Singlet oxygen

Last updated
Singlet oxygen
Dioxygen-3D-ball-&-stick.png
Dioxygen-3D-vdW.png
Names
IUPAC name
Singlet oxygen
Systematic IUPAC name
Dioxidene
Identifiers
3D model (JSmol)
ChEBI
491
  • InChI=1S/O2/c1-2
    Key: MYMOFIZGZYHOMD-UHFFFAOYSA-N
  • O=O
Properties
O2
Molar mass 31.998 g·mol−1
Reacts
Except where otherwise noted, data are given for materials in their standard state (at 25 °C [77 °F], 100 kPa).

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.

Contents

The lowest excited state of the diatomic oxygen molecule is a 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 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.[ citation needed ]

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 lowest 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 superscript "1" indicates 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] This more stable of the two excited states has its two valence electrons spin-paired in one π* orbital while the second π* orbital is empty. This state 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 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]

Molecular orbital diagram of two singlet excited states as well as the triplet ground state of molecular dioxygen. From left to right, the diagrams are for: Dg singlet oxygen (first excited state), S
g singlet oxygen (second excited state), and S
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 p and p* 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 p* antibonding orbitals. Molecular orbital scheme for the three forms of oxygen.png
Molecular orbital diagram of two singlet excited states as well as the triplet ground state of molecular dioxygen. From left to right, the diagrams are for: Δg singlet oxygen (first excited state), Σ
g
singlet oxygen (second excited state), and Σ
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 singlet-singlet transition between the two excited states is spin-allowed but parity-forbidden.

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 relatively long lived (54-86 milliseconds), [13] although interaction with solvents reduces the lifetime to microseconds or even nanoseconds. [14] In 2021, the lifetime of airborne singlet oxygen at air/solid interfaces was measured to be 550 microseconds. [15]

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. [16] [17] [18] The paramagnetism of the 1Δg state is due to a net orbital (and not spin) electronic angular momentum. In a magnetic field the degeneracy of the levels is split into two levels with z projections of angular momenta +1ħ and −1ħ around the molecular axis. The magnetic transition between these levels gives rise to the EPR transition.

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. [19] [9] Large steady state concentrations of singlet oxygen are reported from the reaction of triplet excited state pyruvic acid with dissolved oxygen in water. [20] Singlet oxygen can also be produced in non-photochemical, preparative chemical procedures. One chemical method involves the decomposition of triethylsilyl hydrotrioxide generated in situ from triethylsilane and ozone. [21]

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

Another method uses the aqueous reaction of hydrogen peroxide with sodium hypochlorite: [19]

H2O2 + NaOCl → O2(1Δg) + NaCl + H2O

A third method liberates singlet oxygen via phosphite ozonides, which are, in turn, generated in situ such as triphenyl phosphite ozonide. [22] [23] Phosphite ozonides will decompose to give singlet oxygen: [24]

(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. [24]

Reactions

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. Singlet Oxygenation Citronellol.svg
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.

Because of differences in their electron shells, singlet and triplet oxygen differ in their chemical properties; singlet oxygen is highly reactive. [25] The lifetime of singlet oxygen depends on the medium and pressure. 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. [24] [26]

Organic chemistry

Unlike ground state oxygen, singlet oxygen participates in Diels–Alder [4+2]- and [2+2]-cycloaddition reactions and formal concerted ene reactions. [24] It oxidizes thioethers to sulfoxides. Organometallic complexes are often degraded by singlet oxygen. [27] [28] With some substrates 1,2-dioxetanes are formed; cyclic dienes such as 1,3-cyclohexadiene form [4+2] cycloaddition adducts. [29]

The [4+2]-cycloaddition between singlet oxygen and furans is widely used in organic synthesis. [30] [31]

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. [24] [32] [33] [34]

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

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. [35]

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 associated with decary of singlet oxygen to its triplet state. Singlet oxygen Glow.jpg
Red glow associated with decary of singlet oxygen to its triplet state.

Singlet oxygen luminesces concomitant with its decay to the triplet ground state. This phenomenon was first observed in the thermal degradation of the endo peroxide of rubrene. [36]

Further reading

Related Research Articles

<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">Excited state</span> Quantum states with more energy than the lowest possible amount

In quantum mechanics, an excited state of a system is any quantum state of the system that has a higher energy than the ground state. Excitation refers to an increase in energy level above a chosen starting point, usually the ground state, but sometimes an already excited state. The temperature of a group of particles is indicative of the level of excitation.

<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.

<span class="mw-page-title-main">Sulfur monoxide</span> Chemical compound

Sulfur monoxide is an inorganic compound with formula SO. It is only found as a dilute gas phase. When concentrated or condensed, it converts to S2O2 (disulfur dioxide). It has been detected in space but is rarely encountered intact otherwise.

<span class="mw-page-title-main">Ozonide</span> Polyatomic ion (O3, charge –1), or cyclic compounds made from ozone and alkenes

Ozonide is the polyatomic anion O−3. Cyclic organic compounds formed by the addition of ozone to an alkene are also called ozonides.

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

Photosensitizers are light absorbers that alter 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.

<span class="mw-page-title-main">Triplet oxygen</span> Triplet state of the dioxygen molecule

Triplet oxygen, 3O2, refers to the S = 1 electronic ground state of molecular oxygen (dioxygen). Molecules of triplet oxygen contain two unpaired electrons, making triplet oxygen an unusual example of a stable and commonly encountered diradical: it is more stable as a triplet than a singlet. According to molecular orbital theory, the electron configuration of triplet oxygen has two electrons occupying two π molecular orbitals (MOs) of equal energy (that is, degenerate MOs). In accordance with Hund's rules, they remain unpaired and spin-parallel, which accounts for the paramagnetism of molecular oxygen. These half-filled orbitals are antibonding in character, reducing the overall bond order of the molecule to 2 from the maximum value of 3 that would occur when these antibonding orbitals remain fully unoccupied, as in dinitrogen. The molecular term symbol for triplet oxygen is 3Σ
g
.

In chemistry, a diradical is a molecular species with two electrons occupying molecular orbitals (MOs) which are degenerate. The term "diradical" is mainly used to describe organic compounds, where most diradicals are extremely reactive and in fact rarely isolated. Diradicals are even-electron molecules but have one fewer bond than the number permitted by the octet rule.

<span class="mw-page-title-main">Diatomic carbon</span> Chemical compound

Diatomic carbon (systematically named dicarbon and 2,2λ2-ethene), is a green, gaseous inorganic chemical with the chemical formula C=C (also written [C2] or C2). It is kinetically unstable at ambient temperature and pressure, being removed through autopolymerisation. It occurs in carbon vapor, for example in electric arcs; in comets, stellar atmospheres, and the interstellar medium; and in blue hydrocarbon flames. Diatomic carbon is the second simplest of the allotropes of carbon (after atomic carbon), and is an intermediate participator in the genesis of fullerenes.

<span class="mw-page-title-main">Tris(bipyridine)ruthenium(II) chloride</span> Chemical compound

Tris(bipyridine)ruthenium(II) chloride is the chloride salt coordination complex with the formula [Ru(bpy)3]Cl2. This polypyridine complex is a red crystalline salt obtained as the hexahydrate, although all of the properties of interest are in the cation [Ru(bpy)3]2+, which has received much attention because of its distinctive optical properties. The chlorides can be replaced with other anions, such as PF6.

In spectroscopy and quantum chemistry, the multiplicity of an energy level is defined as 2S+1, where S is the total spin angular momentum. States with multiplicity 1, 2, 3, 4, 5 are respectively called singlets, doublets, triplets, quartets and quintets.

In organic chemistry, the di-π-methane rearrangement is the photochemical rearrangement of a molecule that contains two π-systems separated by a saturated carbon atom. In the aliphatic case, this molecules is a 1,4-diene; in the aromatic case, an allyl-substituted arene. The reaction forms (respectively) an ene- or aryl-substituted cyclopropane. Formally, it amounts to a 1,2 shift of one ene group or the aryl group, followed by bond formation between the lateral carbons of the non-migrating moiety:

There are several known allotropes of oxygen. The most familiar is molecular oxygen, present at significant levels in Earth's atmosphere and also known as dioxygen or triplet oxygen. Another is the highly reactive ozone. Others are:

<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.

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">Photooxygenation</span> Light-induced oxidation reaction

A photooxygenation is a light-induced oxidation reaction in which molecular oxygen is incorporated into the product(s). Initial research interest in photooxygenation reactions arose from Oscar Raab's observations in 1900 that the combination of light, oxygen and photosensitizers is highly toxic to cells. Early studies of photooxygenation focused on oxidative damage to DNA and amino acids, but recent research has led to the application of photooxygenation in organic synthesis and photodynamic therapy.

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

9-Fluorenylidene is an aryl carbene derived from the bridging methylene group of fluorene. Fluorenylidene has the unusual property that the triplet ground state is only 1.1 kcal/mol lower in energy than the singlet state. For this reason, fluorenylidene has been studied extensively in organic chemistry.

The magnesium argide ion, MgAr+ is an ion composed of one ionised magnesium atom, Mg+ and an argon atom. It is important in inductively coupled plasma mass spectrometry and in the study of the field around the magnesium ion. The ionization potential of magnesium is lower than the first excitation state of argon, so the positive charge in MgAr+ will reside on the magnesium atom. Neutral MgAr molecules can also exist in an excited state.

Thermally activated delayed fluorescence (TADF) is a process through which a molecular species in a non-emitting excited state can incorporate surrounding thermal energy to change states and only then undergo light emission. The TADF process usually involves an excited molecular species in a triplet state, which commonly has a forbidden transition to the ground state termed phosphorescence. By absorbing nearby thermal energy the triplet state can undergo reverse intersystem crossing (RISC) converting it to a singlet state, which can then de-excite to the ground state and emit light in a process termed fluorescence. Along with fluorescent and phosphorescent compounds, TADF compounds are one of the three main light-emitting materials used in organic light-emitting diodes (OLEDs). Although most TADF molecules rely on the RISC from a triplet state to a singlet state, some of them take advantage of RISC processes between states with other spin multiplicities instead, for example from a quartet state to a doublet state.

<span class="mw-page-title-main">Triphenyl phosphite ozonide</span> Chemical compound

Triphenyl phosphite ozonide (TPPO) is a chemical compound with the formula PO3(C6H5O)3 that is used to generate singlet oxygen.

References

  1. Wayne RP (1969). "Singlet Molecular Oxygen". In Pitts JN, Hammond GS, Noyes WA (eds.). Advances in Photochemistry. Vol. 7. pp. 311–71. doi:10.1002/9780470133378.ch4. ISBN   9780470133378.
  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   978-1405190886.
  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   978-0-7167-8759-4.
  4. Hill C. "Molecular Term Symbols" (PDF). Archived from the original (PDF) on 5 September 2017. Retrieved 10 October 2016.
  5. Levine IN (1991). Quantum Chemistry (4th ed.). Prentice-Hall. p. 383. ISBN   978-0-205-12770-2.
  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. doi: 10.1111/j.1751-1097.1999.tb08238.x . S2CID   94065922.
  11. Thomas Engel; Philip Reid (2006). Physical Chemistry. PEARSON Benjamin Cummings. p. 580. ISBN   978-0-8053-3842-3.
  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. Physical Mechanisms of Generation and Deactivation of Singlet Oxygen Claude Schweitzer
  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. S2CID   9214506.
  15. Andrés M. Durantini (2021). "Interparticle Delivery and Detection of Volatile Singlet Oxygen at Air/Solid Interfaces". Environmental Science & Technology. 55 (6): 3559–3567. Bibcode:2021EnST...55.3559D. doi:10.1021/acs.est.0c07922. PMID   33660980. S2CID   232114444.
  16. 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 (4): 312–314. Bibcode:2008CPL...457..312H. doi:10.1016/j.cplett.2008.04.031.
  17. 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. CiteSeerX   10.1.1.652.974 . doi:10.1021/jp403648d. PMID   23768193.
  18. Falick AM, et al. (1965). "Paramagnetic resonance spectrum of the 1?g oxygen molecule". J. Chem. Phys. 42 (5): 1837–1838. Bibcode:1965JChPh..42.1837F. doi:10.1063/1.1696199. S2CID   98040975.
  19. 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.
  20. Eugene AJ, Guzman MI (September 2019). "Production of Singlet Oxygen (1O2) during the Photochemistry of Aqueous Pyruvic Acid: The Effects of pH and Photon Flux under Steady-State O2(aq) Concentration". Environmental Science and Technology. 53 (21): 12425–12432. Bibcode:2019EnST...5312425E. doi: 10.1021/acs.est.9b03742 . PMID   31550134.
  21. 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.
  22. Bartlett, Paul D.; Mendenhall, G. David; Durham, Dana L. (October 1980). "Controlled generation of singlet oxygen at low temperatures from triphenyl phosphite ozonide". The Journal of Organic Chemistry. 45 (22): 4269–4271. doi:10.1021/jo01310a001. ISSN   0022-3263.
  23. Housecroft CE, Sharpe AG (2008). "Chapter 15: The group 16 elements". Inorganic Chemistry (3rd ed.). Pearson. p.  438 f. ISBN   9780131755536.
  24. 1 2 3 4 5 Wasserman HH, DeSimone RW, Chia KR, Banwell MG (2001). "Singlet Oxygen". Encyclopedia of Reagents for Organic Synthesis. e-EROS Encyclopedia of Reagents for Organic Synthesis. John Wiley & Sons. doi:10.1002/047084289X.rs035. ISBN   978-0471936237.
  25. Ho RY, Liebman JF, Valentine JS (1995). "Overview of the Energetics and Reactivity of Oxygen". In Foote CS (ed.). 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.
  26. Kuntner N (2018). "Modeling and simulation of electronic excitation in oxygen-helium discharges and plasma-assisted combustion". University of Stuttgart. doi = http://dx.doi.org/10.18419/opus-9925
  27. Clennan EL, Pace A (2005). "Advances in singlet oxygen chemistry". Tetrahedron. 61 (28): 6665–6691. doi:10.1016/j.tet.2005.04.017.
  28. 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.
  29. Carey FA, Sundberg RJ (1985). Structure and mechanisms (2 ed.). New York: Plenum Press. ISBN   978-0306411984.
  30. Montagnon, T.; Kalaitzakis, D.; Triantafyllakis, M.; Stratakis, M.; Vassilikogiannakis, G. (2014). "Furans and Singlet Oxygen - Why There Is More to Come from this Powerful Partnership". Chemical Communications. 50 (98): 15480–15498. doi:10.1039/C4CC02083A. PMID   25316254.
  31. Ghogare, A.A.; Greer, A. (2016). "Using Singlet Oxygen to Synthesise Natural Products and Drugs". Chemical Reviews. 116 (17): 9994–10034. doi:10.1021/acs.chemrev.5b00726. PMID   27128098.
  32. 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.
  33. 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.
  34. 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.
  35. 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.
  36. Franz, Karl A.; Kehr, Wolfgang G.; Siggel, Alfred; Wieczoreck, Jürgen; Adam, Waldemar (2000). "Luminescent Materials". Ullmann's Encyclopedia of Industrial Chemistry. doi:10.1002/14356007.a15_519. ISBN   3-527-30673-0.

Further reading