Aryne

Last updated

In organic chemistry, arynes [1] and benzynes [2] are a class of highly reactive chemical species derived from an aromatic ring by removal of two substituents. Arynes are examples of didehydroarenes (1,2-didehydroarenes in this case), although 1,3- and 1,4-didehydroarenes are also known. [3] [4] [5] Arynes are examples of alkynes under high strain.

Contents

Bonding in arynes

The alkyne representation of benzyne is the most widely encountered. Arynes are usually described as having a strained triple bond. [6]

Resonance contributors1.tif

Geometric constraints on the triple bond in benzyne result in diminished overlap of in-plane p-orbitals, and thus weaker triple bond. [7] The vibrational frequency of the triple bond in benzyne was assigned by Radziszewski to be 1846 cm−1, [8] indicating a weaker triple bond than in unstrained alkyne with vibrational frequency of approximately 2150 cm−1. Nevertheless, benzyne is more like a strained alkyne than a diradical, as seen from the large singlet–triplet gap and alkyne-like reactivity. [3]

Geometric constraint2.tif

The LUMO of aryne lies much lower than the LUMO of unstrained alkynes, which makes it a better energy match for the HOMO of nucleophiles. Hence, benzyne possesses electrophilic character and undergoes reactions with nucleophiles. [9] A detailed MO analysis of benzyne was presented in 1968. [10]

Homo-lumo gap.tif

Generation of arynes

Due to their extreme reactivity, arynes must be generated in situ. Typical of other reactive intermediates, benzyne must be trapped, otherwise it dimerises to biphenylene.

Early routes to benzyne involved dehydrohalogenation of aryl halides:

Benzyne gen1.tif

Such reactions require strong base and high temperatures. 1,2-Disubstituted arenes serve as precursors to benzynes under milder conditions. Benzyne is generated by the dehalogenation of 1-bromo-2-fluorobenzene by magnesium. [11] Anthranilic acid can be converted to 2-diazoniobenzene-1-carboxylate by diazotization and neutralization. Although explosive, [12] this zwitterionic species is a convenient and inexpensive precursor to benzyne. [13]

Benzyne gen2.tif

Another method is based on trimethylsilylaryl triflates. [14] Fluoride displacement of the trimethylsilyl group induces elimination of triflate and release of benzyne:

Benzyne gen3.tif

A hexadehydro Diels-Alder reaction (HDDA) involves cycloaddition of 1,3-diyne and alkyne. [15]

HDDA rxn.tif

N-amination of 1H-benzotriazole with hydroxylamine-O-sulfonic acid generates an intermediate which can be oxidised to benzyne in almost quantitative yield with lead(IV) acetate. [16]

Benzyne Generated from 1H-Benzotriazole.png

Reactions of arynes

Even at low temperatures arynes are extremely reactive. Their reactivity can be classified in three main classes: (1) nucleophilic additions, (2) pericyclic reactions, and (3) bond-insertion.

Nucleophilic additions to arynes

Upon treatment with basic nucleophiles, aryl halides deprotonate alpha to the leaving group, resulting in dehydrohalogenation. Isotope exchange studies indicate that for aryl fluorides and, sometimes, aryl chlorides, the elimination event proceeds in two steps, deprotonation, followed by expulsion of the nucleophile. Thus, the process is formally analogous to the E1cb mechanism of aliphatic compounds. Aryl bromides and iodides, on the other hand, generally appear to undergo elimination by a concerted syn-coplanar E2 mechanism. [17] [18] The resulting benzyne forms addition products, usually by nucleophilic addition and protonation. Generation of the benzyne intermediate is the slow step in the reaction. [19]

Nuc addn to benzyne.tif

"Aryne coupling" reactions allow for generation of biphenyl compounds which are valuable in pharmaceutical industry, agriculture and as ligands in many metal-catalyzed transformations. [20]

Cross-coupling of arynes.tif

The metal–arene product can also add to another aryne, leading to chain-growth polymerization. Using copper(I) cyanide as the initiator to add to the first aryne yielded polymers containing up to about 100 arene units. [21]

When leaving group (LG) and substituent (Y) are mutually ortho or para, only one benzyne intermediate is possible. However, when LG is meta to Y, then regiochemical outcomes (A and B) are possible. If Y is electron withdrawing, then HB is more acidic than HA resulting in regioisomer B being generated. Analogously, if Y is electron donating, regioisomer A is generated, since now HA is the more acidic proton.

Triple bond generation1.png

There are two possible regioisomers of benzyne with substituent (Y): triple bond can be positioned between C2 and C3 or between C3 and C4. Substituents ortho to the leaving group will lead to the triple bond between C2 and C3. Para Y and LG will lead to regioisomer with triple bond between C3 and C4. Meta substituent can afford both regioisomers as described above. In case of triple bond located between C2 and C3, electron withdrawing (EWG) substituents, e.g. CF3, will direct the nucleophile addition to place carbanion as close as possible to EWG. However, electron donating (EDG) substituents, e.g. CH3, will provide little selectivity between products. In the regioisomer where triple bond is located between C3 and C4 the effect of substituent on nucleophile addition is diminished, and mixtures of para and meta products are often obtained. [19]

Regiochemistry2.png

Pericyclic reactions of arynes

Benzyne undergoes rapid dimerization to form biphenylene. Some routes to benzyne lead to especially rapid and high yield of this subsequent reaction. [13] [16] Trimerization gives triphenylene. [22]

Benzynes can undergo [4+2] cyclization reactions. When generated in the presence of anthracene, trypticene results. [11] In this method, the concerted mechanism of the Diels-Alder reaction between benzyne and furan is shown below. Other benzyne [4+2] cycloadditions are thought to proceed via a stepwise mechanism.

4+2 cycloaddition.svg

A classic example is the synthesis of 1,2,3,4-tetraphenylnaphthalene. [23] Tetrabromobenzene can react with butyllithium and furan to form a tetrahydroanthracene [24]

diaryne reaction with furan Diaryne reaction.svg
diaryne reaction with furan

[4+2] cycloadditions of arynes have been commonly applied to natural product total synthesis. The main limitation of such approach, however, is the need to use constrained dienes, such as furan and cyclopentadiene. [14] In 2009 Buszek and co-workers synthesized herbindole A using aryne [4+2]-cycloaddition. [25] 6,7-indolyne undergoes [4+2] cycloaddition with cyclopentadiene to afford complex tetracyclic product.

Buszek.tif

Benzynes undergo [2+2] cycloaddition with a wide range of alkenes. Due to electrophilic nature of benzyne, alkenes bearing electron-donating substituents work best for this reaction. [26]

Due to significant byproduct formation, aryne [2+2] chemistry is rarely utilized in natural product total synthesis. [14] Nevertheless, several examples do exist. In 1982, Stevens and co-workers reported a synthesis of taxodione that utilized [2+2] cycloaddition between an aryne and a ketene acetal. [27]

Stevens.tif

Mori and co-workers performed a palladium-catalyzed [2+2+2]-cocyclization of aryne and diyne in their total synthesis of taiwanins C. [28]

Mori.tif

Bond-insertion reactions of arynes

The first example of aryne σ-bond insertion reaction is the synthesis of melleine in 1973. [29]

Guyot.tif

Other dehydrobenzenes

If benzyne is 1,2-didehydrobenzene, two further isomers are possible: 1,3-didehydrobenzene and 1,4-didehydrobenzene. [3] Their energies in silico are, respectively, 106, 122, and 138 kcal/mol (444, 510 and 577 kJ/mol). [30] The 1,2- and 1,3- isomers have singlet ground states, whereas for 1,4-didehydrobenzene the gap is smaller.

Benzynes o, m, and p.svg

The interconversion of the 1,2-, 1,3- and 1,4-didehydrobenzenes has been studied. [30] [31] A 1,2- to 1,3-didehydrobenzene conversion has been postulated to occur in the pyrolysis (900 °C) of the phenyl substituted aryne precursors [30] as shown below. Extremely high temperatures are required for benzyne interconversion.

Interconversion.tif

1,4-Didehydroarenes

In classical 1,4-didehydrobenzene experiments, heating to 300 °C, [1,6-D2]-A readily equilibrates with [3,2-D2]-B, but does not equilibrate with C or D. The simultaneous migration of deuterium atoms to form B, and the fact that none of C or D is formed can only be explained by a presence of a cyclic and symmetrical intermediate–1,4-didehydrobenzene. [32]

Bergman.tif

Two states were proposed for 1,4-didehydrobenzene: singlet and triplet, with the singlet state lower in energy. [33] [34] Triplet state represents two noninteracting radical centers, and hence should abstract hydrogens at the same rate as phenyl radical. However, singlet state is more stabilized than the triplet, and therefore some of the stabilizing energy will be lost in order to form the transition state for hydrogen cleavage, leading to slower hydrogen abstraction. Chen proposed the use of 1,4-didehydrobenzene analogues that have large singlet-triplet energy gaps to enhance selectivity of enediyne drug candidates. [35]

History

The first evidence for arynes came from the work of Stoermer and Kahlert. In 1902 they observed that upon treatment of 3-bromobenzofuran with base in ethanol 2-ethoxybenzofuran is formed. Based on this observation they postulated an aryne intermediate. [36]

First indication of benzyne. Stoermer - benzyne.svg
First indication of benzyne.

Wittig et al. invoked zwitterionic intermediate in the reaction of fluorobenzene and phenyllithium to give biphenyl. [37] [38] [39] This hypothesis was later confirmed. [40] [41] [42] [43] [44]

Wittig.tif

In 1953 14C labeling experiments provided strong support for the intermediacy of benzyne. [40] John D. Roberts et al. showed that the reaction of chlorobenzene-1-14C and potassium amide gave equal amounts of aniline with 14C incorporation at C-1 and C-2.

C labeling experiment shows equal distribution of products. Roberts 1953.png
C labeling experiment shows equal distribution of products.

Wittig and Pohmer found that benzyne participate in [4+2] cycloaddition reactions. [45]

Capture of benzyne as dienophile in Diels-Alder reaction. 4+2 cycloaddition.svg
Capture of benzyne as dienophile in Diels-Alder reaction.

Additional evidence for the existence of benzyne came from spectroscopic studies. [3] Benzyne has been observed in a "molecular container". [46]

In 2015, a single aryne molecule was imaged by STM. [47]

1,3-Didehydroarenes was first demonstrated in the 1990s when it was generated from 1,3-disubstituted benzene derivatives, such as the peroxy ester 1,3-C6H4(O2C(O)CH3)2. [3]

Breakthroughs on 1,4-didehydrobenzene came in the 1960s, followed from studies on the Bergman cyclization. [32] This theme became topical with the discovery of enediyne "cytostatics", such as calicheamicin, which generates a 1,4-didehydrobenzene. [48]

Examples of benzynes in total synthesis

A variety of natural products have been prepared using arynes as intermediates. [14] Nucleophilic additions to arynes have been widely used in natural product total synthesis. Indeed, nucleophilic additions of arynes are some of the oldest known applications of aryne chemistry. [14] Nucleophilic addition to aryne was used in the attempted synthesis of cryptaustoline (1) and cryptowoline (2). [49]

Kametani1.tif

The synthesis of the tetracyclic meroterpenoid (+)-liphagal involved an aryne intermediate. [50] Their approach employed an aryne cyclization to close the final ring of the natural product. [14]

Stoltz Liphagal.tif

Multicomponent reactions of arynes are powerful transformations that allow for rapid formation of 1,2-disubstituted arenes. Despite their potential utility, examples of multicomponent aryne reactions in natural product synthesis are scarce. [14] A four-component aryne coupling reaction was employed in the synthesis of dehydroaltenuene B. [51]

Barrette synthesis.tif

See also

Related Research Articles

<span class="mw-page-title-main">Alkyne</span> Hydrocarbon compound containing one or more C≡C bonds

In organic chemistry, an alkyne is an unsaturated hydrocarbon containing at least one carbon—carbon triple bond. The simplest acyclic alkynes with only one triple bond and no other functional groups form a homologous series with the general chemical formula CnH2n−2. Alkynes are traditionally known as acetylenes, although the name acetylene also refers specifically to C2H2, known formally as ethyne using IUPAC nomenclature. Like other hydrocarbons, alkynes are generally hydrophobic.

An ylide or ylid is a neutral dipolar molecule containing a formally negatively charged atom (usually a carbanion) directly attached to a heteroatom with a formal positive charge (usually nitrogen, phosphorus or sulfur), and in which both atoms have full octets of electrons. The result can be viewed as a structure in which two adjacent atoms are connected by both a covalent and an ionic bond; normally written X+–Y. Ylides are thus 1,2-dipolar compounds, and a subclass of zwitterions. They appear in organic chemistry as reagents or reactive intermediates.

The 1,3-dipolar cycloaddition is a chemical reaction between a 1,3-dipole and a dipolarophile to form a five-membered ring. The earliest 1,3-dipolar cycloadditions were described in the late 19th century to the early 20th century, following the discovery of 1,3-dipoles. Mechanistic investigation and synthetic application were established in the 1960s, primarily through the work of Rolf Huisgen. Hence, the reaction is sometimes referred to as the Huisgen cycloaddition. 1,3-dipolar cycloaddition is an important route to the regio- and stereoselective synthesis of five-membered heterocycles and their ring-opened acyclic derivatives. The dipolarophile is typically an alkene or alkyne, but can be other pi systems. When the dipolarophile is an alkyne, aromatic rings are generally produced.

An alkyne trimerisation is a [2+2+2] cycloaddition reaction in which three alkyne units react to form a benzene ring. The reaction requires a metal catalyst. The process is of historic interest as well as being applicable to organic synthesis. Being a cycloaddition reaction, it has high atom economy. Many variations have been developed, including cyclisation of mixtures of alkynes and alkenes as well as alkynes and nitriles.

In chemical synthesis, click chemistry is a class of simple, atom-economy reactions commonly used for joining two molecular entities of choice. Click chemistry is not a single specific reaction, but describes a way of generating products that follow examples in nature, which also generates substances by joining small modular units. In many applications, click reactions join a biomolecule and a reporter molecule. Click chemistry is not limited to biological conditions: the concept of a "click" reaction has been used in chemoproteomic, pharmacological, biomimetic and molecular machinery applications. However, they have been made notably useful in the detection, localization and qualification of biomolecules.

The azide-alkyne Huisgen cycloaddition is a 1,3-dipolar cycloaddition between an azide and a terminal or internal alkyne to give a 1,2,3-triazole. Rolf Huisgen was the first to understand the scope of this organic reaction. American chemist Karl Barry Sharpless has referred to this cycloaddition as "the cream of the crop" of click chemistry and "the premier example of a click reaction".

<span class="mw-page-title-main">Bamford–Stevens reaction</span> Synthesis of alkenes by base-catalysed decomposition of tosylhydrazones

The Bamford–Stevens reaction is a chemical reaction whereby treatment of tosylhydrazones with strong base gives alkenes. It is named for the British chemist William Randall Bamford and the Scottish chemist Thomas Stevens Stevens (1900–2000). The usage of aprotic solvents gives predominantly Z-alkenes, while protic solvent gives a mixture of E- and Z-alkenes. As an alkene-generating transformation, the Bamford–Stevens reaction has broad utility in synthetic methodology and complex molecule synthesis.

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

In chemistry, a phosphaalkyne is an organophosphorus compound containing a triple bond between phosphorus and carbon with the general formula R-C≡P. Phosphaalkynes are the heavier congeners of nitriles, though, due to the similar electronegativities of phosphorus and carbon, possess reactivity patterns reminiscent of alkynes. Due to their high reactivity, phosphaalkynes are not found naturally on earth, but the simplest phosphaalkyne, phosphaethyne (H-C≡P) has been observed in the interstellar medium.

In organic chemistry, a cycloalkyne is the cyclic analog of an alkyne. A cycloalkyne consists of a closed ring of carbon atoms containing one or more triple bonds. Cycloalkynes have a general formula CnH2n−4. Because of the linear nature of the C−C≡C−C alkyne unit, cycloalkynes can be highly strained and can only exist when the number of carbon atoms in the ring is great enough to provide the flexibility necessary to accommodate this geometry. Large alkyne-containing carbocycles may be virtually unstrained, while the smallest constituents of this class of molecules may experience so much strain that they have yet to be observed experimentally. Cyclooctyne is the smallest cycloalkyne capable of being isolated and stored as a stable compound. Despite this, smaller cycloalkynes can be produced and trapped through reactions with other organic molecules or through complexation to transition metals.

<span class="mw-page-title-main">Organonickel chemistry</span> Branch of organometallic chemistry

Organonickel chemistry is a branch of organometallic chemistry that deals with organic compounds featuring nickel-carbon bonds. They are used as a catalyst, as a building block in organic chemistry and in chemical vapor deposition. Organonickel compounds are also short-lived intermediates in organic reactions. The first organonickel compound was nickel tetracarbonyl Ni(CO)4, reported in 1890 and quickly applied in the Mond process for nickel purification. Organonickel complexes are prominent in numerous industrial processes including carbonylations, hydrocyanation, and the Shell higher olefin process.

<span class="mw-page-title-main">Vinyl cation</span> Organic cation

The vinyl cation is a carbocation with the positive charge on an alkene carbon. Its empirical formula is C
2
H+
3
. More generally, a vinylic cation is any disubstituted carbon, where the carbon bearing the positive charge is part of a double bond and is sp hybridized. In the chemical literature, substituted vinylic cations are often referred to as vinyl cations, and understood to refer to the broad class rather than the C
2
H+
3
variant alone. The vinyl cation is one of the main types of reactive intermediates involving a non-tetrahedrally coordinated carbon atom, and is necessary to explain a wide variety of observed reactivity trends. Vinyl cations are observed as reactive intermediates in solvolysis reactions, as well during electrophilic addition to alkynes, for example, through protonation of an alkyne by a strong acid. As expected from its sp hybridization, the vinyl cation prefers a linear geometry. Compounds related to the vinyl cation include allylic carbocations and benzylic carbocations, as well as aryl carbocations.

<span class="mw-page-title-main">Germylene</span> Class of germanium (II) compounds

Germylenes are a class of germanium(II) compounds with the general formula :GeR2. They are heavier carbene analogs. However, unlike carbenes, whose ground state can be either singlet or triplet depending on the substituents, germylenes have exclusively a singlet ground state. Unprotected carbene analogs, including germylenes, has a dimerization nature. Free germylenes can be isolated under the stabilization of steric hindrance or electron donation. The synthesis of first stable free dialkyl germylene was reported by Jutzi, et al in 1991.

In organic chemistry, enone–alkene cycloadditions are a version of the [2+2] cycloaddition This reaction involves an enone and alkene as substrates. Although the concerted photochemical [2+2] cycloaddition is allowed, the reaction between enones and alkenes is stepwise and involves discrete diradical intermediates.

The nitrone-olefin (3+2) cycloaddition reaction is the combination of a nitrone with an alkene or alkyne to generate an isoxazoline or isoxazolidine via a (3+2) cycloaddition process. This reaction is a 1,3-dipolar cycloaddition, in which the nitrone acts as the 1,3-dipole, and the alkene or alkyne as the dipolarophile.

<i>tert</i>-Butylphosphaacetylene Chemical compound

tert-Butylphosphaacetylene is an organophosphorus compound. Abbreviated t-BuCP, it was the first example of an isolable phosphaalkyne. Prior to its synthesis, the double bond rule had suggested that elements of Period 3 and higher were unable to form double or triple bonds with lighter main group elements because of weak orbital overlap. The synthesis of t-BuCP discredited much of the double bond rule and opened new studies into the formation of unsaturated phosphorus compounds.

The term bioorthogonal chemistry refers to any chemical reaction that can occur inside of living systems without interfering with native biochemical processes. The term was coined by Carolyn R. Bertozzi in 2003. Since its introduction, the concept of the bioorthogonal reaction has enabled the study of biomolecules such as glycans, proteins, and lipids in real time in living systems without cellular toxicity. A number of chemical ligation strategies have been developed that fulfill the requirements of bioorthogonality, including the 1,3-dipolar cycloaddition between azides and cyclooctynes, between nitrones and cyclooctynes, oxime/hydrazone formation from aldehydes and ketones, the tetrazine ligation, the isocyanide-based click reaction, and most recently, the quadricyclane ligation.

In organic chemistry, the hexadehydro-Diels–Alder (HDDA) reaction is an organic chemical reaction between a diyne and an alkyne to form a reactive benzyne species, via a [4+2] cycloaddition reaction. This benzyne intermediate then reacts with a suitable trapping agent to form a substituted aromatic product. This reaction is a derivative of the established Diels–Alder reaction and proceeds via a similar [4+2] cycloaddition mechanism. The HDDA reaction is particularly effective for forming heavily functionalized aromatic systems and multiple ring systems in one synthetic step.

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

Digermynes are a class of compounds that are regarded as the heavier digermanium analogues of alkynes. The parent member of this entire class is HGeGeH, which has only been characterized computationally, but has revealed key features of the whole class. Because of the large interatomic repulsion between two Ge atoms, only kinetically stabilized digermyne molecules can be synthesized and characterized by utilizing bulky protecting groups and appropriate synthetic methods, for example, reductive coupling of germanium(II) halides.

<span class="mw-page-title-main">1,3-Diphenylisobenzofuran</span> Chemical compound

1,3-Diphenylisobenzofuran is a highly reactive diene that can scavenge unstable and short-lived dienophiles in a Diels-Alder reaction. It is furthermore used as a standard reagent for the determination of singlet oxygen, even in biological systems. Cycloadditions with 1,3-diphenylisobenzofuran and subsequent oxygen cleavage provide access to a variety of polyaromatics.

An organic azide is an organic compound that contains an azide functional group. Because of the hazards associated with their use, few azides are used commercially although they exhibit interesting reactivity for researchers. Low molecular weight azides are considered especially hazardous and are avoided. In the research laboratory, azides are precursors to amines. They are also popular for their participation in the "click reaction" between an azide and an alkyne and in Staudinger ligation. These two reactions are generally quite reliable, lending themselves to combinatorial chemistry.

References

  1. IUPAC , Compendium of Chemical Terminology , 2nd ed. (the "Gold Book") (1997). Online corrected version: (2006) " Aryne ". doi : 10.1351/goldbook.A00465
  2. IUPAC , Compendium of Chemical Terminology , 2nd ed. (the "Gold Book") (1997). Online corrected version: (2006) " Benzynes ". doi : 10.1351/goldbook.B00634
  3. 1 2 3 4 5 Hans Henning Wenk; Michael Winkler; Wolfram Sander (2003). "One Century of Aryne Chemistry". Angew. Chem. Int. Ed. 42 (5): 502–528. doi:10.1002/anie.200390151. PMID   12569480.
  4. IUPAC , Compendium of Chemical Terminology , 2nd ed. (the "Gold Book") (1997). Online corrected version: (2006) " Dehydroarenes ". doi : 10.1351/goldbook.D01574
  5. IUPAC Gold Book entry for "benzynes" identifies "m-benzyne" and "p-benzyne" as erroneous terms for 1,3- and 1,4-didehydrobenzene
  6. Anslyn, E. V.; Dougherty, D. A.: Modern Physical Organic Chemistry, University Science Books, 2006, p612.
  7. Gampe, C. M.; Carreira, E. M. (2012). "Arynes and Cyclohexyne in Natural Product Synthesis". Angew. Chem. Int. Ed. Engl. 51 (16): 3766–78. doi:10.1002/anie.201107485. PMID   22422638.
  8. Radziszewski, J. G.; Hess, B. A. Jr.; Zahradnik, R. (1992). "Infrared Spectrum of o-Benzyne: Experiment and Theory". J. Am. Chem. Soc. 114: 52. doi:10.1021/ja00027a007.
  9. Gilchrist, T. L. Supplement C: The Chemistry of Triple Bonded Functional Groups, Part 1. Patai, S.; Rappaport, Z. Eds., John Wiley & Sons, New York, 1983
  10. Hoffmann, R.; Imamura, A.; Hehre, W. J. (1968). "Benzynes, dehydroconjugated molecules, and the interaction of orbitals separated by a number of intervening sigma bonds". J. Am. Chem. Soc. 90 (6): 1499. doi:10.1021/ja01008a018.
  11. 1 2 Wittig, Georg (1959). "Triptycene". Org. Synth. 39: 75. doi:10.15227/orgsyn.039.0075.
  12. Sullivan, John M. (1971-06-01). "Explosion during preparation of benzenediazonium-2-carboxylate hydrochloride". Journal of Chemical Education. 48 (6): 419. Bibcode:1971JChEd..48..419S. doi:10.1021/ed048p419.3. ISSN   0021-9584.
  13. 1 2 Logullo, Francis M.; Seitz, Arnold M.; Friedman, Lester (1968). "Benzenediazonium-2-Carboxylate and Biphenylene (Benzenediazonium, o-carboxy-, hydroxide, inner salt)". Org. Synth. 48: 12. doi:10.15227/orgsyn.048.0012.
  14. 1 2 3 4 5 6 7 Tadross, P. M.; Stoltz, B. M. (2012). "A Comprehensive History of Arynes in Natural Product Total Synthesis". Chem. Rev. 112 (6): 3550–3577. doi:10.1021/cr200478h. PMID   22443517.
  15. Hoye, T. R.; Baire, B.; Niu, D.; Willoughby, P. H.; Woods, B. P. (2012). "The hexadehydro-Diels–Alder reaction". Nature. 490 (7419): 208–212. Bibcode:2012Natur.490..208H. doi:10.1038/nature11518. PMC   3538845 . PMID   23060191.
  16. 1 2 Campbell, C.D.; C.W. Rees (1969). "Reactive intermediates. Part I. Synthesis and oxidation of 1- and 2-aminobenzotriazole". J. Chem. Soc. C . 1969 (5): 742–747. doi:10.1039/J39690000742.
  17. Panar, Manuel (1961). The Elimination-Addition Mechanism of Nucleophilic Aromatic Substitution. Pasadena, CA: California Institute of Technology (Ph.D. Thesis). pp. 4–5.
  18. H., Lowry, Thomas (1987). Mechanism and theory in organic chemistry . Richardson, Kathleen Schueller (3rd ed.). New York: Harper & Row. pp.  643. ISBN   0-06-044084-8. OCLC   14214254.{{cite book}}: CS1 maint: multiple names: authors list (link)
  19. 1 2 Anslyn, E. V.; Dougherty, D. A. Modern Physical Organic Chemistry. University Science Books, 2006
  20. Diemer, V.; Begaut, M.; Leroux, F. R.; Colobert, F. Eur. J. Org. Chem.2011, 341
  21. Mizukoshi, Yoshihide; Mikami, Koichiro; Uchiyama, Masanobu (2015). "Aryne Polymerization Enabling Straightforward Synthesis of Elusive Poly(ortho-arylene)s". J. Am. Chem. Soc. 137 (1): 74–77. doi:10.1021/ja5112207. PMID   25459083.
  22. Heaney, H.; Millar, I. T. (1960). "Triphenylene". Organic Syntheses . 40: 105; Collected Volumes, vol. 5, 1973, p. 1120.
  23. "1,2,3,4-Tetraphenylnaphthalene". Organic Syntheses. 46: 107. 1966. doi:10.15227/orgsyn.046.0107.
  24. "Use of 1,2,4,5-Tetrabromobenzene as a 1,4-Nenzadiyne Equivalent: Anti- and Syn-1,4,5,8-tetrahydroanthracene 1,4:5,8-diepoxides". Organic Syntheses. 75: 201. 1998. doi:10.15227/orgsyn.075.0201.
  25. Buszek, K. R.; Brown, N.; Kuo, D. (2009). "Concise Total Synthesis of (±)-cis-Trikentrin A and (±)-Herbindole A via Intermolecular Indole Aryne Cycloaddition". Org. Lett. 11 (1): 201–204. doi:10.1021/ol802425m. PMC   2723800 . PMID   19055375.
  26. Pellissier, H.; Santelli, M. Tetrahedron, 2003; 59, 701
  27. Stevens, R. V.; Bisacchi, G. S. J. Org, Chem. 1982; 47, 2396
  28. Sato, Y.; Tamura,T.; Mori, M. Angew. Chem. Int. Ed. 2004; 43, 2436
  29. Guyot, M.; Molho, D. Tetrahedron Lett. 1973; 14, 3433
  30. 1 2 3 Blake, M. E.; Bartlett, K. L.; Jones, M. Jr (2003). "A m-Benzyne to o-Benzyne Conversion Through a 1,2-Shift of a Phenyl Group". J. Am. Chem. Soc. 125 (21): 6485–90. doi:10.1021/ja0213672. PMID   12785789.
  31. Polishchuk, A. L.; Bartlett, K. L.; Friedman, L. A.; Jones, M. Jr (2004). "A p-Benzyne to m-Benzyne Conversion Through a 1,2-Shift of a Phenyl Group. Completion of the Benzyne Cascade". J. Phys. Org. Chem. 17 (9): 798–806. doi:10.1002/poc.797.
  32. 1 2 Richard R. Jones; Robert G. Bergman (1972). "p-Benzyne. Generation as an intermediate in a thermal isomerization reaction and trapping evidence for the 1,4-benzenediyl structure". J. Am. Chem. Soc. 94 (2): 660–661. doi:10.1021/ja00757a071.
  33. Clauberg, H.; Minsek, D. W.; Chen, P. (1992). "Mass and photoelectron spectroscopy of C3H2. .DELTA.Hf of singlet carbenes deviate from additivity by their singlet-triplet gaps". J. Am. Chem. Soc. 114: 99. doi:10.1021/ja00027a014.
  34. Blush, J. A.; Clauberg, H.; Kohn, D. W.; Minsek, D. W.; Zhang, X.; Chen, P. (1992). "Photoionization mass and photoelectron spectroscopy of radicals, carbenes, and biradicals". Acc. Chem. Res. 25 (9): 385. doi:10.1021/ar00021a001.
  35. Chen, P (1996). "Design of Diradical-based Hydrogen Abstraction Agents". Angew. Chem. Int. Ed. Engl. 35 (1314): 1478. doi: 10.1002/anie.199614781 .
  36. Stoermer, R.; Kahlert, B. (1902). "Ueber das 1- und 2-Brom-cumaron". Berichte der Deutschen Chemischen Gesellschaft. 35 (2): 1633–1640. doi:10.1002/cber.19020350286.
  37. Wittig, G.; Pieper, G.; Fuhrmann, G. (1940). "Über die Bildung von Diphenyl aus Fluorbenzol und Phenyl-lithium (IV. Mitteil. über Austauschreaktionen mit Phenyl-lithium)". Berichte der Deutschen Chemischen Gesellschaft (A and B Series). 73 (11): 1193–1197. doi:10.1002/cber.19400731113.
  38. Wittig, Georg (1942). "Phenyl-lithium, der Schlüssel zu einer neuen Chemie metallorganischer Verbindungen". Die Naturwissenschaften. 30 (46–47): 696–703. Bibcode:1942NW.....30..696W. doi:10.1007/BF01489519. S2CID   37148502.
  39. Wittig, G (1954). "Fortschritte auf dem Gebiet der organischen Aniono-Chemie". Angewandte Chemie. 66 (1): 10–17. Bibcode:1954AngCh..66...10W. doi:10.1002/ange.19540660103.
  40. 1 2 Roberts, John D. (1953). "Rearrangement in the Reaction of Chlorobenzene-1-C14With Potassium Amide1". Journal of the American Chemical Society. 75 (13): 3290–3291. doi:10.1021/ja01109a523.
  41. Roberts, John D. (1956). "The Mechanism of Aminations of Halobenzenes 1". Journal of the American Chemical Society. 78 (3): 601–611. doi:10.1021/ja01584a024.
  42. Roberts, John D. (1956). "Orientation in Aminations of Substituted Halobenzenes 1". Journal of the American Chemical Society. 78 (3): 611–614. doi:10.1021/ja01584a025.
  43. Modern Arylation Methods. Edited by Lutz Ackermann 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim ISBN   978-3-527-31937-4
  44. Heaney, H. (1962). "The Benzyne and Related Intermediates". Chemical Reviews. 62 (2): 81–97. doi:10.1021/cr60216a001.
  45. Wittig, G.; Pohmer, L. Angew. Chem. 1955; 67(13), 348.
  46. Warmuth, R.; Yoon (2001). "Recent highlights in hemicarcerand chemistry". Acc. Chem. Res. 34 (2): 96. doi:10.1021/ar980082k. PMID   11263868.
  47. On-surface; Pérez, E.Guitián; Peña, L.Gross (2015). "On-surface generation and imaging of arynes by atomic force microscopy". Nature Chemistry. 7 (8): 623–8. Bibcode:2015NatCh...7..623P. doi:10.1038/nchem.2300. PMID   26201737.
  48. Galm, U; Hager, MH; Van Lanen, SG; Ju, J; Thorson, JS; Shen, B (Feb 2005). "Antitumor antibiotics: bleomycin, enediynes, and mitomycin". Chemical Reviews. 105 (2): 739–58. doi:10.1021/cr030117g. PMID   15700963.
  49. Kametani, T.; Ogasawara, K. J. J. Chem. Soc., C 1967, 2208
  50. Day, J. J.; McFadden, R. M.; Virgil, S. C.; Kolding, H.; Alleva, J. L.; Stoltz, B. M. (2011). "The catalytic enantioselective total synthesis of (+)-liphagal". Angew. Chem. Int. Ed. 50 (30): 6814–8. doi:10.1002/anie.201101842. PMC   3361906 . PMID   21671325.
  51. Soorukram, D.; Qu, T.; Barrett, A. G. M. (2008). "Four-Component Benzyne Coupling Reactions: A Concise Total Synthesis of Dehydroaltenuene B". Org. Lett. 10 (17): 3833–3835. doi:10.1021/ol8015435. PMID   18672878.