Vinyl iodide functional group

Last updated
General structure of vinyl iodides Vinyl iodide general structure.png
General structure of vinyl iodides

In organic chemistry, a vinyl iodide (also known as an iodoalkene) functional group is an alkene with one or more iodide substituents. Vinyl iodides are versatile molecules that serve as important building blocks and precursors in organic synthesis. They are commonly used in carbon-carbon forming reactions in transition-metal catalyzed cross-coupling reactions, such as Stille reaction, Heck reaction, Sonogashira coupling, and Suzuki coupling. [1] Synthesis of well-defined geometry or complexity vinyl iodide is important in stereoselective synthesis of natural products and drugs.

Contents

Properties

Vinyl iodides are generally stable under nucleophilic conditions. In SN2 reactions, back-attack is difficult because of steric clash of R groups on carbon adjacent to electrophilic center (see figure 1a). [2] In addition, the lone pair on iodide donates into the ╥* of the alkene, which reduces electrophilic character on the carbon as a result of decreased positive charge. Also, this stereoelectronic effect strengthens the C-I bond, thus making removal of the iodide difficult (see figure 1b). [3] In SN1 case, dissociation is difficult because of the strengthened C-I bond and loss of the iodide will generate an unstable carbocation(see figure 1c) [2]

Figure 1. Properties of vinyl iodide.jpg
Figure 1.

In cross-coupling reactions, typically vinyl iodides react faster and under more mild conditions than vinyl chloride and vinyl bromide. The order of reactivity is based on the strength of carbon-halogen bond. C-I bond is the weakest of the halogens, the bond dissociation energies of C-I is 57.6kcal/mol, while fluoride, chloride and bromide are 115, 83.7, 72.1 kcal/mol respectively. [4] As a result of having weaker bond, vinyl iodide does not polymerize as easily as its vinyl halide counterparts, but rather decompose and release iodide. [5] It is generally believed that vinyl iodide cannot survive common reduction conditions, which reduces the vinyl iodide to an olefin or unsaturated alkane. [6] However, there is evidence in literature, in which a propargyl alcohol's alkyne was reduced in presence of a vinyl iodide using hydrogen over Pd/CaCO3 or Crabtree's catalyst. [7]

Parker's group reduction methods Figure 3 Parker group work.jpg
Parker's group reduction methods

Other applications

Scheme 1. Magnesium-halogen exchange Scheme 1..jpg
Scheme 1. Magnesium-halogen exchange

Besides using vinyl iodides as useful substrates in transition metal cross-coupling reaction, they can also undergo elimination with a strong base to give corresponding alkyne, and they can be converted to suitable vinyl Grignard reagents. Vinyl iodides are converted to Grignard reagents by magnesium-halogen exchange (see Scheme 1a). [8] The scope of this synthetic method is limited since it requires higher temperatures and longer reaction time, which affects functional group tolerance. However, vinyl iodide with electron withdrawing group can enhance rate of exchange(see Scheme 1b). [8] Also addition of lithium chloride helps enhance magnesium-halogen exchange (see Scheme 1c). It is predicted lithium chloride breaks up aggregates in organomagnesium reagents. [9]

Methods of synthesis

Nomenclature Stereo-regio isomers.jpg
Nomenclature

Vinyl iodides are synthesized by methods such as iodination and substitution reaction. Vinyl iodides with well-defined geometry (regiochemistry and stereochemistry) are important in synthesis since many natural products and drugs that have specific structure and dimension. Example of regiochemistry is whether the iodide is positioned in either alpha or beta position on the olefin. Stereochemistry such as E-Z notation or cis-trans alkene geometry is important since some transition metal cross-coupling reactions, such as the Suzuki coupling, can retain olefin geometry. In synthesis, it is useful to introduce vinyl iodide at various positions to be set up for a coupling reaction at the next synthetic step. Below are various means and methods in introducing and synthesizing vinyl iodides.

Synthesis from alkynes

The common and simplest approach to make vinyl iodide is addition of one equivalent HI to alkyne. This generally makes 2-iodo-1-alkenes or α-vinyl iodide by Markovnikov's rule. However, this reaction does not happen at good rates or very high stereoselectively. [10] As a result, most synthetic methods often involve a hydrometalation step before addition of I+ source.

α-vinyl iodides

Introducing an α-vinyl iodide from a terminal position of an alkyne is a difficult step. in addition, the vinyl metal intermediate can be mildly nucleophilic, for example vinyl aluminum, can form C-C bonds under catalytic conditions. However, Hoveyda group have demonstrated using nickel-based catalyst (Ni(dppp)Cl2), DIBAL-H with N-iodosuccinimide (NIS), selectively favor α-vinyl iodide with little to no byproducts. [11] Also they observed reverse selectivity for β with Ni(PPh3)2Cl2 in their hydroalumination reactions under same conditions with little or no byproducts. The advantage of this method is that is inexpensive (and commercially available), scalable and one-pot reaction.

Hoveyda'group hydroalumination method Figure5 hydroalumination.jpg
Hoveyda'group hydroalumination method

Another method doesn't involve hydrometalation but hydroiodation with I2/hydrophosphine binary system, which was developed by Ogawa's group. [12]

Ogawa's group Hydroiodation method with I2/hydrophosphine Hydroiodation with I2 and hydrophosphine.jpg
Ogawa's group Hydroiodation method with I2/hydrophosphine

The hydroiodation proceeds by Markovnikov-type adduct, no reaction is observed without addition of hydrophoshine. In a plausible mechanism proposed by Ogawa's group, the hydrophosphine reacts with HI to form an intermediate complex that coordinate HI to do Markovnikov hydroiodation on the alkene. The advantage of this system is the conditions are mild, can tolerate wide range of functional groups.

Mechanism Proposed by Ogawa's group Mechanism Proposed by Ogawa's group.jpg
Mechanism Proposed by Ogawa's group

β-vinyl iodides

They are generally more methods in making β-vinyl iodides versus α-vinyl iodides using hydrometalation (with aluminum with DIBAL-H (hydroalumination), with boron (hydroboration), with HZrCp2Cl (hydrozirconation)). [13] However, hydrometalation with alkyne with various functional groups often react poorly with side products. The Chong groups have demonstrated using hydrostannation, using Bu3SnH with palladium catalyst with high E stereoselectivity. [13] They observed using sterically bulky ligands gave higher regioselectivity for β-vinyl iodide. The advantage of this technique is this technique can tolerate a wide range of functional groups.

Chong's group hydrostannation method Chong's group hydrostannation method.jpg
Chong's group hydrostannation method

Z selective β-vinyl iodides are slightly more difficult to introduce than E-β-vinyl iodides, often requiring more than one step. Hydroalumination and hydroboration usually proceed by syn fashion, therefore selectively favors E geometry. The Oshima group have demonstrated using hydroindation with HInCl selectively favors Z geometry. [14] They suggested that the reaction proceeds by a radical mechanism. They predict that HInCl adds to alkyne by radical addition in a Z geometry. It does not isomerized to E geometry because of low reactivity of radical InCl2 with intermediate complex (no second addition). If second addition occurs then isomerization will occur through diindium intermediate. They confirm a radical mechanism in a mechanistic study with alkyne and alkene cyclization.

Oshima's group hydroindation method Oshima's group hydroindation method.jpg
Oshima's group hydroindation method

Substitution

Substitution is perhaps most useful method in introducing vinyl iodide into the molecule. Halogen-exchange can be useful since vinyl iodides are more reactivity than other vinyl halides. Buchwald group demonstrates a halogen-exchange from vinyl bromide to vinyl iodide with copper catalyst under mild conditions. [15] It is possible that this method can tolerate various functional groups since these conditions were tested aryl halides initially. The scope of this exchange for regiochemistry and stereochemistry is currently unexplored.

Buchwald's group halogen exchange method Buchwald's group halogen exchange method.jpg
Buchwald's group halogen exchange method

Halogen-exchange can also be done with zirconium derivatives that retain olefin’s geometry [16]

Marek's group Halogen exchange method Marek's group Halogen exchange method.jpg
Marek's group Halogen exchange method

The Marek group have further investigated using zirconium catalyst on E or Z vinyl ethers, which selective for E-vinyl ethers. [16] The zirconium's oxophilic nature allows elimination alkoxy group at the β position to form intermediate vinyl zirconium complex. The E geometry selectivity is not cause by sterics but rather the reaction itself is not concerted. In a mechanistic study, they observed isomerization, which suggest E geometry product is more favored than Z geometry. The difference of results between halogen exchange and E-vinyl ether reaction is that only when there is a presence of an oxonium intermediate, is isomerization observed.

Marek's group zirconium vinyl iodide synthesis from vinyl ether Marek's group zirconium vinyl iodide synthesis from vinyl ether.jpg
Marek's group zirconium vinyl iodide synthesis from vinyl ether
Scheme 2. Scheme 2 vinyl borate to viny iodide.jpg
Scheme 2.

An interesting substitution reaction is vinyl boronic acid to vinyl iodide done by Brown's group. [17] Depending on order of addition of iodide or base, vinyl borate can yield different stereoisomers of vinyl iodide (see scheme 2a). The Whiting group, however, noticed that Brown's method was not applicable to more sterically hindered boronic esters (no reaction). [18] They proposed that the iodide source was not electropositive enough. So they decided to use ICl which is more polar than I2, in which, they observed similar results (see scheme 2b).

Radical substitution of carboxylic acid to iodide is demonstrated by a modified Hunsdiecker reaction. [19] Homolytic cleavage of O-I bond generates CO2 and vinyl radical. Vinyl radical recombines with iodide radical to form vinyl iodide.

Modified Hunsdiecker reaction Modified Hunsdiecker reaction.jpg
Modified Hunsdiecker reaction

Iododesilylation

Iododesilylation is a substitution reaction of silyl group for iodide. The advantages of iododesilylation are that it avoids toxic tin reagent and intermediate vinyl silyl are stable, nontoxic and easily handled and stored. Vinyl silyl can be made from terminal alkyne or other methods.

The Kishi's group reported a mild preparation of vinyl iodide from vinyl silyl using NIS in mixture of acetonitrile and chloroacetonitrile. [20] They observed retention of olefin geometry in some vinyl silyl substrates while inversion in others. They reasoned that the R group's size had an effect on the geometry of the olefin. If the R group is small, the solvent acetonitrile can participate in the reaction leading to inversion of the olefin's geometry. If the R group is big, the solvent is unable to participate, leading to retention of olefin's geometry

Kishi's group iododesilylation method Kishi's group iododesilylation method.jpg
Kishi's group iododesilylation method

Zakarian's group then decided to run the reaction in HFIP, which gave high retention of olefin geometry. [21] They reasoned that HFIP is a low nucleophilicity solvent, unlike acetonitrile. In addition, they observed accelerated reaction rate because HFIP activate NIS by hydrogen bonding.

Zakarian's group HFIP iododesilylation method Zakarian's group HFIP iododesilylation method.jpg
Zakarian's group HFIP iododesilylation method

Unfortunately, iododesilylation under those conditions (above) can potentially yield multiple byproducts in highly functionalized molecules with oxygen functional groups. Vilarrasa and Costa's group hypothesized that radical reactions producing HI and I2 help facilitate cleavage in alcohol's protecting group and may add into other alkene bonds. [22] They experimented with the use of silver additives such as silver acetate and silver carbonate in which the silver can react with the excess iodide to form silver iodide. They achieved better conversions with these conditions.

Name reactions

Some famous vinyl iodide synthesis methods involve conversion of aldehyde or ketone to vinyl iodide. Barton's hydrazone iodination method involves addition of hydrazines to aldehyde or ketone to form hydrazone. Then the hydrazone is converted to vinyl iodide by addition of iodide and DBU. [23] [24] This method has been used in natural product synthesis of Taxol by Danishefsky [25] and Cortistatin A by Shair. [26] Another method is the Takai olefination which uses iodoform and chromium(II) chloride to make vinyl iodide from aldehyde with high stereoselectivity for E geometry. [27] For high stereoselectivity for Z geometry, Stork-Zhao olefination proceeds by Wittig-like reaction. High yields and Z stereoselectivity occurred at low temperature and at the presence of HMPA. [28]

Stork-Zhao Olefination Stork-Zhao Olefination.jpg
Stork-Zhao Olefination

Below is example of employing both Takai olefination and Stork-Zhao olefination in total synthesis of (+)-3-(E)- and (+)-3-(Z)-Pinnatifidenyne. [29]

Employment of Takai and Stork-Zhao olefination by Kim's group Employment of Takai and Stork-Zhao olefination by Kim's group.jpg
Employment of Takai and Stork-Zhao olefination by Kim's group

Elimination method

Vinyl iodides are rarely by made an elimination reaction of vicinal diiodide because it tends to decompose to alkene and iodide. [30] The Baker group have shown using decarboxylation, elimination can occur. [31]

Baker's group elimination method Baker's group elimination method.jpg
Baker's group elimination method

See also

Related Research Articles

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

In organic chemistry, an alkene is a hydrocarbon containing a carbon–carbon double bond. The double bond may be internal or in the terminal position. Terminal alkenes are also known as α-olefins.

In chemistry, an electrophile is a chemical species that forms bonds with nucleophiles by accepting an electron pair. Because electrophiles accept electrons, they are Lewis acids. Most electrophiles are positively charged, have an atom that carries a partial positive charge, or have an atom that does not have an octet of electrons.

The Heck reaction is the chemical reaction of an unsaturated halide with an alkene in the presence of a base and a palladium catalyst to form a substituted alkene. It is named after Tsutomu Mizoroki and Richard F. Heck. Heck was awarded the 2010 Nobel Prize in Chemistry, which he shared with Ei-ichi Negishi and Akira Suzuki, for the discovery and development of this reaction. This reaction was the first example of a carbon-carbon bond-forming reaction that followed a Pd(0)/Pd(II) catalytic cycle, the same catalytic cycle that is seen in other Pd(0)-catalyzed cross-coupling reactions. The Heck reaction is a way to substitute alkenes.

The Simmons–Smith reaction is an organic cheletropic reaction involving an organozinc carbenoid that reacts with an alkene to form a cyclopropane. It is named after Howard Ensign Simmons, Jr. and Ronald D. Smith. It uses a methylene free radical intermediate that is delivered to both carbons of the alkene simultaneously, therefore the configuration of the double bond is preserved in the product and the reaction is stereospecific.

Organopalladium chemistry is a branch of organometallic chemistry that deals with organic palladium compounds and their reactions. Palladium is often used as a catalyst in the reduction of alkenes and alkynes with hydrogen. This process involves the formation of a palladium-carbon covalent bond. Palladium is also prominent in carbon-carbon coupling reactions, as demonstrated in tandem reactions.

<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.

In organic chemistry, hydroboration refers to the addition of a hydrogen-boron bond to certain double and triple bonds involving carbon. This chemical reaction is useful in the organic synthesis of organic compounds.

In organic chemistry, a vinyl halide is a compound with the formula CH2=CHX (X = halide). The term vinyl is often used to describe any alkenyl group. For this reason, alkenyl halides with the formula RCH=CHX are sometimes called vinyl halides. From the perspective of applications, the dominant member of this class of compounds is vinyl chloride, which is produced on the scale of millions of tons per year as a precursor to polyvinyl chloride. Polyvinyl fluoride is another commercial product. Related compounds include vinylidene chloride and vinylidene fluoride.

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

Organozinc chemistry is the study of the physical properties, synthesis, and reactions of organozinc compounds, which are organometallic compounds that contain carbon (C) to zinc (Zn) chemical bonds.

A carbometallation is any reaction where a carbon-metal bond reacts with a carbon-carbon π-bond to produce a new carbon-carbon σ-bond and a carbon-metal σ-bond. The resulting carbon-metal bond can undergo further carbometallation reactions or it can be reacted with a variety of electrophiles including halogenating reagents, carbonyls, oxygen, and inorganic salts to produce different organometallic reagents. Carbometallations can be performed on alkynes and alkenes to form products with high geometric purity or enantioselectivity, respectively. Some metals prefer to give the anti-addition product with high selectivity and some yield the syn-addition product. The outcome of syn and anti- addition products is determined by the mechanism of the carbometallation.

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

Takai olefination in organic chemistry describes the organic reaction of an aldehyde with a diorganochromium compound to form an alkene. It is a name reaction, referencing Kazuhiko Takai, who first reported it in 1986. In the original reaction, the organochromium species is generated from iodoform or bromoform and an excess of chromium(II) chloride and the product is a vinyl halide. One main advantage of this reaction is the E-configuration of the double bond that is formed. According to the original report, existing alternatives such as the Wittig reaction only gave mixtures.

<span class="mw-page-title-main">Reductions with samarium(II) iodide</span>

Reductions with samarium(II) iodide involve the conversion of various classes of organic compounds into reduced products through the action of samarium(II) iodide, a mild one-electron reducing agent.

Desulfonylation reactions are chemical reactions leading to the removal of a sulfonyl group from organic compounds. As the sulfonyl functional group is electron-withdrawing, methods for cleaving the sulfur–carbon bonds of sulfones are typically reductive in nature. Olefination or replacement with hydrogen may be accomplished using reductive desulfonylation methods.

<span class="mw-page-title-main">Organotantalum chemistry</span> Chemistry of compounds containing a carbon-to-tantalum bond

Organotantalum chemistry is the chemistry of chemical compounds containing a carbon-to-tantalum chemical bond. A wide variety of compound have been reported, initially with cyclopentadienyl and CO ligands. Oxidation states vary from Ta(V) to Ta(-I).

Cobalt(II)–porphyrin catalysis is a process in which a Co(II) porphyrin complex acts as a catalyst, inducing and accelerating a chemical reaction.

The Mukaiyama hydration is an organic reaction involving formal addition of an equivalent of water across an olefin by the action of catalytic bis(acetylacetonato)cobalt(II) complex, phenylsilane and atmospheric oxygen to produce an alcohol with Markovnikov selectivity.

Clark Landis is an American chemist, whose research focuses on organic and inorganic chemistry. He is currently a Professor of Chemistry at the University of Wisconsin–Madison. He was awarded the ACS Award in Organometallic Chemistry in 2010, and is a fellow of the American Chemical Society and the American Association for the Advancement of Science.

In organic chemistry, hydrovinylation is the formal insertion of an alkene into the C-H bond of ethylene. The more general reaction, hydroalkenylation, is the formal insertion of an alkene into the C-H bond of any terminal alkene. The reaction is catalyzed by metal complexes. A representative reaction is the conversion of styrene and ethylene to 3-phenybutene:

A phosphetane is a 4-membered organophosphorus heterocycle. The parent phosphetane molecule, which has the formula C3H7P, is one atom larger than phosphiranes, one smaller than phospholes, and is the heavy-atom analogue of azetidines. The first known phosphetane synthesis was reported in 1957 by Kosolapoff and Struck, but the method was both inefficient and hard to reproduce, with yields rarely exceeding 1%. A far more efficient method was reported in 1962 by McBride, whose method allowed for the first studies into the physical and chemical properties of phosphetanes. Phosphetanes are a well understood class of molecules that have found broad applications as chemical building blocks, reagents for organic/inorganic synthesis, and ligands in coordination chemistry.

Chao-Jun "C.-J." Li, a Canadian chemist, is E. B. Eddy Professor of Chemistry and Canada Research Chair in Green Chemistry at McGill University, Montréal. He is known for his pioneering works in Green Solvent and Green Syntheses.

References

  1. Xie, Meihua; Wang, Jialiang; Zhang, Wei; Wang, Shaowu (2009-06-15). "Regio- and stereospecific synthesis of vinyl halides via carbozincation of acetylenic sulfones followed by halogenation". Journal of Organometallic Chemistry. 694 (14): 2258–2262. doi:10.1016/j.jorganchem.2009.03.006. ISSN   0022-328X.
  2. 1 2 Klein, David R. (2011-08-24). Organic Chemistry. Wiley. ISBN   978-1-118-13750-5.
  3. MEHTA, BHUPINDER; MEHTA, MANJU (2005-01-01). ORGANIC CHEMISTRY. PHI Learning. ISBN   978-81-203-2441-1.
  4. Blanksby, Stephen J.; Ellison, G. Barney (2003-04-01). "Bond Dissociation Energies of Organic Molecules". Accounts of Chemical Research. 36 (4): 255–263. doi:10.1021/ar020230d. ISSN   0001-4842.
  5. Herman, Jan A.; Roberge, Pierre (December 1962). "X‐ray induced polymerization of vinyl iodide in solution". Journal of Polymer Science. 62 (174). doi:10.1002/pol.1962.1206217444. ISSN   0022-3832.
  6. Zhang, Xing; Liu, Jun; Sun, Xue; Du, Yuguo (2013-02-04). "An efficient cis-reduction of alkyne to alkene in the presence of a vinyl iodide: stereoselective synthesis of the C22–C31 fragment of leiodolide A". Tetrahedron. 69 (5): 1553–1558. doi:10.1016/j.tet.2012.12.008. ISSN   0040-4020.
  7. Denton, Richard W.; Parker, Kathlyn A. (2009-07-02). "Functional Group Compatibility. Propargyl Alcohol Reduction in the Presence of a Vinyl Iodide". Organic Letters. 11 (13): 2722–2723. doi:10.1021/ol900927a. ISSN   1523-7060. PMC   2726658 . PMID   19476372.
  8. 1 2 Rottländer, Mario; Boymond, Laure; Cahiez, Gérard; Knochel, Paul (1999-02-01). "Stereoselective Preparation of Functionalized Alkenylmagnesium Reagents via an Iodine−Magnesium Exchange Reaction". The Journal of Organic Chemistry. 64 (4): 1080–1081. doi:10.1021/jo981941l. ISSN   0022-3263.
  9. Ren, Hongjun; Krasovskiy, Arkady; Knochel, Paul (2004-11-01). "Stereoselective Preparation of Functionalized Acyclic Alkenylmagnesium Reagents Using i- PrMgCl·LiCl". Organic Letters. 6 (23): 4215–4217. doi:10.1021/ol048363h. ISSN   1523-7060.
  10. Kropp, Paul J.; Crawford, Scott D. (June 1994). "Surface-Mediated Reactions. 4. Hydrohalogenation of Alkynes". The Journal of Organic Chemistry. 59 (11): 3102–3112. doi:10.1021/jo00090a031. ISSN   0022-3263.
  11. Gao, Fang; Hoveyda, Amir H. (2010-08-18). "α-Selective Ni-Catalyzed Hydroalumination of Aryl- and Alkyl-Substituted Terminal Alkynes: Practical Syntheses of Internal Vinyl Aluminums, Halides, or Boronates". Journal of the American Chemical Society. 132 (32): 10961–10963. doi:10.1021/ja104896b. ISSN   0002-7863. PMC   2921967 . PMID   20698643.
  12. Kawaguchi, Shin-ichi; Ogawa, Akiya (2010-05-07). "Highly Selective Hydroiodation of Alkynes Using an Iodine−Hydrophosphine Binary System". Organic Letters. 12 (9): 1893–1895. doi:10.1021/ol1005246. ISSN   1523-7060.
  13. 1 2 Darwish, Alla; Chong, J. Michael (2012-01-14). "Synthesis of E-vinyl iodides via Pd-catalyzed hydrostannation of terminal alkynes". Tetrahedron. 68 (2): 654–658. doi:10.1016/j.tet.2011.10.104. ISSN   0040-4020.
  14. Takami, Kazuaki; Mikami, Satoshi; Yorimitsu, Hideki; Shinokubo, Hiroshi; Oshima, Koichiro (2003-08-01). "Triethylborane-Mediated Hydrogallation and Hydroindation: Novel Access to Organogalliums and Organoindiums". The Journal of Organic Chemistry. 68 (17): 6627–6631. doi:10.1021/jo0344790. ISSN   0022-3263.Takami, Kazuaki, et al. "Triethylborane-mediated hydrogallation and hydroindation: Novel access to organogalliums and organoindiums." The Journal of Organic Chemistry 68.17 (2003): 6627-6631.
  15. Klapars, Artis; Buchwald, Stephen L. (2002-12-01). "Copper-Catalyzed Halogen Exchange in Aryl Halides: An Aromatic Finkelstein Reaction". Journal of the American Chemical Society. 124 (50): 14844–14845. doi:10.1021/ja028865v. ISSN   0002-7863.
  16. 1 2 Liard, Annie; Marek, Ilan (2000-10-01). "Stereoselective Preparation of E Vinyl Zirconium Derivatives from E or Z Enol Ethers". The Journal of Organic Chemistry. 65 (21): 7218–7220. doi:10.1021/jo005561n. ISSN   0022-3263.
  17. Brown, Herbert C.; Hamaoka, Tsutomu; Ravindran, N. (1973-10-23). "ChemInform Abstract: REACTION OF ALKENYLBORONIC ACIDS WITH IODINE UNDER THE INFLUENCE OF BASE, A SIMPLE PROCEDURE FOR THE STEREOSPECIFIC CONVERSION OF TERMINAL ALKYNES INTO TRANS‐1‐ALKENYL IODIDES VIA HYDROBORATION". Chemischer Informationsdienst. 4 (43). doi:10.1002/chin.197343207. ISSN   0009-2975.
  18. Stewart, Sarah K; Whiting, Andrew (1995-05-29). "Stereoselective synthesis of vinyl iodides from vinylboronate pinacol esters using ICI". Tetrahedron Letters. 36 (22): 3929–3932. doi:10.1016/0040-4039(95)00644-R. ISSN   0040-4039.
  19. Das, Jaya Prakash; Roy, Sujit (2002-11-01). "Catalytic Hunsdiecker Reaction of α,β-Unsaturated Carboxylic Acids: How Efficient Is the Catalyst?". The Journal of Organic Chemistry. 67 (22): 7861–7864. doi:10.1021/jo025868h. ISSN   0022-3263.
  20. Stamos, Dean P; Taylor, Andrew G; Kishi, Yoshito (1996-11-25). "A mild preparation of vinyliodides from vinylsilanes". Tetrahedron Letters. 37 (48): 8647–8650. doi:10.1016/S0040-4039(96)02000-X. ISSN   0040-4039.
  21. Ilardi, Elizabeth A.; Stivala, Craig E.; Zakarian, Armen (2008-05-01). "Hexafluoroisopropanol as a Unique Solvent for Stereoselective Iododesilylation of Vinylsilanes". Organic Letters. 10 (9): 1727–1730. doi:10.1021/ol800341z. ISSN   1523-7060.
  22. Sidera, Mireia; Costa, Anna M.; Vilarrasa, Jaume (2011-09-16). "Iododesilylation of TIPS-, TBDPS-, and TBS-Substituted Alkenes in Connection with the Synthesis of Amphidinolides B/D". Organic Letters. 13 (18): 4934–4937. doi:10.1021/ol2020187. ISSN   1523-7060.
  23. Barton, D. H. R.; O'Brien, R. E.; Sternhell, S. (1962-01-01). "88. A new reaction of hydrazones". Journal of the Chemical Society (0): 470–476. doi:10.1039/JR9620000470. ISSN   0368-1769.
  24. Barton, Derek H. R.; Bashiardes, George; Fourrey, Jean-Louis (1988-01-01). "Studies on the oxidation of hydrazones with iodine and with phenylselenenyl bromide in the presence of strong organic bases; an improved procedure for the synthesis of vinyl iodides and phenyl-vinyl selenides". Tetrahedron. 44 (1): 147–162. doi:10.1016/S0040-4020(01)85102-4. ISSN   0040-4020.
  25. Danishefsky, Samuel J.; Masters, John J.; Young, Wendy B.; Link, J. T.; Snyder, Lawrence B.; Magee, Thomas V.; Jung, David K.; Isaacs, Richard C. A.; Bornmann, William G.; Alaimo, Cheryl A.; Coburn, Craig A.; Di Grandi, Martin J. (1996-01-01). "Total Synthesis of Baccatin III and Taxol". Journal of the American Chemical Society. 118 (12): 2843–2859. doi:10.1021/ja952692a. ISSN   0002-7863.Danishefsky, Samuel J., et al. "Total synthesis of baccatin III and taxol." Journal of the American Chemical Society 118.12 (1996): 2843-2859
  26. Lee, Hong Myung; Nieto-Oberhuber, Cristina; Shair, Matthew D. (2008-12-17). "Enantioselective Synthesis of (+)-Cortistatin A, a Potent and Selective Inhibitor of Endothelial Cell Proliferation". Journal of the American Chemical Society. 130 (50): 16864–16866. doi:10.1021/ja8071918. ISSN   0002-7863.
  27. Takai, K.; Nitta, K.; Utimoto, K. (November 1986). "Simple and selective method for aldehydes (RCHO) -> (E)-haloalkenes (RCH:CHX) conversion by means of a haloform-chromous chloride system". Journal of the American Chemical Society. 108 (23): 7408–7410. doi:10.1021/ja00283a046. ISSN   0002-7863.
  28. Stork, Gilbert; Zhao, Kang (1989-01-01). "A stereoselective synthesis of (Z)-1-iodo-1-alkenes". Tetrahedron Letters. 30 (17): 2173–2174. doi:10.1016/S0040-4039(00)99640-0. ISSN   0040-4039.
  29. Kim, Hyoungsu; Choi, Won Jun; Jung, Jaeyoon; Kim, Sanghee; Kim, Deukjoon (2003-08-01). "Construction of Eight-Membered Ether Rings by Olefin Geometry-Dependent Internal Alkylation: First Asymmetric Total Syntheses of (+)-3-( E )- and (+)-3-( Z )-Pinnatifidenyne". Journal of the American Chemical Society. 125 (34): 10238–10240. doi:10.1021/ja035538u. ISSN   0002-7863.
  30. Katritzky, Alan R.; Meth-Cohn, Otto; Rees, Charles Wayne (1995-12-15). Comprehensive Organic Functional Group Transformations: Synthesis: carbon with one heteroatom attached by a single bond. Elsevier. ISBN   978-0-08-042323-4.
  31. Baker, Raymond; Castro, Jose L. (1990-01-01). "Total synthesis of (+)-macbecin I". Journal of the Chemical Society, Perkin Transactions 1 (1): 47–65. doi:10.1039/P19900000047. ISSN   1364-5463.