenoids in photosynthetic systems is to prevent damage caused by produced singlet oxygen by either removing excess light energy from chlorophyll molecules or quenching the singlet oxygen molecules directly.

In mammalian biology, singlet oxygen is one of the reactive oxygen species, which is linked to oxidation of LDL cholesterol and resultant cardiovascular effects. Polyphenol antioxidants can scavenge and reduce concentrations of reactive oxygen species and may prevent such deleterious oxidative effects.[29]

Ingestion of pigments capable of producing singlet oxygen with activation by light can produce severe photosensitivity of skin (see phototoxicity, photosensitivity in humans, photodermatitis, phytophotodermatitis). This is especially a concern in herbivorous animals (see Photosensitivity in animals).

Singlet oxygen is the active species in photodynamic therapy.

Analytical and physical chemistry

Red glow of singlet oxygen passing into triplet state.

Direct detection of singlet oxygen is possible using sensitive laser spectroscopy [30] or through its extremely weak phosphorescence at 1270 nm, which is not visible.[31] However, at high singlet oxygen concentrations, the fluorescence of the singlet oxygen "dimol" species—simultaneous emission from two singlet oxygen molecules upon collision—can be observed as a red glow at 634 nm.[32]

References

  1. Wayne RP (1969). Pitts JN, Hammond GS, Noyes WA, eds. "Singlet Molecular Oxygen". Advances in Photochemistry. 7: 311–71. doi:10.1002/9780470133378.ch4.
  2. 1 2 3 4 5 Klán P, Wirz J (2009). Photochemistry of Organic Compounds: From Concepts to Practice (Repr. 2010 ed.). Chichester, West Sussex, U.K.: Wiley. ISBN 1405190884.
  3. 1 2 3 4 5 6 Atkins P, de Paula J (2006). Atkins' Physical Chemistry (8th ed.). W.H.Freeman. pp. 482–3. ISBN 0-7167-8759-8.
  4. Hill C. "Molecular Term Symbols" (PDF). Retrieved 10 October 2016.
  5. Levine IN (1991). Quantum Chemistry (4th ed.). Prentice-Hall. p. 383. ISBN 0-205-12770-3.
  6. Frimer AA (1985). Singlet Oxygen: Volume I, Physical-Chemical Aspects. Boca Raton, Fla.: CRC Press. pp. 4–7. ISBN 9780849364396.
  7. For triplet ground state on right side of diagram, see C.E.Housecroft and A.G.Sharpe Inorganic Chemistry, 2nd ed. (Pearson Prentice-Hall 2005), p.35 ISBN 0130-39913-2
  8. For changes in singlet states on left and in centre, see F. Albert Cotton and Geoffrey Wilkinson. Advanced Inorganic Chemistry, 5th ed. (John Wiley 1988), p.452 ISBN 0-471-84997-9
  9. 1 2 3 Schweitzer C, Schmidt R (May 2003). "Physical Mechanisms of Generation and Deactivation of Singlet Oxygen". Chemical Reviews. 103 (5): 1685–757. doi:10.1021/cr010371d. PMID 12744692.
  10. 1 2 3 Weldon, Dean; Poulsen, Tina D.; Mikkelsen, Kurt V.; Ogilby, Peter R. (1999). "Singlet Sigma: The "Other" Singlet Oxygen in Solution". Photochemistry and Photobiology. 70 (4): 369–379. Retrieved 26 March 2018.
  11. Thomas Engel; Philip Reid (2006). Physical Chemistry. PEARSON Benjamin Cummings. p. 580–. ISBN 0-8053-3842-X.
  12. Guy P. Brasseur; Susan Solomon (January 15, 2006). Aeronomy of the Middle Atmosphere: Chemistry and Physics of the Stratosphere and Mesosphere. Springer Science & Business Media. pp. 220–. ISBN 978-1-4020-3824-2.
  13. Chemistry of Singlet Oxygen O2(a1DELTAg) in the Upper Atmosphere John M. Plane, Research Report, Air Force Research Laboratory, European Office of Aerospace Research and Development (2012)
  14. Wilkinson F, Helman WP, Ross AB (1995). "Rate constants for the decay and reactions of the lowest electronically excited singlet state of molecular oxygen in solution. An expanded and revised compilation". J. Phys. Chem. Ref. Data. 24 (2): 663–677. Bibcode:1995JPCRD..24..663W. doi:10.1063/1.555965.
  15. Hasegawa K, Yamada K, Sasase R, Miyazaki R, Kikuchi A, Yagi M (2008). "Direct measurements of absolute concentration and lifetime of singlet oxygen in the gas phase by electron paramagnetic resonance". Chemical Physics Letters. 457: 312–314. Bibcode:2008CPL...457..312H. doi:10.1016/j.cplett.2008.04.031.
  16. Ruzzi M, Sartori E, Moscatelli A, Khudyakov IV, Turro NJ (June 2013). "Time-resolved EPR study of singlet oxygen in the gas phase". The Journal of Physical Chemistry A. 117 (25): 5232–40. Bibcode:2013JPCA..117.5232R. doi:10.1021/jp403648d. PMID 23768193.
  17. Falick AM, et al. (1965). "Paramagnetic resonance spectrum of the 1?g oxygen molecule". J. Chem. Phys. 42: 1837–1838. Bibcode:1965JChPh..42.1837F. doi:10.1063/1.1696199.
  18. 1 2 Greer A (2006). "Christopher Spencer Foote's Discovery of the Role of Singlet Oxygen [1O2 (1Δg)] in Photosensitized Oxidation Reactions". Acc. Chem. Res. 39 (11): 797–804. doi:10.1021/ar050191g. PMID 17115719.
  19. Corey EJ, Mehrotra MM, Khan AU (April 1986). "Generation of 1Δg from triethylsilane and ozone". Journal of the American Chemical Society. 108 (9): 2472–3. doi:10.1021/ja00269a070. PMID 22175617.
  20. Housecroft CE, Sharpe AG (2008). "Chapter 15: The group 16 elements". Inorganic Chemistry (3rd ed.). Pearson. p. 438f. ISBN 9780131755536.
  21. 1 2 3 4 5 Wasserman HH, DeSimone RW, Chia KR, Banwell MG (2001). "Singlet Oxygen". e-EROS Encyclopedia of Reagents for Organic Synthesis. John Wiley & Sons. doi:10.1002/047084289X.rs035.
  22. Ho RY, Liebman JF, Valentine JS (1995). "Overview of the Energetics and Reactivity of Oxygen". In Foote CS. Active Oxygen in Chemistry. London: Blackie Academic & Professional. pp. 1–23. doi:10.1007/978-94-007-0874-7_1. ISBN 978-0-7514-0371-8.
  23. Clennan EL, Pace A (2005). "Advances in singlet oxygen chemistry". Tetrahedron. 61 (28): 6665–6691. doi:10.1016/j.tet.2005.04.017.
  24. Ogilby PR (August 2010). "Singlet oxygen: there is indeed something new under the sun". Chemical Society Reviews. 39 (8): 3181–209. doi:10.1039/b926014p. PMID 20571680.
  25. Carey FA, Sundberg RJ (1985). Structure and mechanisms (2 ed.). New York: Plenum Press. ISBN 0306411989.
  26. Stephenson LM, Grdina MJ, Orfanopoulos M (November 1980). "Mechanism of the ene reaction between singlet oxygen and olefins". Accounts of Chemical Research. 13 (11): 419–425. doi:10.1021/ar50155a006.
  27. This reaction is not a true ene reaction, because it is not concerted; singlet oxygen forms an "epoxide oxide" exciplex, which then abstracts the hydrogen. See Alberti et al, op. cit.
  28. Alsters PL, Jary W, Nardello-Rataj V, Jean-Marie A (2009). "Dark Singlet Oxygenation of β-Citronellol: A Key Step in the Manufacture of Rose Oxide". Organic Process Research & Development. 14: 259–262. doi:10.1021/op900076g.
  29. Karp G, van der Geer P (2004). Cell and molecular biology: concepts and experiments (4th ed., Wiley International ed.). New York: J. Wiley & Sons. p. 223. ISBN 978-0471656654.
  30. Földes T, Čermák P, Macko M, Veis P, Macko P (January 2009). "Cavity ring-down spectroscopy of singlet oxygen generated in microwave plasma". Chemical Physics Letters. 467 (4–6): 233–236. Bibcode:2009CPL...467..233F. doi:10.1016/j.cplett.2008.11.040.
  31. Nosaka Y, Daimon T, Nosaka, AY, Murakami Y (2004). "Singlet oxygen formation in photocatalytic TiO₂ aqueous suspension". Phys. Chem. Chem. Phys. 6 (11): 2917–2918. Bibcode:2004PCCP....6.2917N. doi:10.1039/B405084C.
  32. Mulliken RS (1928). "Interpretation of the atmospheric oxygen bands; electronic levels of the oxygen molecule". Nature. 122: 505. Bibcode:1928Natur.122..505M. doi:10.1038/122505a0.

Further reading

  • Bodner, G.M. (2002) Lecture Demonstration Movie Sheets: 8.4 Liquid Oxygen—Paramagnetism and Color, West Lafayette, IN, USA: Purdue University Department of Chemistry, see and , accessed 11 August 2015; alternatively, see Bodner, G.M.; K. Keyes & T.J. Greenbowe (1995) Purdue University Lecture Demonstration Manual, 2nd Edn, p. TBD, New York, NY, USA: John Wiley and Sons. [Earlier appearing reference on magnetic properties of oxygen states.]
This article is issued from Wikipedia. The text is licensed under Creative Commons - Attribution - Sharealike. Additional terms may apply for the media files.