Jump to content

Alkene: Difference between revisions

From Wikipedia, the free encyclopedia
Content deleted Content added
 
(36 intermediate revisions by 18 users not shown)
Line 1: Line 1:
{{Short description|1=Hydrocarbon compound containing one or more C=C bonds}}
{{Short description|1=Hydrocarbon compound containing one or more C=C bonds}}
{{distinguish|alkane|alkyne}}
{{Distinguish|alkane|alkyne}}
{{Redirect|Olefin|the material|olefin fiber}}
{{Redirect|Olefin|the material|olefin fiber}}
{{Use dmy dates|date=August 2022}}
{{Use dmy dates|date=August 2022}}


[[Image:Ethylene-3D-vdW.png|thumb|right|120px|A 3D model of [[ethylene]], the simplest alkene.]]
[[Image:Ethylene-3D-vdW.png|thumb|right|120px|A 3D model of [[ethylene]], the simplest alkene]]


In [[organic chemistry]], an '''alkene''' is a [[hydrocarbon]] containing a [[carbon]]–carbon [[double bond]].<ref name="Wade">{{cite book | last = Wade | first = L.G. | title = Organic Chemistry | url = https://archive.org/details/organicchemistry00lgwa_203 | url-access = limited | publisher = Pearson [[Prentice Hall]] | date = 2006 |edition=6th| pages = [https://archive.org/details/organicchemistry00lgwa_203/page/n322 279] | isbn = 978-1-4058-5345-3 }}</ref> The double bond may be internal or in the terminal position. Terminal alkenes are also known as [[Alpha-olefin|α-olefins]].
In [[organic chemistry]], an '''alkene''', or '''olefin''', is a [[hydrocarbon]] containing a [[carbon]]–carbon [[double bond]].<ref name="Wade">{{cite book | last = Wade | first = L.G. | title = Organic Chemistry | url = https://archive.org/details/organicchemistry00lgwa_203 | url-access = limited | publisher = Pearson [[Prentice Hall]] | date = 2006 |edition=6th| pages = [https://archive.org/details/organicchemistry00lgwa_203/page/n322 279] | isbn = 978-1-4058-5345-3 }}</ref> The double bond may be internal or in the terminal position. Terminal alkenes are also known as [[Alpha-olefin|α-olefins]].


The [[International Union of Pure and Applied Chemistry]] (IUPAC) [[Preferred IUPAC name|recommends using]] the name "alkene" only for [[Open-chain compound|acyclic]] hydrocarbons with just one double bond; '''alkadiene''', '''alkatriene''', etc., or '''[[polyene]]''' for acyclic hydrocarbons with two or more double bonds; '''cycloalkene''', '''cycloalkadiene''', etc. for [[Cyclic compound|cyclic]] ones; and "olefin" for the general class – cyclic or acyclic, with one or more double bonds.<ref name=PAC1995.alkenes/><ref name=PAC1995.olefins/><ref name="PAC1995">{{cite journal| title=Glossary of Class Names of Organic Compounds and Reactive Intermediates Based on Structure (IUPAC Recommendations 1995) | last1 = Moss | first1 = G. P. | journal= [[Pure and Applied Chemistry]] | year=1995| volume=67 | pages=1307–1375| doi=10.1351/pac199567081307| last2=Smith| first2=P. A. S.| last3=Tavernier| first3=D.| issue=8–9 | s2cid = 95004254 | doi-access=free}}</ref>
The [[International Union of Pure and Applied Chemistry]] (IUPAC) [[Preferred IUPAC name|recommends using]] the name "alkene" only for [[Open-chain compound|acyclic]] hydrocarbons with just one double bond; '''alkadiene''', '''alkatriene''', etc., or '''[[polyene]]''' for acyclic hydrocarbons with two or more double bonds; '''cycloalkene''', '''cycloalkadiene''', etc. for [[Cyclic compound|cyclic]] ones; and "olefin" for the general class – cyclic or acyclic, with one or more double bonds.<ref name=PAC1995.alkenes/><ref name=PAC1995.olefins/><ref name="PAC1995">{{cite journal| title=Glossary of Class Names of Organic Compounds and Reactive Intermediates Based on Structure (IUPAC Recommendations 1995) | last1 = Moss | first1 = G. P. | journal= [[Pure and Applied Chemistry]] | year=1995| volume=67 | pages=1307–75| doi=10.1351/pac199567081307| last2=Smith| first2=P. A. S.| last3=Tavernier| first3=D.| issue=8–9 | s2cid = 95004254 | doi-access=free}}</ref>


Acyclic alkenes, with only one double bond and no other [[functional group]]s (also known as '''mono-enes''') form a [[homologous series]] of [[hydrocarbon]]s with the general formula {{chem2|C_{''n''}H_{2''n''} }} with ''n'' being 2 or more (which is two [[hydrogen]]s less than the corresponding [[alkane]]). When ''n'' is four or more, [[isomer]]s are possible, distinguished by the position and [[cis–trans isomerism|conformation]] of the double bond.
Acyclic alkenes, with only one double bond and no other [[functional group]]s (also known as '''mono-enes''') form a [[homologous series]] of [[hydrocarbon]]s with the general formula {{chem2|C_{''n''}H_{2''n''} }} with ''n'' being a >1 natural number (which is two [[hydrogen]]s less than the corresponding [[alkane]]). When ''n'' is four or more, [[isomer]]s are possible, distinguished by the position and [[cis–trans isomerism|conformation]] of the double bond.


Alkenes are generally colorless [[polarity (chemistry)|non-polar]] compounds, somewhat similar to alkanes but more reactive. The first few members of the series are gases or liquids at room temperature. The simplest alkene, [[ethylene]] ({{chem2|C2H4}}) (or "ethene" in the [[IUPAC name|IUPAC nomenclature]]) is the [[organic compound]] produced on the largest scale industrially.<ref name="cenews">{{cite journal |title=Production: Growth is the Norm |journal=Chemical and Engineering News |volume=84 |issue=28 |pages=59–236 |date=10 July 2006 |doi=10.1021/cen-v084n034.p059}}</ref>
Alkenes are generally colorless [[polarity (chemistry)|non-polar]] compounds, somewhat similar to alkanes but more reactive. The first few members of the series are gases or liquids at room temperature. The simplest alkene, [[ethylene]] ({{chem2|C2H4}}) (or "ethene" in the [[IUPAC name|IUPAC nomenclature]]) is the [[organic compound]] produced on the largest scale industrially.<ref name="cenews">{{cite journal |title=Production: Growth is the Norm |journal=Chemical and Engineering News |volume=84 |issue=28 |pages=59–236 |date=10 July 2006 |doi=10.1021/cen-v084n034.p059}}</ref>
Line 42: Line 42:
As predicted by the [[VSEPR theory|VSEPR]] model of [[electron]] pair repulsion, the [[molecular geometry]] of alkenes includes [[bond angle]]s about each carbon atom in a double bond of about 120°. The angle may vary because of [[steric strain]] introduced by [[nonbonded interactions]] between [[functional group]]s attached to the carbon atoms of the double bond. For example, the C–C–C bond angle in [[propylene]] is 123.9°.
As predicted by the [[VSEPR theory|VSEPR]] model of [[electron]] pair repulsion, the [[molecular geometry]] of alkenes includes [[bond angle]]s about each carbon atom in a double bond of about 120°. The angle may vary because of [[steric strain]] introduced by [[nonbonded interactions]] between [[functional group]]s attached to the carbon atoms of the double bond. For example, the C–C–C bond angle in [[propylene]] is 123.9°.


For bridged alkenes, [[Bredt's rule]] states that a double bond cannot occur at the bridgehead of a bridged ring system unless the rings are large enough.<ref name=Bansal /> Following Fawcett and defining ''S'' as the total number of non-bridgehead atoms in the rings,<ref>{{cite journal |title=Bredt's Rule of Double Bonds in Atomic-Bridged-Ring Structures |first=Frank S. |last=Fawcett |journal=[[Chem. Rev.]] |year=1950 |volume=47 |issue=2 |pages=219–274 |doi=10.1021/cr60147a003 |pmid=24538877}}</ref> bicyclic systems require ''S''&nbsp;≥&nbsp;7 for stability<ref name=Bansal>{{cite book |title=Organic Reaction Mechanisms |chapter=Bredt's Rule |pages=14–16 |first=Raj K. |last=Bansal |edition=3rd |year=1998 |publisher=[[McGraw-Hill Education]] |isbn=978-0-07-462083-0 |chapter-url=https://books.google.com/books?id=bga3xjLVCo0C&pg=PT29}}</ref> and tricyclic systems require ''S''&nbsp;≥&nbsp;11.<ref>{{Cite book |year=2010 |chapter=Bredt's Rule |title=Comprehensive Organic Name Reactions and Reagents |volume=116 |pages=525–528 |doi=10.1002/9780470638859.conrr116 |isbn=978-0-470-63885-9}}</ref>
For bridged alkenes, [[Bredt's rule]] states that a double bond cannot occur at the bridgehead of a bridged ring system unless the rings are large enough.<ref name=Bansal /> Following Fawcett and defining ''S'' as the total number of non-bridgehead atoms in the rings,<ref>{{cite journal |title=Bredt's Rule of Double Bonds in Atomic-Bridged-Ring Structures |first=Frank S. |last=Fawcett |journal=[[Chem. Rev.]] |year=1950 |volume=47 |issue=2 |pages=219–274 |doi=10.1021/cr60147a003 |pmid=24538877}}</ref> bicyclic systems require ''S''&nbsp;≥&nbsp;7 for stability<ref name=Bansal>{{cite book |title=Organic Reaction Mechanisms |chapter=Bredt's Rule |pages=14–16 |first=Raj K. |last=Bansal |edition=3rd |year=1998 |publisher=[[McGraw-Hill Education]] |isbn=978-0-07-462083-0 |chapter-url=https://books.google.com/books?id=bga3xjLVCo0C&pg=PT29}}</ref> and tricyclic systems require ''S''&nbsp;≥&nbsp;11.<ref>{{Cite book |year=2010 |chapter=Bredt's Rule |title=Comprehensive Organic Name Reactions and Reagents |volume=116 |pages=525–8 |doi=10.1002/9780470638859.conrr116 |isbn=978-0-470-63885-9}}</ref>


=== Isomerism ===
=== Isomerism ===
Line 61: Line 61:
Many of the physical properties of alkenes and [[alkane]]s are similar: they are colorless, nonpolar, and combustible. The [[physical state]] depends on [[molecular mass]]: like the corresponding saturated hydrocarbons, the simplest alkenes ([[ethylene]], [[propylene]], and [[butene]]) are gases at room temperature. Linear alkenes of approximately five to sixteen carbon atoms are liquids, and higher alkenes are waxy solids. The melting point of the solids also increases with increase in molecular mass.
Many of the physical properties of alkenes and [[alkane]]s are similar: they are colorless, nonpolar, and combustible. The [[physical state]] depends on [[molecular mass]]: like the corresponding saturated hydrocarbons, the simplest alkenes ([[ethylene]], [[propylene]], and [[butene]]) are gases at room temperature. Linear alkenes of approximately five to sixteen carbon atoms are liquids, and higher alkenes are waxy solids. The melting point of the solids also increases with increase in molecular mass.


Alkenes generally have stronger smells than their corresponding alkanes. Ethylene has a sweet and musty odor. Strained alkenes, in particular, like norbornene and [[Trans-Cyclooctene|''trans''-cyclooctene]] are known to have strong, unpleasant odors, a fact consistent with the stronger π complexes they form with metal ions including copper.<ref>{{Cite journal|last1=Duan|first1=Xufang|last2=Block|first2=Eric|last3=Li|first3=Zhen|last4=Connelly|first4=Timothy|last5=Zhang|first5=Jian|last6=Huang|first6=Zhimin|last7=Su|first7=Xubo|last8=Pan|first8=Yi|last9=Wu|first9=Lifang|date=2012-02-28|title=Crucial role of copper in detection of metal-coordinating odorants|journal=Proceedings of the National Academy of Sciences of the United States of America|volume=109|issue=9|pages=3492–3497|doi=10.1073/pnas.1111297109|issn=0027-8424|pmc=3295281|pmid=22328155|bibcode=2012PNAS..109.3492D|doi-access=free}}</ref>
Alkenes generally have stronger smells than their corresponding alkanes. Ethylene has a sweet and musty odor. Strained alkenes, in particular, like norbornene and [[Trans-Cyclooctene|''trans''-cyclooctene]] are known to have strong, unpleasant odors, a fact consistent with the stronger π complexes they form with metal ions including copper.<ref>{{Cite journal|last1=Duan|first1=Xufang|last2=Block|first2=Eric|last3=Li|first3=Zhen|last4=Connelly|first4=Timothy|last5=Zhang|first5=Jian|last6=Huang|first6=Zhimin|last7=Su|first7=Xubo|last8=Pan|first8=Yi|last9=Wu|first9=Lifang|date=2012-02-28|title=Crucial role of copper in detection of metal-coordinating odorants|journal=Proceedings of the National Academy of Sciences of the United States of America|volume=109|issue=9|pages=3492–7|doi=10.1073/pnas.1111297109 |pmc=3295281|pmid=22328155|bibcode=2012PNAS..109.3492D|doi-access=free}}</ref>


=== Boiling and melting points ===
=== Boiling and melting points ===
This is a list showing the boiling points and melting points of various saturated and unsaturated hydrocarbons.<ref name="alkene properties">{{cite web|url=https://chem.libretexts.org/Bookshelves/Organic_Chemistry/Supplemental_Modules_(Organic_Chemistry)/Alkenes/Properties_of_Alkenes/Physical_Properties_of_Alkenes|title=Physical Properties of Alkenes|last1=Nguyen|first1=Trung|last2=Clark|first2=Jim|date=April 23, 2019|access-date=May 27, 2019|website=Chemistry LibreTexts}}</ref><ref>{{cite web|url=http://chemistry.elmhurst.edu/vchembook/501hcboilingpts.html|title=BOILING POINTS AND STRUCTURES OF HYDROCARBONS|last=Ophardt|first=Charles|date=2003|access-date=May 27, 2019|website=Virtual Chembook}}</ref> This shows that the effect of saturation is relatively small.
Below is a list of the boiling and melting points of various alkenes with the corresponding alkane and alkyne analogues.<ref name="alkene properties">{{cite web|url=https://chem.libretexts.org/Bookshelves/Organic_Chemistry/Supplemental_Modules_(Organic_Chemistry)/Alkenes/Properties_of_Alkenes/Physical_Properties_of_Alkenes|title=Physical Properties of Alkenes|last1=Nguyen|first1=Trung|last2=Clark|first2=Jim|date=April 23, 2019|access-date=May 27, 2019|website=Chemistry LibreTexts}}</ref><ref>{{cite web|url=http://chemistry.elmhurst.edu/vchembook/501hcboilingpts.html|title=Boiling Points and Structures of Hydrocarbons |last=Ophardt|first=Charles|date=2003|access-date=May 27, 2019|website=Virtual Chembook}}</ref>
[[File:Benzol.JPG|thumb|A bottle of [[benzene]]. Benzene is a liquid at room temperature and pressure.]]
{| class="wikitable"
{| class="wikitable"
|+Melting and boiling points in [[Celsius|°C]]
|+Melting and boiling points in [[Celsius|°C]]
Line 77: Line 76:
|Name||ethane||ethylene||acetylene
|Name||ethane||ethylene||acetylene
|-
|-
|Melting point||−183||-169||-80.7
|Melting point||−183||−169||−80.7
|-
|-
|Boiling point||-89||-104||-84.7
|Boiling point||−89||−104||−84.7
|-
|-
| rowspan="3" |3
| rowspan="3" |3
|Name||propane||propylene||propyne
|Name||propane||propylene||propyne
|-
|-
|Melting point||-190||-185||-102.7
|Melting point||−190||−185||−102.7
|-
|-
|Boiling point||-42||-47||-23.2
|Boiling point||−42||−47||−23.2
|-
|-
| rowspan="3" |4
| rowspan="3" |4
|Name||butane||1-butene||1-butyne
|Name||butane||1-butene||1-butyne
|-
|-
|Melting point||-138||-185.3||-125.7
|Melting point||−138||−185.3||−125.7
|-
|-
|Boiling point||-0.5||-6.2||8.0
|Boiling point||−0.5||−6.2||8.0
|-
|-
| rowspan="3" |5
| rowspan="3" |5
|Name||pentane||1-pentene||1-pentyne
|Name||pentane||1-pentene||1-pentyne
|-
|-
|Melting point||-130||-165.2||-90.0
|Melting point||−130||−165.2||−90.0
|-
|-
|Boiling point||36||29.9||40.1
|Boiling point||36||29.9||40.1
|}
|}


Just like their [[saturated hydrocarbon|saturated]] counterparts, the unsaturated hydrocarbons are [[non-polar]]. The intermolecular forces between unsaturated hydrocarbon molecules are dominantly weak [[Van der Waals force]]s. The boiling points and melting points of unsaturated hydrocarbons are usually similar to their saturated counterparts with same number of carbons.

The melting and boiling points of unsaturated hydrocarbons compared to saturated ones are determined by two opposing factors. On the one hand, the strength of [[Van der Waals force]] depends on the number of electrons in a molecule. Unsaturated hydrocarbons have fewer electrons than saturated ones, so the boiling and melting points may decrease as [[intermolecular force]] decreases. On the other hand, the delocalized π electrons existing in the unsaturated hydrocarbons make the electrons flow more easily within one molecule, so temporary dipoles are easier to form. Thus, the Van der Waals force may also increase due to delocalization of electrons. It turns out that alkynes are more affected by electron delocalization and usually have higher boiling points than alkanes with the same number of carbon. Alkenes are more affected by number of electrons and have lower boiling points than alkanes.<ref name="alkene properties"/>

The boiling and melting points also depend on the stereochemistry, i.e. [[cis–trans isomerism]] alkenes, although the effect is subtle.

For longer chains of unsaturated hydrocarbons, the effects above still apply. In longer chains, the stereochemical "zig-zag" effect of unsaturated hydrocarbons becomes the dominant effect, so unsaturated long chain hydrocarbons usually have lower boiling and melting points.<ref>{{cite web|url=http://chemistry.elmhurst.edu/vchembook/551fattyacids.html|title=Fatty Acids|last=Ophardt|first=Charles|date=2003|access-date=May 29, 2019|website=Virtual Chembook}}</ref> The melting point differences between saturated and unsaturated [[fat]]s inside human body also leads to health issues.

Unsaturated hydrocarbons have low solubility in water.
=== Infrared spectroscopy ===
=== Infrared spectroscopy ===
The stretching of C=C bond will give an [[infrared|IR]] absorption peak at 1670–1600&nbsp;[[wave number|cm<sup>−1</sup>]], while the bending of C=C bond absorbs between 1000 and 650&nbsp;cm<sup>−1</sup> wavelength. The stretching of C≡C bond absorbs 2100–2140&nbsp;cm<sup>−1</sup>(monosubstituted) and 2190–2260&nbsp;cm<sup>−1</sup>(disubstituted).<ref>{{cite web|url=https://www.sigmaaldrich.com/technical-documents/articles/biology/ir-spectrum-table.html|title=IR Spectrum Table & Chart|author=<!--Not stated-->|website=Sigma-Aldrich|access-date=May 5, 2019}}</ref> The strength of these absorption peaks varies with the place and number of the double or triple bonds.
The stretching of C=C bond will give an [[infrared|IR]] absorption peak at 1670–1600&nbsp;[[wave number|cm<sup>−1</sup>]], while the bending of C=C bond absorbs between 1000 and 650&nbsp;cm<sup>−1</sup> wavelength.

Because of the [[delocalization|delocalized π electrons]] in [[aromatic]] groups, the bending of C=C bond in these groups usually absorbs between 1500 and 1700&nbsp;cm<sup>−1</sup>.<ref>{{cite web|url=https://webspectra.chem.ucla.edu/irtable.html|title=Table of IR Absorptions|last=Merlic|first=Craig A.|website=Webspectra|access-date=May 5, 2019}}</ref>


=== NMR spectroscopy ===
=== NMR spectroscopy ===
In <sup>1</sup>H [[NMR]] spectroscopy, the [[hydrogen]] bonded to the carbon adjacent to double bonds will give a [[chemical shift|δ<sub>H</sub>]] of 4.5–6.5&nbsp;[[parts per million|ppm]]. The double bond will also [[Electromagnetic shielding|deshield]] the hydrogen attached to the carbons adjacent to sp<sup>2</sup> carbons, and this generates δ<sub>H</sub>=1.6–2.&nbsp;ppm peaks.<ref>{{cite web|url=http://www2.ups.edu/faculty/hanson/Spectroscopy/NMR/HNMRshift.htm|title=Overview of Chemical Shifts in H-NMR|last=Hanson|first=John|website=ups.edu|access-date=May 5, 2019}}</ref> Since the π bondings will make cis/trans isomers, the unsaturated hydrocarbon isomers will appear differently due to different [[J-coupling]] effect. Cis [[Vicinal (chemistry)|vicinal]] hydrogens will have coupling constants in the range of 6–14&nbsp;[[Hz]], whereas the trans will have coupling constants of 11–18&nbsp;Hz.<ref name="NMR of Alkenes">{{cite web|url=https://chem.libretexts.org/Bookshelves/Organic_Chemistry/Supplemental_Modules_(Organic_Chemistry)/Alkenes/Properties_of_Alkenes/Nuclear_Magnetic_Resonance_(NMR)_of_Alkenes|author=<!--Not stated-->|title=Nuclear Magnetic Resonance (NMR) of Alkenes|website=Chemistry LibreTexts|date=April 23, 2019|access-date=May 5, 2019}}</ref>
In <sup>1</sup>H [[NMR]] spectroscopy, the [[hydrogen]] bonded to the carbon adjacent to double bonds will give a [[chemical shift|δ<sub>H</sub>]] of 4.5–6.5&nbsp;[[parts per million|ppm]]. The double bond will also [[Electromagnetic shielding|deshield]] the hydrogen attached to the carbons adjacent to sp<sup>2</sup> carbons, and this generates δ<sub>H</sub>=1.6–2.&nbsp;ppm peaks.<ref>{{cite web|url=http://www2.ups.edu/faculty/hanson/Spectroscopy/NMR/HNMRshift.htm|title=Overview of Chemical Shifts in H-NMR|last=Hanson|first=John|website=ups.edu|access-date=May 5, 2019}}</ref> Cis/trans isomers are distinguishable due to different [[J-coupling]] effect. Cis [[Vicinal (chemistry)|vicinal]] hydrogens will have coupling constants in the range of 6–14&nbsp;[[Hz]], whereas the trans will have coupling constants of 11–18&nbsp;Hz.<ref name="NMR of Alkenes">{{cite web|url=https://chem.libretexts.org/Bookshelves/Organic_Chemistry/Supplemental_Modules_(Organic_Chemistry)/Alkenes/Properties_of_Alkenes/Nuclear_Magnetic_Resonance_(NMR)_of_Alkenes|author=<!--Not stated-->|title=Nuclear Magnetic Resonance (NMR) of Alkenes|website=Chemistry LibreTexts|date=April 23, 2019|access-date=May 5, 2019}}</ref>


In <sup>13</sup>C NMR spectroscopy, compared to the saturated hydrocarbons, double bonds also deshield the carbons, making them have low field shift. C=C double bonds usually have chemical shift of about 100–170&nbsp;ppm.<ref name="NMR of Alkenes" />
In their <sup>13</sup>C NMR spectra of alkenes, double bonds also deshield the carbons, making them have low field shift. C=C double bonds usually have chemical shift of about 100–170&nbsp;ppm.<ref name="NMR of Alkenes" />


==Reactions==
=== Combustion ===
Like most other [[hydrocarbon]]s, alkenes [[combustion|combust]] to give carbon dioxide and water.
Alkenes are relatively stable compounds, but are more reactive than [[alkane]]s. Most reactions of alkenes involve additions to this pi bond, forming new [[sigma bond|single bonds]]. Alkenes serve as a feedstock for the [[petrochemical industry]] because they can participate in a wide variety of reactions, prominently polymerization and alkylation.


The combustion of alkenes release less energy than burning same [[molarity]] of saturated ones with same number of carbons.
Except for ethylene, alkenes have two sites of reactivity: the carbon–carbon pi-bond and the presence of [[allylic]] CH centers. The former dominates but the allylic sites are important too.
This trend can be clearly seen in the list of [[enthalpy of combustion|standard enthalpy of combustion]] of hydrocarbons.<ref>{{cite web|url=http://www2.ucdsb.on.ca/tiss/stretton/database/organic_thermo.htm|title=Organic Compounds: Physical and Thermochemical Data|website=ucdsb.on.ca|access-date=May 5, 2019}}</ref>
{| class="wikitable"
|+Combustion energies of various hydrocarbons
!Number of<br>carbons
!Substance
!Type
!Formula
!H<sub>c</sub><sup>ø</sup><br>(kJ/mol)
|-
| rowspan="3" |2
|[[ethane]]
|saturated
|C<sub>2</sub>H<sub>6</sub>
| −1559.7
|-
|[[ethylene]]
|unsaturated
|C<sub>2</sub>H<sub>4</sub>
| −1410.8
|-
|[[acetylene]]
|unsaturated
|C<sub>2</sub>H<sub>2</sub>
| −1300.8
|-
| rowspan="3" |3
|[[propane]]
|saturated
|CH<sub>3</sub>CH<sub>2</sub>CH<sub>3</sub>
| −2219.2
|-
|[[propene]]
|unsaturated
|CH<sub>3</sub>CH=CH<sub>2</sub>
| −2058.1
|-
|[[propyne]]
|unsaturated
|CH<sub>3</sub>C≡CH
| −1938.7
|-
| rowspan="3" |4
|[[butane]]
|saturated
|CH<sub>3</sub>CH<sub>2</sub>CH<sub>2</sub>CH<sub>3</sub>
| −2876.5
|-
|[[1-butene]]
|unsaturated
|CH<sub>2</sub>=CH−CH<sub>2</sub>CH<sub>3</sub>
| −2716.8
|-
|[[1-butyne]]
|unsaturated
|CH≡C-CH<sub>2</sub>CH<sub>3</sub>
| −2596.6
|}


==Reactions==
Alkenes are relatively stable compounds, but are more reactive than [[alkane]]s. Most reactions of alkenes involve additions to this pi bond, forming new [[sigma bond|single bonds]]. Alkenes serve as a feedstock for the [[petrochemical industry]] because they can participate in a wide variety of reactions, prominently polymerization and alkylation. Except for ethylene, alkenes have two sites of reactivity: the carbon–carbon pi-bond and the presence of [[allylic]] CH centers. The former dominates but the allylic sites are important too.


=== Addition to the unsaturated bonds ===
=== Addition to the unsaturated bonds ===
[[Image:Ear.png|400px|thumb|typical electrophilic addition reaction of [[ethylene]]]]
[[Image:Ear.png|400px|thumb|typical electrophilic addition reaction of [[ethylene]]]]
Double bonds are often the locus of reactions of unsaturated hydrocarbons.

[[Hydrogenation]] involves the addition of [[hydrogen|H<sub>2</sub>]] resulting in an alkane. The equation of hydrogenation of [[ethylene]] to form [[ethane]] is:
[[Hydrogenation]] involves the addition of [[hydrogen|H<sub>2</sub>]] resulting in an alkane. The equation of hydrogenation of [[ethylene]] to form [[ethane]] is:
:H<sub>2</sub>C=CH<sub>2</sub> + H<sub>2</sub>→H<sub>3</sub>C−CH<sub>3</sub>
:H<sub>2</sub>C=CH<sub>2</sub> + H<sub>2</sub>→H<sub>3</sub>C−CH<sub>3</sub>
Line 138: Line 183:
Similar to hydrogen, halogens added to double bonds.
Similar to hydrogen, halogens added to double bonds.
:H<sub>2</sub>C=CH<sub>2</sub> + Br<sub>2</sub>→H<sub>2</sub>CBr−CH<sub>2</sub>Br
:H<sub>2</sub>C=CH<sub>2</sub> + Br<sub>2</sub>→H<sub>2</sub>CBr−CH<sub>2</sub>Br
[[Halonium ion]]s are intermediates.
[[Halonium ion]]s are intermediates. These reactions do not require catalysts.


[[File:Biadamantylidene-bromonium-ion-from-xtal-1994-2D-skeletal.png|170px|thumb|Structure of a [[bromonium ion]]]]
[[File:Biadamantylidene-bromonium-ion-from-xtal-1994-2D-skeletal.png|170px|thumb|Structure of a [[bromonium ion]]]]


[[Bromine test]] is used to test the saturation of hydrocarbons.<ref>{{Cite book | title = The Systematic Identification of Organic Compounds | author = R.L. Shriner, C.K.F. Hermann, T.C. Morrill, D.Y. Curtin, and R.C. Fuson | publisher = John Wiley & Sons | date = 1997| isbn = 0-471-59748-1}}</ref> The bromine test can also be used as an indication of the [[degree of unsaturation]] for unsaturated hydrocarbons. [[Bromine number]] is defined as gram of bromine able to react with 100g of product.<ref>{{cite web|url=https://www.hach.com/asset-get.download.jsa?id=3980387617|website=Hach company|title=Bromine Number|access-date=May 5, 2019}}</ref> Similar as hydrogenation, the halogenation of bromine is also depend on the number of π bond. A higher bromine number indicates higher degree of unsaturation.
[[Bromine test]] is used to test the saturation of hydrocarbons.<ref>{{Cite book | title = The Systematic Identification of Organic Compounds | first1 = R.L. |last1=Shriner | first2 =C.K.F. |last2=Hermann | first3 =T.C. |last3=Morrill | first4 =D.Y. |last4=Curtin | first5 =R.C. |last5=Fuson | publisher = Wiley | date = 1997| isbn = 0-471-59748-1}}</ref> The bromine test can also be used as an indication of the [[degree of unsaturation]] for unsaturated hydrocarbons. [[Bromine number]] is defined as gram of bromine able to react with 100g of product.<ref>{{cite web|url=https://www.hach.com/asset-get.download.jsa?id=3980387617|website=Hach company|title=Bromine Number|access-date=May 5, 2019}}</ref> Similar as hydrogenation, the halogenation of bromine is also depend on the number of π bond. A higher bromine number indicates higher degree of unsaturation.


The π bonds of alkenes hydrocarbons are also susceptible to [[hydration reaction]]. The reaction usually involves [[strong acid]] as [[catalyst]].<ref>{{cite web|url=https://www.chemguide.co.uk/physical/catalysis/hydrate.html|title=The Mechanism for the Acid Catalysed Hydration of Ethene |last=Clark|first=Jim|website=Chemguide|date=November 2007|access-date=May 6, 2019}}</ref> The first step in hydration often involves formation of a [[carbocation]]. The net result of the reaction will be an [[Alcohol (chemistry)|alcohol]]. The reaction equation for hydration of ethylene is:
The π bonds of alkenes hydrocarbons are also susceptible to [[hydration reaction|hydration]]. The reaction usually involves [[strong acid]] as [[catalyst]].<ref>{{cite web|url=https://www.chemguide.co.uk/physical/catalysis/hydrate.html|title=The Mechanism for the Acid Catalysed Hydration of Ethene |last=Clark|first=Jim|website=Chemguide|date=November 2007|access-date=May 6, 2019}}</ref> The first step in hydration often involves formation of a [[carbocation]]. The net result of the reaction will be an [[Alcohol (chemistry)|alcohol]]. The reaction equation for hydration of ethylene is:
:H<sub>2</sub>C=CH<sub>2</sub> + H<sub>2</sub>O→{{coloredlink|black|ethanol|H<sub>3</sub>C-CH<sub>2</sub>OH}}
:H<sub>2</sub>C=CH<sub>2</sub> + H<sub>2</sub>O→{{coloredlink|black|ethanol|H<sub>3</sub>C-CH<sub>2</sub>OH}}


Line 155: Line 200:
{{Main|cycloaddition}}
{{Main|cycloaddition}}
[[File:Diels-Alder (1,3-butadiene + ethylene) red.svg|right|thumb|a Diels-Alder reaction]]
[[File:Diels-Alder (1,3-butadiene + ethylene) red.svg|right|thumb|a Diels-Alder reaction]]
Alkenes add to [[diene]]s to give [[cyclohexene]]s. This conversion is an example of a [[Diels-Alder reaction]]. Such reaction proceed with retention of stereochemistry. The rates are sensitive to electron-withdrawing or electron-donating substituents. When irradiated by UV-light, alkenes dimerize to give [[cyclobutane]]s.<ref>{{March6th}}</ref>

[[File:4+2 cycloaddition cyclopentadiene O2.svg|350px|center|alt=Generation of singlet oxygen and its [4+2]-cycloaddition with cyclopentadiene]]
[[File:4+2 cycloaddition cyclopentadiene O2.svg|350px|center|alt=Generation of singlet oxygen and its [4+2]-cycloaddition with cyclopentadiene]]
Alkenes add to [[diene]]s to give [[cyclohexene]]s. This conversion is an example of a [[Diels-Alder reaction]]. Such reaction proceed with retention of stereochemistry. The rates are sensitive to electron-withdrawing or electron-donating substituents. When irradiated by UV-light, alkenes dimerize to give [[cyclobutane]]s.<ref>{{March6th}}</ref> Another example is the [[Ene reaction#Singlet-oxygen ene reaction|Schenck ene reaction]], in which singlet oxygen reacts with an [[allyl]]ic structure to give a transposed allyl [[peroxide]]:

Another example is the [[Ene reaction#Singlet-oxygen ene reaction|Schenck ene reaction]], in which singlet oxygen reacts with an [[allyl]]ic structure to give a transposed allyl [[peroxide]]:


[[File:Schenck ene reaction.svg|200px|center|alt=Reaction of singlet oxygen with an allyl structure to give allyl peroxide]]
[[File:Schenck ene reaction.svg|200px|center|alt=Reaction of singlet oxygen with an allyl structure to give allyl peroxide]]
Line 165: Line 207:
==== Oxidation ====
==== Oxidation ====
Alkenes react with [[Peroxy acid|percarboxylic acids]] and even hydrogen peroxide to yield [[epoxide]]s:
Alkenes react with [[Peroxy acid|percarboxylic acids]] and even hydrogen peroxide to yield [[epoxide]]s:
:<chem>RCH=CH2 + RCO3H -> RCHOCH2 + RCO2H</chem>
:{{chem2|RCH\dCH2 + RCO3H -> RCHOCH2 + RCO2H}}


For ethylene, the [[epoxidation]] is conducted on a very large scale industrially using oxygen in the presence of catalysts:
For ethylene, the [[epoxidation]] is conducted on a very large scale industrially using oxygen in the presence of silver-based catalysts:
:<chem>C2H4 + 1/2 O2 -> C2H4O</chem>
:{{chem2|C2H4 + 1/2 O2 -> C2H4O}}


Alkenes react with ozone, leading to the scission of the double bond. The process is called [[ozonolysis]]. Often the reaction procedure includes a mild reductant, such as dimethylsulfide ({{chem2|SMe2}}):
Alkenes react with ozone, leading to the scission of the double bond. The process is called [[ozonolysis]]. Often the reaction procedure includes a mild reductant, such as dimethylsulfide ({{chem2|SMe2}}):
:<chem>RCH=CHR' + O3 + SMe2 -> RCHO + R'CHO + O=SMe2</chem>
:{{chem2|RCH\dCHR' + O3 + SMe2 -> RCHO + R'CHO + O\dSMe2}}
:<chem>R2C=CHR' + O3 -> R2CHO + R'CHO + O=SMe2</chem>
:{{chem2|R2C\dCHR' + O3 -> R2CHO + R'CHO + O\dSMe2}}


When treated with a hot concentrated, acidified solution of {{chem2|[[potassium permanganate|KMnO4]]}}, alkenes are cleaved to form [[ketone]]s and/or [[carboxylic acid]]s. The stoichiometry of the reaction is sensitive to conditions. This reaction and the ozonolysis can be used to determine the position of a double bond in an unknown alkene.
When treated with a hot concentrated, acidified solution of {{chem2|[[potassium permanganate|KMnO4]]}}, alkenes are cleaved to form [[ketone]]s and/or [[carboxylic acid]]s. The stoichiometry of the reaction is sensitive to conditions. This reaction and the ozonolysis can be used to determine the position of a double bond in an unknown alkene.
Line 180: Line 222:
This reaction is called [[dihydroxylation]].
This reaction is called [[dihydroxylation]].


In the presence of an appropriate [[photosensitiser]], such as [[methylene blue]] and light, alkenes can undergo reaction with reactive oxygen species generated by the photosensitiser, such as [[hydroxyl radical]]s, [[singlet oxygen]] or [[superoxide]] ion. Reactions of the excited sensitizer can involve electron or hydrogen transfer, usually with a reducing substrate (Type I reaction) or interaction with oxygen (Type II reaction).<ref>{{cite journal |last1=Baptista |first1=Maurício S. |last2=Cadet |first2=Jean |last3=Mascio |first3=Paolo Di |last4=Ghogare |first4=Ashwini A. |last5=Greer |first5=Alexander |last6=Hamblin |first6=Michael R. |last7=Lorente |first7=Carolina |last8=Nunez |first8=Silvia Cristina |last9=Ribeiro |first9=Martha Simões |last10=Thomas |first10=Andrés H. |last11=Vignoni |first11=Mariana |last12=Yoshimura |first12=Tania Mateus |title=Type I and Type II Photosensitized Oxidation Reactions: Guidelines and Mechanistic Pathways |journal=Photochemistry and Photobiology |date=2017 |volume=93 |issue=4 |pages=912–919 |doi=10.1111/php.12716|pmid=28084040 |pmc=5500392 |doi-access=free }}</ref> These various alternative processes and reactions can be controlled by choice of specific reaction conditions, leading to a wide range of products. A common example is the [4+2]-[[cycloaddition]] of singlet oxygen with a [[diene]] such as [[cyclopentadiene]] to yield an [[endoperoxide]]:
In the presence of an appropriate [[photosensitiser]], such as [[methylene blue]] and light, alkenes can undergo reaction with reactive oxygen species generated by the photosensitiser, such as [[hydroxyl radical]]s, [[singlet oxygen]] or [[superoxide]] ion. Reactions of the excited sensitizer can involve electron or hydrogen transfer, usually with a reducing substrate (Type I reaction) or interaction with oxygen (Type II reaction).<ref>{{cite journal |last1=Baptista |first1=Maurício S. |last2=Cadet |first2=Jean |last3=Mascio |first3=Paolo Di |last4=Ghogare |first4=Ashwini A. |last5=Greer |first5=Alexander |last6=Hamblin |first6=Michael R. |last7=Lorente |first7=Carolina |last8=Nunez |first8=Silvia Cristina |last9=Ribeiro |first9=Martha Simões |last10=Thomas |first10=Andrés H. |last11=Vignoni |first11=Mariana |last12=Yoshimura |first12=Tania Mateus |title=Type I and Type II Photosensitized Oxidation Reactions: Guidelines and Mechanistic Pathways |journal=Photochemistry and Photobiology |date=2017 |volume=93 |issue=4 |pages=912–9 |doi=10.1111/php.12716|pmid=28084040 |pmc=5500392 |doi-access=free }}</ref> These various alternative processes and reactions can be controlled by choice of specific reaction conditions, leading to a wide range of products. A common example is the [4+2]-[[cycloaddition]] of singlet oxygen with a [[diene]] such as [[cyclopentadiene]] to yield an [[endoperoxide]]:


===Polymerization===
===Polymerization===
{{main|polyolefin}}
{{main|polyolefin}}
Terminal alkenes are precursors to [[polymer]]s via processes termed [[polymerization]]. Some polymerizations are of great economic significance, as they generate as the plastics [[polyethylene]] and [[polypropylene]]. Polymers from alkene are usually referred to as ''[[polyolefin]]s'' although they contain no olefins. Polymerization can proceed via diverse mechanisms. [[Conjugated system|conjugated]] [[diene]]s such as [[buta-1,3-diene]] and [[isoprene]] (2-methylbuta-1,3-diene) also produce polymers, one example being natural rubber.
Terminal alkenes are precursors to [[polymer]]s via processes termed [[polymerization]]. Some polymerizations are of great economic significance, as they generate the plastics [[polyethylene]] and [[polypropylene]]. Polymers from alkene are usually referred to as ''[[polyolefin]]s'' although they contain no olefins. Polymerization can proceed via diverse mechanisms. [[Conjugated system|Conjugated]] [[diene]]s such as [[buta-1,3-diene]] and [[isoprene]] (2-methylbuta-1,3-diene) also produce polymers, one example being natural rubber.


=== Allylic substitution ===
=== Allylic substitution ===
Line 196: Line 238:
[[File:DCDmodel.png|thumb|The [[Dewar-Chatt-Duncanson model]] for alkene-metal bonding.]]
[[File:DCDmodel.png|thumb|The [[Dewar-Chatt-Duncanson model]] for alkene-metal bonding.]]
:[[Image:Ni(cod)2.png|thumb|right|220px|Structure of [[bis(cyclooctadiene)nickel(0)]], a metal–alkene complex]]
:[[Image:Ni(cod)2.png|thumb|right|220px|Structure of [[bis(cyclooctadiene)nickel(0)]], a metal–alkene complex]]
In [[transition metal alkene complex]]es, alkenes serve as ligands for metals.<ref>{{cite web|url=http://www.ilpi.com/organomet/alkene.html|title=Alkene Complexes|last=Toreki|first=Rob|date=March 31, 2015|access-date=May 29, 2019|website=Organometallic HyperTextbook}}</ref> In this case, the π electron density is donated{{clarify|date=September 2023}} to the metal d orbitals. The stronger the donation is, the stronger the [[back bonding]] from the metal d orbital to π* anti-bonding orbital of the alkene. This effect lowers the bond order of the alkene and increases the C-C [[bond length]]. One example is the complex {{chem2|PtCl3(C2H4)]-}}. These complexes are related to the mechanisms of metal-catalyzed reactions of unsaturated hydrocarbons.<ref name=JFH>{{cite book | title=Organotransition Metal Chemistry: From Bonding to Catalysis | publisher=University Science Books | author=Hartwig, John | year=2010 | location=New York | pages=1160 | isbn=978-1-938787-15-7}}</ref>
In [[transition metal alkene complex]]es, alkenes serve as ligands for metals.<ref>{{cite web|url=http://www.ilpi.com/organomet/alkene.html|title=Alkene Complexes|last=Toreki|first=Rob|date=March 31, 2015|access-date=May 29, 2019|website=Organometallic HyperTextbook}}</ref> In this case, the π electron density is donated{{clarify|date=September 2023}} to the metal d orbitals. The stronger the donation is, the stronger the [[back bonding]] from the metal d orbital to π* anti-bonding orbital of the alkene. This effect lowers the bond order of the alkene and increases the C-C [[bond length]]. One example is the complex {{chem2|PtCl3(C2H4)]-}}. These complexes are related to the mechanisms of metal-catalyzed reactions of unsaturated hydrocarbons.<ref name=JFH>{{cite book | title=Organotransition Metal Chemistry: From Bonding to Catalysis | publisher=University Science Books | last=Hartwig |first=John | year=2010 | location=New York | pages=1160 | isbn=978-1-938787-15-7}}</ref>


===Reaction overview===
===Reaction overview===
Line 255: Line 297:
|reagent: ozone
|reagent: ozone
|-
|-
|[[Olefin metathesis]]
| [[Olefin metathesis]]
| alkenes
| alkenes
| two alkenes rearrange to form two new alkenes
| two alkenes rearrange to form two new alkenes
|-
|-
|[[Diels–Alder reaction]]
| [[Diels–Alder reaction]]
| cyclohexenes
| cyclohexenes
| cycloaddition with a diene
| cycloaddition with a diene
|-
|-
|[[Pauson–Khand reaction]]
| [[Pauson–Khand reaction]]
| cyclopentenones
| cyclopentenones
| cycloaddition with an alkyne and CO
| cycloaddition with an alkyne and CO
|-
|-
|[[Hydroboration–oxidation]]
| [[Hydroboration–oxidation]]
| alcohols
| alcohols
| reagents: borane, then a peroxide
| reagents: borane, then a peroxide
|-
|-
|[[Oxymercuration-reduction]]
| [[Oxymercuration-reduction]]
| alcohols
| alcohols
| electrophilic addition of mercuric acetate, then reduction
| electrophilic addition of mercuric acetate, then reduction
Line 304: Line 346:


===Industrial methods===
===Industrial methods===
Alkenes are produced by hydrocarbon [[cracking (chemistry)|cracking]]. Raw materials are mostly [[natural gas condensate]] components (principally ethane and propane) in the US and Mideast and [[naphtha]] in Europe and Asia. Alkanes are broken apart at high temperatures, often in the presence of a [[zeolite]] catalyst, to produce a mixture of primarily aliphatic alkenes and lower molecular weight alkanes. The mixture is feedstock and temperature dependent, and separated by fractional distillation. This is mainly used for the manufacture of small alkenes (up to six carbons).<ref name="Wade2">{{cite book | last = Wade | first = L.G. | title = Organic Chemistry | url = https://archive.org/details/organicchemistry00wade_388 | url-access = limited | publisher = Pearson [[Prentice Hall]] | date = 2006 |edition=6th| pages = [https://archive.org/details/organicchemistry00wade_388/page/n351 309] | isbn = 978-1-4058-5345-3 }}</ref>
Alkenes are produced by hydrocarbon [[cracking (chemistry)|cracking]]. Raw materials are mostly [[natural-gas condensate]] components (principally ethane and propane) in the US and Mideast and [[naphtha]] in Europe and Asia. Alkanes are broken apart at high temperatures, often in the presence of a [[zeolite]] catalyst, to produce a mixture of primarily aliphatic alkenes and lower molecular weight alkanes. The mixture is feedstock and temperature dependent, and separated by fractional distillation. This is mainly used for the manufacture of small alkenes (up to six carbons).<ref name="Wade2">{{cite book | last = Wade | first = L.G. | title = Organic Chemistry | url = https://archive.org/details/organicchemistry00wade_388 | url-access = limited | publisher = Pearson [[Prentice Hall]] | date = 2006 |edition=6th| pages = [https://archive.org/details/organicchemistry00wade_388/page/n351 309] | isbn = 978-1-4058-5345-3 }}</ref>
[[Image:OctaneCracking.svg|500px|center|Cracking of ''n''-octane to give pentane and propene]]
[[Image:OctaneCracking.svg|500px|center|Cracking of ''n''-octane to give pentane and propene]]


Line 314: Line 356:


===Elimination reactions===
===Elimination reactions===
One of the principal methods for alkene synthesis in the laboratory is the room [[elimination reaction|elimination]] of alkyl halides, alcohols, and similar compounds. Most common is the β-elimination via the E2 or E1 mechanism,<ref name="PataiBook1964">{{cite book | last = Saunders | first = W. H. | editor = Patai, Saul | title = The Chemistry of Alkenes| publisher = Wiley Interscience | year = 1964 | pages = 149–150 }}</ref> but α-eliminations are also known.
One of the principal methods for alkene synthesis in the laboratory is the [[elimination reaction]] of alkyl halides, alcohols, and similar compounds. Most common is the β-elimination via the E2 or E1 mechanism.<ref name="PataiBook1964">{{cite book | last = Saunders | first = W. H. | editor = Patai, Saul | title = The Chemistry of Alkenes| chapter=Elimination Reactions in Solution|publisher = Wiley Interscience |series=PATAI'S Chemistry of Functional Groups | year = 1964 | pages = 149–201|doi=10.1002/9780470771044 | isbn = 978-0-470-77104-4 }}</ref> A commercially significant example is the production of [[vinyl chloride]].


The E2 mechanism provides a more reliable β-elimination method than E1 for most alkene syntheses. Most E2 eliminations start with an alkyl halide or alkyl sulfonate ester (such as a [[tosylate]] or [[triflate]]). When an alkyl halide is used, the reaction is called a [[dehydrohalogenation]]. For unsymmetrical products, the more substituted alkenes (those with fewer hydrogens attached to the C=C) tend to predominate (see [[Zaitsev's rule]]). Two common methods of elimination reactions are dehydrohalogenation of alkyl halides and dehydration of alcohols. A typical example is shown below; note that if possible, the H is ''anti'' to the leaving group, even though this leads to the less stable ''Z''-isomer.<ref name=Cram1956>{{cite journal
The E2 mechanism provides a more reliable β-elimination method than E1 for most alkene syntheses. Most E2 eliminations start with an alkyl halide or alkyl sulfonate ester (such as a [[tosylate]] or [[triflate]]). When an alkyl halide is used, the reaction is called a [[dehydrohalogenation]]. For unsymmetrical products, the more substituted alkenes (those with fewer hydrogens attached to the C=C) tend to predominate (see [[Zaitsev's rule]]). Two common methods of elimination reactions are dehydrohalogenation of alkyl halides and dehydration of alcohols. A typical example is shown below; note that if possible, the H is ''anti'' to the leaving group, even though this leads to the less stable ''Z''-isomer.<ref name=Cram1956>{{cite journal
| author = Cram, D.J.
| last1 = Cram |first1 = D.J.
| year = 1956
| year = 1956
| title = Studies in Stereochemistry. XXV. Eclipsing Effects in the E2 Reaction1
| title = Studies in Stereochemistry. XXV. Eclipsing Effects in the E2 Reaction1
Line 323: Line 365:
| volume = 78
| volume = 78
| issue = 4
| issue = 4
| pages = 790–796
| pages = 790–6
| doi = 10.1021/ja01585a024
| doi = 10.1021/ja01585a024
| last2 = Greene
| last2 = Greene
Line 336: Line 378:
:CH<sub>3</sub>CH<sub>2</sub>OH → H<sub>2</sub>C=CH<sub>2</sub> + H<sub>2</sub>O
:CH<sub>3</sub>CH<sub>2</sub>OH → H<sub>2</sub>C=CH<sub>2</sub> + H<sub>2</sub>O


An alcohol may also be converted to a better leaving group (e.g., [[xanthate]]), so as to allow a milder ''syn''-elimination such as the [[Chugaev elimination]] and the [[Grieco elimination]]. Related reactions include eliminations by β-haloethers (the [[Boord olefin synthesis]]) and esters ([[ester pyrolysis]]).
An alcohol may also be converted to a better leaving group (e.g., [[xanthate]]), so as to allow a milder ''syn''-elimination such as the [[Chugaev elimination]] and the [[Grieco elimination]]. Related reactions include eliminations by β-haloethers (the [[Boord olefin synthesis]]) and esters ([[ester pyrolysis]]). [[Diphosphorus tetraiodide]] will deoxygenate [[glycol]]s to alkenes.


Alkenes can be prepared indirectly from alkyl [[amine]]s. The amine or ammonia is not a suitable leaving group, so the amine is first either [[alkylation|alkylated]] (as in the [[Hofmann elimination]]) or oxidized to an [[amine oxide]] (the [[Cope reaction]]) to render a smooth elimination possible. The Cope reaction is a ''syn''-elimination that occurs at or below 150&nbsp;°C, for example:<ref name="CopeElimination1973">{{cite journal | author = Bach, R.D. | title=Mechanism of the Cope elimination | journal=J. Org. Chem. | year=1973| volume=38| pages=1742–3 | doi=10.1021/jo00949a029 | last2 = Andrzejewski | first2 = Denis | last3 = Dusold | first3 = Laurence R. | issue = 9 }}</ref>
Alkenes can be prepared indirectly from alkyl [[amine]]s. The amine or ammonia is not a suitable leaving group, so the amine is first either [[alkylation|alkylated]] (as in the [[Hofmann elimination]]) or oxidized to an [[amine oxide]] (the [[Cope reaction]]) to render a smooth elimination possible. The Cope reaction is a ''syn''-elimination that occurs at or below 150&nbsp;°C, for example:<ref name="CopeElimination1973">{{cite journal | last1 = Bach |first1=R.D. | title=Mechanism of the Cope elimination | journal=J. Org. Chem. | year=1973| volume=38| pages=1742–3 | doi=10.1021/jo00949a029 | last2 = Andrzejewski | first2 = Denis | last3 = Dusold | first3 = Laurence R. | issue = 9 }}</ref>


[[Image:CopeEliminationExample.svg|300px|center|Synthesis of cyclooctene via Cope elimination]]
[[Image:CopeEliminationExample.svg|300px|center|Synthesis of cyclooctene via Cope elimination]]
Line 347: Line 389:


===Synthesis from carbonyl compounds===
===Synthesis from carbonyl compounds===
Another important method for alkene synthesis involves construction of a new carbon–carbon double bond by coupling of a carbonyl compound (such as an [[aldehyde]] or [[ketone]]) to a [[carbanion]] equivalent. Such reactions are sometimes called ''olefinations''. The most well-known of these methods is the [[Wittig reaction]], but other related methods are known, including the [[Horner–Wadsworth–Emmons reaction]].
Another important class of methods for alkene synthesis involves construction of a new carbon–carbon double bond by coupling or condensation of a carbonyl compound (such as an [[aldehyde]] or [[ketone]]) to a [[carbanion]] or its equivalent. Pre-eminent is the [[aldol condensation]]. Knoevenagel condensations are a related class of reactions that convert carbonyls into alkenes.Well-known methods are called ''olefinations''. The [[Wittig reaction]] is illustrative, but other related methods are known, including the [[Horner–Wadsworth–Emmons reaction]].

The Wittig reaction involves reaction of an aldehyde or ketone with a [[Wittig reagent]] (or phosphorane) of the type Ph<sub>3</sub>P=CHR to produce an alkene and [[Triphenylphosphine oxide|Ph<sub>3</sub>P=O]]. The Wittig reagent is itself prepared easily from [[triphenylphosphine]] and an alkyl halide. The reaction is quite general and many functional groups are tolerated, even esters, as in this example:<ref name="Snider2006">{{cite journal| last1=Snider| first1=Barry B. | title=Synthesis of ent-Thallusin | journal=Org. Lett.| year=2006| volume=8| pages=2123–6|doi=10.1021/ol0605777| pmid=16671797| last2=Matsuo| first2=Y| last3=Snider| first3=BB| issue=10| pmc=2518398 }}</ref>
The Wittig reaction involves reaction of an aldehyde or ketone with a [[Wittig reagent]] (or phosphorane) of the type Ph<sub>3</sub>P=CHR to produce an alkene and [[Triphenylphosphine oxide|Ph<sub>3</sub>P=O]]. The Wittig reagent is itself prepared easily from [[triphenylphosphine]] and an alkyl halide.<ref>{{cite book | last = Crowell | first = Thomas I.|chapter=Alkene-Forming Condensation Reactions |series=PATAI'S Chemistry of Functional Groups| editor = Patai, Saul | title = The Chemistry of Alkenes| publisher = Wiley Interscience | year = 1964 | pages = 241–270|doi=10.1002/9780470771044.ch4 | isbn = 978-0-470-77104-4}}</ref>


[[Image:Wittig reaction example.svg|350px|center|A typical example of the Wittig reaction]]
[[Image:Wittig reaction example.svg|350px|center|A typical example of the Wittig reaction]]
Line 362: Line 404:
The formation of longer alkenes via the step-wise polymerisation of smaller ones is appealing, as [[ethylene]] (the smallest alkene) is both inexpensive and readily available, with hundreds of millions of tonnes produced annually. The [[Ziegler–Natta process]] allows for the formation of very long chains, for instance those used for [[polyethylene]]. Where shorter chains are wanted, as they for the production of [[surfactant]]s, then processes incorporating a [[olefin metathesis]] step, such as the [[Shell higher olefin process]] are important.
The formation of longer alkenes via the step-wise polymerisation of smaller ones is appealing, as [[ethylene]] (the smallest alkene) is both inexpensive and readily available, with hundreds of millions of tonnes produced annually. The [[Ziegler–Natta process]] allows for the formation of very long chains, for instance those used for [[polyethylene]]. Where shorter chains are wanted, as they for the production of [[surfactant]]s, then processes incorporating a [[olefin metathesis]] step, such as the [[Shell higher olefin process]] are important.
Olefin metathesis is also used commercially for the interconversion of ethylene and 2-butene to propylene. Rhenium- and molybdenum-containing [[heterogeneous catalysis]] are used in this process:<ref name=KO>{{cite encyclopedia|encyclopedia=Kirk-Othmer Encyclopedia of Chemical Technology |author=Lionel Delaude |author2=Alfred F. Noels|year=2005| doi=10.1002/0471238961.metanoel.a01|place=Weinheim|publisher=Wiley-VCH|isbn = 978-0-471-23896-6|chapter = Metathesis|pages=metanoel.a01 }}</ref>
Olefin metathesis is also used commercially for the interconversion of ethylene and 2-butene to propylene. Rhenium- and molybdenum-containing [[heterogeneous catalysis]] are used in this process:<ref name=KO>{{cite encyclopedia|encyclopedia=Kirk-Othmer Encyclopedia of Chemical Technology |first1=Lionel |last1=Delaude |first2=Alfred F. |last2=Noels|year=2005| doi=10.1002/0471238961.metanoel.a01|place=Weinheim|publisher=Wiley-VCH|isbn = 978-0-471-23896-6|chapter = Metathesis|pages=metanoel.a01 }}</ref>
:CH<sub>2</sub>=CH<sub>2</sub> + CH<sub>3</sub>CH=CHCH<sub>3</sub> → 2 CH<sub>2</sub>=CHCH<sub>3</sub>
:CH<sub>2</sub>=CH<sub>2</sub> + CH<sub>3</sub>CH=CHCH<sub>3</sub> → 2 CH<sub>2</sub>=CHCH<sub>3</sub>


Transition metal catalyzed [[hydrovinylation]] is another important alkene synthesis process starting from alkene itself.<ref name=Vogt2010>{{cite journal | author = Vogt, D. | year = 2010 | pages = 7166–8 | title = Cobalt-Catalyzed Asymmetric Hydrovinylation | issue = 40 | pmid = 20672269 | journal = Angew. Chem. Int. Ed. | volume = 49 | doi = 10.1002/anie.201003133 }}</ref> It involves the addition of a hydrogen and a vinyl group (or an alkenyl group) across a double bond.
Transition metal catalyzed [[hydrovinylation]] is another important alkene synthesis process starting from alkene itself.<ref name=Vogt2010>{{cite journal | last = Vogt |first=D. | year = 2010 | pages = 7166–8 | title = Cobalt-Catalyzed Asymmetric Hydrovinylation | issue = 40 | pmid = 20672269 | journal = Angew. Chem. Int. Ed. | volume = 49 | doi = 10.1002/anie.201003133 }}</ref> It involves the addition of a hydrogen and a vinyl group (or an alkenyl group) across a double bond.


===From alkynes===
===From alkynes===
Reduction of [[alkyne]]s is a useful method for the [[stereoselectivity|stereoselective]] synthesis of disubstituted alkenes. If the ''cis''-alkene is desired, [[hydrogenation]] in the presence of [[Lindlar's catalyst]] (a heterogeneous catalyst that consists of palladium deposited on calcium carbonate and treated with various forms of lead) is commonly used, though hydroboration followed by hydrolysis provides an alternative approach. Reduction of the alkyne by [[sodium]] metal in liquid [[ammonia]] gives the ''trans''-alkene.<ref name="ZweifelNantz">{{cite book | last = Zweifel | first = George S. |author2=Nantz, Michael H.| title = Modern Organic Synthesis: An Introduction | url = https://archive.org/details/modernorganicsyn00zwei | url-access = limited | publisher = W. H. Freeman & Co. | location = New York | year = 2007 | pages = [https://archive.org/details/modernorganicsyn00zwei/page/n373 366] | isbn = 978-0-7167-7266-8 }}</ref>
Reduction of [[alkyne]]s is a useful method for the [[stereoselectivity|stereoselective]] synthesis of disubstituted alkenes. If the ''cis''-alkene is desired, [[hydrogenation]] in the presence of [[Lindlar's catalyst]] (a heterogeneous catalyst that consists of palladium deposited on calcium carbonate and treated with various forms of lead) is commonly used, though hydroboration followed by hydrolysis provides an alternative approach. Reduction of the alkyne by [[sodium]] metal in liquid [[ammonia]] gives the ''trans''-alkene.<ref name="ZweifelNantz">{{cite book | last1 = Zweifel | first1 = George S. |last2=Nantz |first2=Michael H.| title = Modern Organic Synthesis: An Introduction | url = https://archive.org/details/modernorganicsyn00zwei | url-access = limited | publisher = W. H. Freeman | year = 2007 | pages = [https://archive.org/details/modernorganicsyn00zwei/page/n373 366] | isbn = 978-0-7167-7266-8 }}</ref>


[[Image:AlkyneToAlkene.png|600px|center|Synthesis of ''cis''- and ''trans''-alkenes from alkynes]]
[[Image:AlkyneToAlkene.png|600px|center|Synthesis of ''cis''- and ''trans''-alkenes from alkynes]]
Line 375: Line 417:


===Rearrangements and related reactions===
===Rearrangements and related reactions===
Alkenes can be synthesized from other alkenes via [[rearrangement reaction]]s. Besides [[olefin metathesis]] (described [[#Olefin metathesis|above]]), many [[pericyclic reaction]]s can be used such as the [[ene reaction]] and the [[Cope rearrangement]].
Alkenes can be synthesized from other alkenes via [[rearrangement reaction]]s. Besides [[olefin metathesis]] (described [[#Synthesis from alkenes|above]]), many [[pericyclic reaction]]s can be used such as the [[ene reaction]] and the [[Cope rearrangement]].


[[Image:3,3copeexpansion.svg|180px|center|Cope rearrangement of divinylcyclobutane to cyclooctadiene]]
[[Image:3,3copeexpansion.svg|180px|center|Cope rearrangement of divinylcyclobutane to cyclooctadiene]]
Line 412: Line 454:


==Natural occurrence==
==Natural occurrence==
Alkenes are pervasive in nature.
Alkenes are prevalent in nature.
Plants are the main natural source of alkenes in the form of [[terpene]]s. Many of the most vivid natural pigments are terpenes; e.g. [[lycopene]] (red in tomatoes) and [[carotene]] (orange of carrots). The simplest of all alkenes, [[ethylene|ethylene (plant hormone)]] is a [[signaling molecule]] that influences the ripening of plants.
Plants are the main natural source of alkenes in the form of [[terpene]]s.<ref name="Ninkuu_2021">{{cite journal |last1=Ninkuu |first1=Vincent |last2=Zhang |first2=Lin |last3=Yan |first3=Jianpei |last4=Fu |first4=Zenchao |last5=Yang |first5=Tengfeng |last6=Zeng |first6=Hongmei |display-authors=3 |date=June 2021 |title=Biochemistry of Terpenes and Recent Advances in Plant Protection |journal=International Journal of Molecular Sciences |volume=22 |issue=11 |pages=5710 |doi=10.3390/ijms22115710 |doi-access=free |pmid=34071919 |pmc=8199371 }}</ref> Many of the most vivid natural pigments are terpenes; e.g. [[lycopene]] (red in tomatoes), [[carotene]] (orange in carrots), and [[xanthophyll]]s (yellow in egg yolk). The simplest of all alkenes, [[ethylene (plant hormone)|ethylene]] is a [[signaling molecule]] that influences the ripening of plants.


<gallery caption="Selected unsaturated compounds in nature">
<gallery caption="Selected unsaturated compounds in nature">
Line 421: Line 463:
File:Squalene.svg|[[Squalene]], a [[triterpene]] and universal precursor to natural [[steroid]]s.
File:Squalene.svg|[[Squalene]], a [[triterpene]] and universal precursor to natural [[steroid]]s.
</gallery>
</gallery>



==IUPAC Nomenclature==
==IUPAC Nomenclature==
Line 449: Line 490:
[[Image:Cis-trans example.svg|thumb|300px|center|The difference between ''cis-'' and ''trans-'' isomers]]
[[Image:Cis-trans example.svg|thumb|300px|center|The difference between ''cis-'' and ''trans-'' isomers]]


More generally, ''cis''–''trans'' isomerism will exist if each of the two carbons of in the double bond has two different atoms or groups attached to it. Accounting for these cases, the IUPAC recommends the more general [[E–Z notation]], instead of the ''cis'' and ''trans'' prefixes. This notation considers the group with highest [[Cahn-Ingold-Prelog priority rule|CIP priority]] in each of the two carbons. If these two groups are on opposite sides of the double bond's plane, the configuration is labeled ''E'' (from the [[German language|German]] ''entgegen'' meaning "opposite"); if they are on the same side, it is labeled ''Z'' (from German ''zusammen'', "together"). This labeling may be taught with mnemonic "''Z'' means 'on ze zame zide'".<ref name=murr2014>John E. McMurry (2014): ''[https://books.google.com/books?id=KDIeCgAAQBAJ&pg=PA189 Organic Chemistry with Biological Applications]''; 3rd edition. 1224 pages. {{isbn|9781285842912}}</ref>
More generally, ''cis''–''trans'' isomerism will exist if each of the two carbons of in the double bond has two different atoms or groups attached to it. Accounting for these cases, the IUPAC recommends the more general [[E–Z notation]], instead of the ''cis'' and ''trans'' prefixes. This notation considers the group with highest [[Cahn-Ingold-Prelog priority rule|CIP priority]] in each of the two carbons. If these two groups are on opposite sides of the double bond's plane, the configuration is labeled ''E'' (from the [[German language|German]] ''entgegen'' meaning "opposite"); if they are on the same side, it is labeled ''Z'' (from German ''zusammen'', "together"). This labeling may be taught with mnemonic "''Z'' means 'on ze zame zide'".<ref name=murr2014>{{cite book |first=John E. |last=McMurry |date=2014 |page=[https://books.google.com/books?id=KDIeCgAAQBAJ&pg=PA189 189] |title=Organic Chemistry with Biological Applications |publisher=Cengage Learning |edition=3rd |isbn=978-1-285-84291-2}}</ref>


[[Image:EZalkenes2.png|400px|center|thumb|The difference between ''E'' and ''Z'' isomers]]
[[Image:EZalkenes2.png|400px|center|thumb|The difference between ''E'' and ''Z'' isomers]]
Line 455: Line 496:
===Groups containing C=C double bonds===
===Groups containing C=C double bonds===
IUPAC recognizes two names for hydrocarbon groups containing carbon–carbon double bonds, the [[vinyl group]] and the [[allyl]] group.<ref name="PAC1995"/>
IUPAC recognizes two names for hydrocarbon groups containing carbon–carbon double bonds, the [[vinyl group]] and the [[allyl]] group.<ref name="PAC1995"/>
[[Image:AlkeneGroups.png|200px|center]]
[[Image:AlkenylGroups.png|200px|center]]


==See also==
==See also==
Line 468: Line 509:


==Nomenclature links==
==Nomenclature links==
* Rule A-3. Unsaturated Compounds and Univalent Radicals [http://www.acdlabs.com/iupac/nomenclature/79/r79_53.htm] IUPAC Blue Book.
*[http://www.acdlabs.com/iupac/nomenclature/79/r79_53.htm Rule A-3. Unsaturated Compounds and Univalent Radicals] [[IUPAC Color Books#Blue Book|IUPAC Blue Book]].
* Rule A-4. Bivalent and Multivalent Radicals [http://www.acdlabs.com/iupac/nomenclature/79/r79_78.htm] IUPAC Blue Book.
*[http://www.acdlabs.com/iupac/nomenclature/79/r79_78.htm Rule A-4. Bivalent and Multivalent Radicals] IUPAC Blue Book.
* Rules A-11.3, A-11.4, A-11.5 Unsaturated monocyclic hydrocarbons and substituents [http://www.acdlabs.com/iupac/nomenclature/79/r79_60.htm] IUPAC Blue Book.
*[http://www.acdlabs.com/iupac/nomenclature/79/r79_60.htm Rules A-11.3, A-11.4, A-11.5 Unsaturated monocyclic hydrocarbons and substituents] IUPAC Blue Book.
* Rule A-23. Hydrogenated Compounds of Fused Polycyclic Hydrocarbons [http://www.acdlabs.com/iupac/nomenclature/79/r79_73.htm] IUPAC Blue Book.
*[http://www.acdlabs.com/iupac/nomenclature/79/r79_73.htm Rule A-23. Hydrogenated Compounds of Fused Polycyclic Hydrocarbons] IUPAC Blue Book.


== References ==
== References ==

Latest revision as of 15:51, 8 June 2024

A 3D model of ethylene, the simplest alkene

In organic chemistry, an alkene, or olefin, is a hydrocarbon containing a carbon–carbon double bond.[1] The double bond may be internal or in the terminal position. Terminal alkenes are also known as α-olefins.

The International Union of Pure and Applied Chemistry (IUPAC) recommends using the name "alkene" only for acyclic hydrocarbons with just one double bond; alkadiene, alkatriene, etc., or polyene for acyclic hydrocarbons with two or more double bonds; cycloalkene, cycloalkadiene, etc. for cyclic ones; and "olefin" for the general class – cyclic or acyclic, with one or more double bonds.[2][3][4]

Acyclic alkenes, with only one double bond and no other functional groups (also known as mono-enes) form a homologous series of hydrocarbons with the general formula CnH2n with n being a >1 natural number (which is two hydrogens less than the corresponding alkane). When n is four or more, isomers are possible, distinguished by the position and conformation of the double bond.

Alkenes are generally colorless non-polar compounds, somewhat similar to alkanes but more reactive. The first few members of the series are gases or liquids at room temperature. The simplest alkene, ethylene (C2H4) (or "ethene" in the IUPAC nomenclature) is the organic compound produced on the largest scale industrially.[5]

Aromatic compounds are often drawn as cyclic alkenes, however their structure and properties are sufficiently distinct that they are not classified as alkenes or olefins.[3] Hydrocarbons with two overlapping double bonds (C=C=C) are called allenes—the simplest such compound is itself called allene—and those with three or more overlapping bonds (C=C=C=C, C=C=C=C=C, etc.) are called cumulenes.

Structural isomerism[edit]

Alkenes having four or more carbon atoms can form diverse structural isomers. Most alkenes are also isomers of cycloalkanes. Acyclic alkene structural isomers with only one double bond follow:[6]

  • C2H4: ethylene only
  • C3H6: propylene only
  • C4H8: 3 isomers: 1-butene, 2-butene, and isobutylene
  • C5H10: 5 isomers: 1-pentene, 2-pentene, 2-methyl-1-butene, 3-methyl-1-butene, 2-methyl-2-butene
  • C6H12: 13 isomers: 1-hexene, 2-hexene, 3-hexene, 2-methyl-1-pentene, 3-methyl-1-pentene, 4-methyl-1-pentene, 2-methyl-2-pentene, 3-methyl-2-pentene, 4-methyl-2-pentene, 2,3-dimethyl-1-butene, 3,3-dimethyl-1-butene, 2,3-dimethyl-2-butene, 2-ethyl-1-butene

Many of these molecules exhibit cistrans isomerism. There may also be chiral carbon atoms particularly within the larger molecules (from C5). The number of potential isomers increases rapidly with additional carbon atoms.

Structure and bonding[edit]

Bonding[edit]

Ethylene (ethene), showing the pi bond in green

A carbon–carbon double bond consists of a sigma bond and a pi bond. This double bond is stronger than a single covalent bond (611 kJ/mol for C=C vs. 347 kJ/mol for C–C),[1] but not twice as strong. Double bonds are shorter than single bonds with an average bond length of 1.33 Å (133 pm) vs 1.53 Å for a typical C-C single bond.[7]

Each carbon atom of the double bond uses its three sp2 hybrid orbitals to form sigma bonds to three atoms (the other carbon atom and two hydrogen atoms). The unhybridized 2p atomic orbitals, which lie perpendicular to the plane created by the axes of the three sp2 hybrid orbitals, combine to form the pi bond. This bond lies outside the main C–C axis, with half of the bond on one side of the molecule and a half on the other. With a strength of 65 kcal/mol, the pi bond is significantly weaker than the sigma bond.

Rotation about the carbon–carbon double bond is restricted because it incurs an energetic cost to break the alignment of the p orbitals on the two carbon atoms. Consequently cis or trans isomers interconvert so slowly that they can be freely handled at ambient conditions without isomerization. More complex alkenes may be named with the EZ notation for molecules with three or four different substituents (side groups). For example, of the isomers of butene, the two methyl groups of (Z)-but-2-ene (a.k.a. cis-2-butene) appear on the same side of the double bond, and in (E)-but-2-ene (a.k.a. trans-2-butene) the methyl groups appear on opposite sides. These two isomers of butene have distinct properties.

Shape[edit]

As predicted by the VSEPR model of electron pair repulsion, the molecular geometry of alkenes includes bond angles about each carbon atom in a double bond of about 120°. The angle may vary because of steric strain introduced by nonbonded interactions between functional groups attached to the carbon atoms of the double bond. For example, the C–C–C bond angle in propylene is 123.9°.

For bridged alkenes, Bredt's rule states that a double bond cannot occur at the bridgehead of a bridged ring system unless the rings are large enough.[8] Following Fawcett and defining S as the total number of non-bridgehead atoms in the rings,[9] bicyclic systems require S ≥ 7 for stability[8] and tricyclic systems require S ≥ 11.[10]

Isomerism[edit]

In organic chemistry,the prefixes cis- and trans- are used to describe the positions of functional groups attached to carbon atoms joined by a double bond. In Latin, cis and trans mean "on this side of" and "on the other side of" respectively. Therefore, if the functional groups are both on the same side of the carbon chain, the bond is said to have cis- configuration, otherwise (i.e. the functional groups are on the opposite side of the carbon chain), the bond is said to have trans- configuration.

For there to be cis- and trans- configurations, there must be a carbon chain, or at least one functional group attached to each carbon is the same for both. E- and Z- configuration can be used instead in a more general case where all four functional groups attached to carbon atoms in a double bond are different. E- and Z- are abbreviations of German words zusammen (together) and entgegen (opposite). In E- and Z-isomerism, each functional group is assigned a priority based on the Cahn–Ingold–Prelog priority rules. If the two groups with higher priority are on the same side of the double bond, the bond is assigned Z- configuration, otherwise (i.e. the two groups with higher priority are on the opposite side of the double bond), the bond is assigned E- configuration. Cis- and trans- configurations do not have a fixed relationship with E- and Z-configurations.

Physical properties[edit]

Many of the physical properties of alkenes and alkanes are similar: they are colorless, nonpolar, and combustible. The physical state depends on molecular mass: like the corresponding saturated hydrocarbons, the simplest alkenes (ethylene, propylene, and butene) are gases at room temperature. Linear alkenes of approximately five to sixteen carbon atoms are liquids, and higher alkenes are waxy solids. The melting point of the solids also increases with increase in molecular mass.

Alkenes generally have stronger smells than their corresponding alkanes. Ethylene has a sweet and musty odor. Strained alkenes, in particular, like norbornene and trans-cyclooctene are known to have strong, unpleasant odors, a fact consistent with the stronger π complexes they form with metal ions including copper.[11]

Boiling and melting points[edit]

Below is a list of the boiling and melting points of various alkenes with the corresponding alkane and alkyne analogues.[12][13]

Melting and boiling points in °C
Number of
carbons
Alkane Alkene Alkyne
2 Name ethane ethylene acetylene
Melting point −183 −169 −80.7
Boiling point −89 −104 −84.7
3 Name propane propylene propyne
Melting point −190 −185 −102.7
Boiling point −42 −47 −23.2
4 Name butane 1-butene 1-butyne
Melting point −138 −185.3 −125.7
Boiling point −0.5 −6.2 8.0
5 Name pentane 1-pentene 1-pentyne
Melting point −130 −165.2 −90.0
Boiling point 36 29.9 40.1

Infrared spectroscopy[edit]

The stretching of C=C bond will give an IR absorption peak at 1670–1600 cm−1, while the bending of C=C bond absorbs between 1000 and 650 cm−1 wavelength.

NMR spectroscopy[edit]

In 1H NMR spectroscopy, the hydrogen bonded to the carbon adjacent to double bonds will give a δH of 4.5–6.5 ppm. The double bond will also deshield the hydrogen attached to the carbons adjacent to sp2 carbons, and this generates δH=1.6–2. ppm peaks.[14] Cis/trans isomers are distinguishable due to different J-coupling effect. Cis vicinal hydrogens will have coupling constants in the range of 6–14 Hz, whereas the trans will have coupling constants of 11–18 Hz.[15]

In their 13C NMR spectra of alkenes, double bonds also deshield the carbons, making them have low field shift. C=C double bonds usually have chemical shift of about 100–170 ppm.[15]

Combustion[edit]

Like most other hydrocarbons, alkenes combust to give carbon dioxide and water.

The combustion of alkenes release less energy than burning same molarity of saturated ones with same number of carbons. This trend can be clearly seen in the list of standard enthalpy of combustion of hydrocarbons.[16]

Combustion energies of various hydrocarbons
Number of
carbons
Substance Type Formula Hcø
(kJ/mol)
2 ethane saturated C2H6 −1559.7
ethylene unsaturated C2H4 −1410.8
acetylene unsaturated C2H2 −1300.8
3 propane saturated CH3CH2CH3 −2219.2
propene unsaturated CH3CH=CH2 −2058.1
propyne unsaturated CH3C≡CH −1938.7
4 butane saturated CH3CH2CH2CH3 −2876.5
1-butene unsaturated CH2=CH−CH2CH3 −2716.8
1-butyne unsaturated CH≡C-CH2CH3 −2596.6

Reactions[edit]

Alkenes are relatively stable compounds, but are more reactive than alkanes. Most reactions of alkenes involve additions to this pi bond, forming new single bonds. Alkenes serve as a feedstock for the petrochemical industry because they can participate in a wide variety of reactions, prominently polymerization and alkylation. Except for ethylene, alkenes have two sites of reactivity: the carbon–carbon pi-bond and the presence of allylic CH centers. The former dominates but the allylic sites are important too.

Addition to the unsaturated bonds[edit]

typical electrophilic addition reaction of ethylene

Hydrogenation involves the addition of H2 resulting in an alkane. The equation of hydrogenation of ethylene to form ethane is:

H2C=CH2 + H2→H3C−CH3

Hydrogenation reactions usually require catalysts to increase their reaction rate. The total number of hydrogens that can be added to an unsaturated hydrocarbon depends on its degree of unsaturation.

Similar to hydrogen, halogens added to double bonds.

H2C=CH2 + Br2→H2CBr−CH2Br

Halonium ions are intermediates. These reactions do not require catalysts.

Structure of a bromonium ion

Bromine test is used to test the saturation of hydrocarbons.[17] The bromine test can also be used as an indication of the degree of unsaturation for unsaturated hydrocarbons. Bromine number is defined as gram of bromine able to react with 100g of product.[18] Similar as hydrogenation, the halogenation of bromine is also depend on the number of π bond. A higher bromine number indicates higher degree of unsaturation.

The π bonds of alkenes hydrocarbons are also susceptible to hydration. The reaction usually involves strong acid as catalyst.[19] The first step in hydration often involves formation of a carbocation. The net result of the reaction will be an alcohol. The reaction equation for hydration of ethylene is:

H2C=CH2 + H2O→H3C-CH2OH
Example of hydrohalogenation: addition of HBr to an alkene

Hydrohalogenation involves addition of H−X to unsaturated hydrocarbons. This reaction results in new C−H and C−X σ bonds. The formation of the intermediate carbocation is selective and follows Markovnikov's rule. The hydrohalogenation of alkene will result in haloalkane. The reaction equation of HBr addition to ethylene is:

H2C=CH2 + HBr → H3C−CH2Br

Cycloaddition[edit]

a Diels-Alder reaction
Generation of singlet oxygen and its [4+2]-cycloaddition with cyclopentadiene

Alkenes add to dienes to give cyclohexenes. This conversion is an example of a Diels-Alder reaction. Such reaction proceed with retention of stereochemistry. The rates are sensitive to electron-withdrawing or electron-donating substituents. When irradiated by UV-light, alkenes dimerize to give cyclobutanes.[20] Another example is the Schenck ene reaction, in which singlet oxygen reacts with an allylic structure to give a transposed allyl peroxide:

Reaction of singlet oxygen with an allyl structure to give allyl peroxide

Oxidation[edit]

Alkenes react with percarboxylic acids and even hydrogen peroxide to yield epoxides:

RCH=CH2 + RCO3H → RCHOCH2 + RCO2H

For ethylene, the epoxidation is conducted on a very large scale industrially using oxygen in the presence of silver-based catalysts:

C2H4 + 1/2 O2 → C2H4O

Alkenes react with ozone, leading to the scission of the double bond. The process is called ozonolysis. Often the reaction procedure includes a mild reductant, such as dimethylsulfide (SMe2):

RCH=CHR' + O3 + SMe2 → RCHO + R'CHO + O=SMe2
R2C=CHR' + O3 → R2CHO + R'CHO + O=SMe2

When treated with a hot concentrated, acidified solution of KMnO4, alkenes are cleaved to form ketones and/or carboxylic acids. The stoichiometry of the reaction is sensitive to conditions. This reaction and the ozonolysis can be used to determine the position of a double bond in an unknown alkene.

The oxidation can be stopped at the vicinal diol rather than full cleavage of the alkene by using osmium tetroxide or other oxidants:

This reaction is called dihydroxylation.

In the presence of an appropriate photosensitiser, such as methylene blue and light, alkenes can undergo reaction with reactive oxygen species generated by the photosensitiser, such as hydroxyl radicals, singlet oxygen or superoxide ion. Reactions of the excited sensitizer can involve electron or hydrogen transfer, usually with a reducing substrate (Type I reaction) or interaction with oxygen (Type II reaction).[21] These various alternative processes and reactions can be controlled by choice of specific reaction conditions, leading to a wide range of products. A common example is the [4+2]-cycloaddition of singlet oxygen with a diene such as cyclopentadiene to yield an endoperoxide:

Polymerization[edit]

Terminal alkenes are precursors to polymers via processes termed polymerization. Some polymerizations are of great economic significance, as they generate the plastics polyethylene and polypropylene. Polymers from alkene are usually referred to as polyolefins although they contain no olefins. Polymerization can proceed via diverse mechanisms. Conjugated dienes such as buta-1,3-diene and isoprene (2-methylbuta-1,3-diene) also produce polymers, one example being natural rubber.

Allylic substitution[edit]

The presence of a C=C π bond in unsaturated hydrocarbons weakens the dissociation energy of the allylic C−H bonds. Thus, these groupings are susceptible to free radical substitution at these C-H sites as well as addition reactions at the C=C site. In the presence of radical initiators, allylic C-H bonds can be halogenated.[22] The presence of two C=C bonds flanking one methylene, i.e., doubly allylic, results in particularly weak HC-H bonds. The high reactivity of these situations is the basis for certain free radical reactions, manifested in the chemistry of drying oils.

Metathesis[edit]

Alkenes undergo olefin metathesis, which cleaves and interchanges the substituents of the alkene. A related reaction is ethenolysis:[23]

Metal complexation[edit]

The Dewar-Chatt-Duncanson model for alkene-metal bonding.
Structure of bis(cyclooctadiene)nickel(0), a metal–alkene complex

In transition metal alkene complexes, alkenes serve as ligands for metals.[24] In this case, the π electron density is donated[clarification needed] to the metal d orbitals. The stronger the donation is, the stronger the back bonding from the metal d orbital to π* anti-bonding orbital of the alkene. This effect lowers the bond order of the alkene and increases the C-C bond length. One example is the complex PtCl3(C2H4)]. These complexes are related to the mechanisms of metal-catalyzed reactions of unsaturated hydrocarbons.[23]

Reaction overview[edit]

Reaction name Product Comment
Hydrogenation alkanes addition of hydrogen
Hydroalkenylation alkenes hydrometalation / insertion / beta-elimination by metal catalyst
Halogen addition reaction 1,2-dihalide electrophilic addition of halogens
Hydrohalogenation (Markovnikov) haloalkanes addition of hydrohalic acids
Anti-Markovnikov hydrohalogenation haloalkanes free radicals mediated addition of hydrohalic acids
Hydroamination amines addition of N−H bond across C−C double bond
Hydroformylation aldehydes industrial process, addition of CO and H2
Hydrocarboxylation and Koch reaction carboxylic acid industrial process, addition of CO and H2O.
Carboalkoxylation ester industrial process, addition of CO and alcohol.
alkylation ester industrial process: alkene alkylating carboxylic acid with silicotungstic acid the catalyst.
Sharpless bishydroxylation diols oxidation, reagent: osmium tetroxide, chiral ligand
Woodward cis-hydroxylation diols oxidation, reagents: iodine, silver acetate
Ozonolysis aldehydes or ketones reagent: ozone
Olefin metathesis alkenes two alkenes rearrange to form two new alkenes
Diels–Alder reaction cyclohexenes cycloaddition with a diene
Pauson–Khand reaction cyclopentenones cycloaddition with an alkyne and CO
Hydroboration–oxidation alcohols reagents: borane, then a peroxide
Oxymercuration-reduction alcohols electrophilic addition of mercuric acetate, then reduction
Prins reaction 1,3-diols electrophilic addition with aldehyde or ketone
Paterno–Büchi reaction oxetanes photochemical reaction with aldehyde or ketone
Epoxidation epoxide electrophilic addition of a peroxide
Cyclopropanation cyclopropanes addition of carbenes or carbenoids
Hydroacylation ketones oxidative addition / reductive elimination by metal catalyst
Hydrophosphination phosphines

Synthesis[edit]

Industrial methods[edit]

Alkenes are produced by hydrocarbon cracking. Raw materials are mostly natural-gas condensate components (principally ethane and propane) in the US and Mideast and naphtha in Europe and Asia. Alkanes are broken apart at high temperatures, often in the presence of a zeolite catalyst, to produce a mixture of primarily aliphatic alkenes and lower molecular weight alkanes. The mixture is feedstock and temperature dependent, and separated by fractional distillation. This is mainly used for the manufacture of small alkenes (up to six carbons).[25]

Cracking of n-octane to give pentane and propene
Cracking of n-octane to give pentane and propene

Related to this is catalytic dehydrogenation, where an alkane loses hydrogen at high temperatures to produce a corresponding alkene.[1] This is the reverse of the catalytic hydrogenation of alkenes.

Dehydrogenation of butane to give butadiene and isomers of butene
Dehydrogenation of butane to give butadiene and isomers of butene

This process is also known as reforming. Both processes are endothermic and are driven towards the alkene at high temperatures by entropy.

Catalytic synthesis of higher α-alkenes (of the type RCH=CH2) can also be achieved by a reaction of ethylene with the organometallic compound triethylaluminium in the presence of nickel, cobalt, or platinum.

Elimination reactions[edit]

One of the principal methods for alkene synthesis in the laboratory is the elimination reaction of alkyl halides, alcohols, and similar compounds. Most common is the β-elimination via the E2 or E1 mechanism.[26] A commercially significant example is the production of vinyl chloride.

The E2 mechanism provides a more reliable β-elimination method than E1 for most alkene syntheses. Most E2 eliminations start with an alkyl halide or alkyl sulfonate ester (such as a tosylate or triflate). When an alkyl halide is used, the reaction is called a dehydrohalogenation. For unsymmetrical products, the more substituted alkenes (those with fewer hydrogens attached to the C=C) tend to predominate (see Zaitsev's rule). Two common methods of elimination reactions are dehydrohalogenation of alkyl halides and dehydration of alcohols. A typical example is shown below; note that if possible, the H is anti to the leaving group, even though this leads to the less stable Z-isomer.[27]

An example of an E2 Elimination
An example of an E2 Elimination

Alkenes can be synthesized from alcohols via dehydration, in which case water is lost via the E1 mechanism. For example, the dehydration of ethanol produces ethylene:

CH3CH2OH → H2C=CH2 + H2O

An alcohol may also be converted to a better leaving group (e.g., xanthate), so as to allow a milder syn-elimination such as the Chugaev elimination and the Grieco elimination. Related reactions include eliminations by β-haloethers (the Boord olefin synthesis) and esters (ester pyrolysis). Diphosphorus tetraiodide will deoxygenate glycols to alkenes.

Alkenes can be prepared indirectly from alkyl amines. The amine or ammonia is not a suitable leaving group, so the amine is first either alkylated (as in the Hofmann elimination) or oxidized to an amine oxide (the Cope reaction) to render a smooth elimination possible. The Cope reaction is a syn-elimination that occurs at or below 150 °C, for example:[28]

Synthesis of cyclooctene via Cope elimination
Synthesis of cyclooctene via Cope elimination

The Hofmann elimination is unusual in that the less substituted (non-Zaitsev) alkene is usually the major product.

Alkenes are generated from α-halosulfones in the Ramberg–Bäcklund reaction, via a three-membered ring sulfone intermediate.

Synthesis from carbonyl compounds[edit]

Another important class of methods for alkene synthesis involves construction of a new carbon–carbon double bond by coupling or condensation of a carbonyl compound (such as an aldehyde or ketone) to a carbanion or its equivalent. Pre-eminent is the aldol condensation. Knoevenagel condensations are a related class of reactions that convert carbonyls into alkenes.Well-known methods are called olefinations. The Wittig reaction is illustrative, but other related methods are known, including the Horner–Wadsworth–Emmons reaction.

The Wittig reaction involves reaction of an aldehyde or ketone with a Wittig reagent (or phosphorane) of the type Ph3P=CHR to produce an alkene and Ph3P=O. The Wittig reagent is itself prepared easily from triphenylphosphine and an alkyl halide.[29]

A typical example of the Wittig reaction
A typical example of the Wittig reaction

Related to the Wittig reaction is the Peterson olefination, which uses silicon-based reagents in place of the phosphorane. This reaction allows for the selection of E- or Z-products. If an E-product is desired, another alternative is the Julia olefination, which uses the carbanion generated from a phenyl sulfone. The Takai olefination based on an organochromium intermediate also delivers E-products. A titanium compound, Tebbe's reagent, is useful for the synthesis of methylene compounds; in this case, even esters and amides react.

A pair of ketones or aldehydes can be deoxygenated to generate an alkene. Symmetrical alkenes can be prepared from a single aldehyde or ketone coupling with itself, using titanium metal reduction (the McMurry reaction). If different ketones are to be coupled, a more complicated method is required, such as the Barton–Kellogg reaction.

A single ketone can also be converted to the corresponding alkene via its tosylhydrazone, using sodium methoxide (the Bamford–Stevens reaction) or an alkyllithium (the Shapiro reaction).

Synthesis from alkenes[edit]

The formation of longer alkenes via the step-wise polymerisation of smaller ones is appealing, as ethylene (the smallest alkene) is both inexpensive and readily available, with hundreds of millions of tonnes produced annually. The Ziegler–Natta process allows for the formation of very long chains, for instance those used for polyethylene. Where shorter chains are wanted, as they for the production of surfactants, then processes incorporating a olefin metathesis step, such as the Shell higher olefin process are important.

Olefin metathesis is also used commercially for the interconversion of ethylene and 2-butene to propylene. Rhenium- and molybdenum-containing heterogeneous catalysis are used in this process:[30]

CH2=CH2 + CH3CH=CHCH3 → 2 CH2=CHCH3

Transition metal catalyzed hydrovinylation is another important alkene synthesis process starting from alkene itself.[31] It involves the addition of a hydrogen and a vinyl group (or an alkenyl group) across a double bond.

From alkynes[edit]

Reduction of alkynes is a useful method for the stereoselective synthesis of disubstituted alkenes. If the cis-alkene is desired, hydrogenation in the presence of Lindlar's catalyst (a heterogeneous catalyst that consists of palladium deposited on calcium carbonate and treated with various forms of lead) is commonly used, though hydroboration followed by hydrolysis provides an alternative approach. Reduction of the alkyne by sodium metal in liquid ammonia gives the trans-alkene.[32]

Synthesis of cis- and trans-alkenes from alkynes
Synthesis of cis- and trans-alkenes from alkynes

For the preparation multisubstituted alkenes, carbometalation of alkynes can give rise to a large variety of alkene derivatives.

Rearrangements and related reactions[edit]

Alkenes can be synthesized from other alkenes via rearrangement reactions. Besides olefin metathesis (described above), many pericyclic reactions can be used such as the ene reaction and the Cope rearrangement.

Cope rearrangement of divinylcyclobutane to cyclooctadiene
Cope rearrangement of divinylcyclobutane to cyclooctadiene

In the Diels–Alder reaction, a cyclohexene derivative is prepared from a diene and a reactive or electron-deficient alkene.


Application[edit]

Unsaturated hydrocarbons are widely used to produce plastics, medicines, and other useful materials.

Name Structure Use
Ethylene
  • Monomers for synthesizing polyethylene
1,3-butadiene
vinyl chloride
  • Precursor to PVC
styrene

Natural occurrence[edit]

Alkenes are prevalent in nature. Plants are the main natural source of alkenes in the form of terpenes.[33] Many of the most vivid natural pigments are terpenes; e.g. lycopene (red in tomatoes), carotene (orange in carrots), and xanthophylls (yellow in egg yolk). The simplest of all alkenes, ethylene is a signaling molecule that influences the ripening of plants.

IUPAC Nomenclature[edit]

Although the nomenclature is not followed widely, according to IUPAC, an alkene is an acyclic hydrocarbon with just one double bond between carbon atoms.[2] Olefins comprise a larger collection of cyclic and acyclic alkenes as well as dienes and polyenes.[3]

To form the root of the IUPAC names for straight-chain alkenes, change the -an- infix of the parent to -en-. For example, CH3-CH3 is the alkane ethANe. The name of CH2=CH2 is therefore ethENe.

For straight-chain alkenes with 4 or more carbon atoms, that name does not completely identify the compound. For those cases, and for branched acyclic alkenes, the following rules apply:

  1. Find the longest carbon chain in the molecule. If that chain does not contain the double bond, name the compound according to the alkane naming rules. Otherwise:
  2. Number the carbons in that chain starting from the end that is closest to the double bond.
  3. Define the location k of the double bond as being the number of its first carbon.
  4. Name the side groups (other than hydrogen) according to the appropriate rules.
  5. Define the position of each side group as the number of the chain carbon it is attached to.
  6. Write the position and name of each side group.
  7. Write the names of the alkane with the same chain, replacing the "-ane" suffix by "k-ene".

The position of the double bond is often inserted before the name of the chain (e.g. "2-pentene"), rather than before the suffix ("pent-2-ene").

The positions need not be indicated if they are unique. Note that the double bond may imply a different chain numbering than that used for the corresponding alkane: (H
3
C)
3
C–CH
2
CH
3
is "2,2-dimethyl pentane", whereas (H
3
C)
3
C–CH=CH
2
is "3,3-dimethyl 1-pentene".

More complex rules apply for polyenes and cycloalkenes.[4]

Naming substituted hex-1-enes

Cistrans isomerism[edit]

If the double bond of an acyclic mono-ene is not the first bond of the chain, the name as constructed above still does not completely identify the compound, because of cistrans isomerism. Then one must specify whether the two single C–C bonds adjacent to the double bond are on the same side of its plane, or on opposite sides. For monoalkenes, the configuration is often indicated by the prefixes cis- (from Latin "on this side of") or trans- ("across", "on the other side of") before the name, respectively; as in cis-2-pentene or trans-2-butene.

The difference between cis- and trans- isomers

More generally, cistrans isomerism will exist if each of the two carbons of in the double bond has two different atoms or groups attached to it. Accounting for these cases, the IUPAC recommends the more general E–Z notation, instead of the cis and trans prefixes. This notation considers the group with highest CIP priority in each of the two carbons. If these two groups are on opposite sides of the double bond's plane, the configuration is labeled E (from the German entgegen meaning "opposite"); if they are on the same side, it is labeled Z (from German zusammen, "together"). This labeling may be taught with mnemonic "Z means 'on ze zame zide'".[34]

The difference between E and Z isomers

Groups containing C=C double bonds[edit]

IUPAC recognizes two names for hydrocarbon groups containing carbon–carbon double bonds, the vinyl group and the allyl group.[4]

See also[edit]

Nomenclature links[edit]

References[edit]

  1. ^ a b c Wade, L.G. (2006). Organic Chemistry (6th ed.). Pearson Prentice Hall. pp. 279. ISBN 978-1-4058-5345-3.
  2. ^ a b IUPAC, Compendium of Chemical Terminology, 2nd ed. (the "Gold Book") (1997). Online corrected version: (2006–) "alkenes". doi:10.1351/goldbook.A00224
  3. ^ a b c IUPAC, Compendium of Chemical Terminology, 2nd ed. (the "Gold Book") (1997). Online corrected version: (2006–) "olefins". doi:10.1351/goldbook.O04281
  4. ^ a b c Moss, G. P.; Smith, P. A. S.; Tavernier, D. (1995). "Glossary of Class Names of Organic Compounds and Reactive Intermediates Based on Structure (IUPAC Recommendations 1995)". Pure and Applied Chemistry. 67 (8–9): 1307–75. doi:10.1351/pac199567081307. S2CID 95004254.
  5. ^ "Production: Growth is the Norm". Chemical and Engineering News. 84 (28): 59–236. 10 July 2006. doi:10.1021/cen-v084n034.p059.
  6. ^ Sloane, N. J. A. (ed.). "Sequence A000631 (Number of ethylene derivatives with n carbon atoms)". The On-Line Encyclopedia of Integer Sequences. OEIS Foundation.
  7. ^ Smith, Michael B.; March, Jerry (2007), Advanced Organic Chemistry: Reactions, Mechanisms, and Structure (6th ed.), New York: Wiley-Interscience, p. 23, ISBN 978-0-471-72091-1
  8. ^ a b Bansal, Raj K. (1998). "Bredt's Rule". Organic Reaction Mechanisms (3rd ed.). McGraw-Hill Education. pp. 14–16. ISBN 978-0-07-462083-0.
  9. ^ Fawcett, Frank S. (1950). "Bredt's Rule of Double Bonds in Atomic-Bridged-Ring Structures". Chem. Rev. 47 (2): 219–274. doi:10.1021/cr60147a003. PMID 24538877.
  10. ^ "Bredt's Rule". Comprehensive Organic Name Reactions and Reagents. Vol. 116. 2010. pp. 525–8. doi:10.1002/9780470638859.conrr116. ISBN 978-0-470-63885-9.
  11. ^ Duan, Xufang; Block, Eric; Li, Zhen; Connelly, Timothy; Zhang, Jian; Huang, Zhimin; Su, Xubo; Pan, Yi; Wu, Lifang (28 February 2012). "Crucial role of copper in detection of metal-coordinating odorants". Proceedings of the National Academy of Sciences of the United States of America. 109 (9): 3492–7. Bibcode:2012PNAS..109.3492D. doi:10.1073/pnas.1111297109. PMC 3295281. PMID 22328155.
  12. ^ Nguyen, Trung; Clark, Jim (23 April 2019). "Physical Properties of Alkenes". Chemistry LibreTexts. Retrieved 27 May 2019.
  13. ^ Ophardt, Charles (2003). "Boiling Points and Structures of Hydrocarbons". Virtual Chembook. Retrieved 27 May 2019.
  14. ^ Hanson, John. "Overview of Chemical Shifts in H-NMR". ups.edu. Retrieved 5 May 2019.
  15. ^ a b "Nuclear Magnetic Resonance (NMR) of Alkenes". Chemistry LibreTexts. 23 April 2019. Retrieved 5 May 2019.
  16. ^ "Organic Compounds: Physical and Thermochemical Data". ucdsb.on.ca. Retrieved 5 May 2019.
  17. ^ Shriner, R.L.; Hermann, C.K.F.; Morrill, T.C.; Curtin, D.Y.; Fuson, R.C. (1997). The Systematic Identification of Organic Compounds. Wiley. ISBN 0-471-59748-1.
  18. ^ "Bromine Number". Hach company. Retrieved 5 May 2019.
  19. ^ Clark, Jim (November 2007). "The Mechanism for the Acid Catalysed Hydration of Ethene". Chemguide. Retrieved 6 May 2019.
  20. ^ Smith, Michael B.; March, Jerry (2007), Advanced Organic Chemistry: Reactions, Mechanisms, and Structure (6th ed.), New York: Wiley-Interscience, ISBN 978-0-471-72091-1
  21. ^ Baptista, Maurício S.; Cadet, Jean; Mascio, Paolo Di; Ghogare, Ashwini A.; Greer, Alexander; Hamblin, Michael R.; Lorente, Carolina; Nunez, Silvia Cristina; Ribeiro, Martha Simões; Thomas, Andrés H.; Vignoni, Mariana; Yoshimura, Tania Mateus (2017). "Type I and Type II Photosensitized Oxidation Reactions: Guidelines and Mechanistic Pathways". Photochemistry and Photobiology. 93 (4): 912–9. doi:10.1111/php.12716. PMC 5500392. PMID 28084040.
  22. ^ Oda, Masaji; Kawase, Takeshi; Kurata, Hiroyuki (1996). "1,3,5-Cyclooctatriene". Organic Syntheses. 73: 240. doi:10.15227/orgsyn.073.0240.
  23. ^ a b Hartwig, John (2010). Organotransition Metal Chemistry: From Bonding to Catalysis. New York: University Science Books. p. 1160. ISBN 978-1-938787-15-7.
  24. ^ Toreki, Rob (31 March 2015). "Alkene Complexes". Organometallic HyperTextbook. Retrieved 29 May 2019.
  25. ^ Wade, L.G. (2006). Organic Chemistry (6th ed.). Pearson Prentice Hall. pp. 309. ISBN 978-1-4058-5345-3.
  26. ^ Saunders, W. H. (1964). "Elimination Reactions in Solution". In Patai, Saul (ed.). The Chemistry of Alkenes. PATAI'S Chemistry of Functional Groups. Wiley Interscience. pp. 149–201. doi:10.1002/9780470771044. ISBN 978-0-470-77104-4.
  27. ^ Cram, D.J.; Greene, Frederick D.; Depuy, C. H. (1956). "Studies in Stereochemistry. XXV. Eclipsing Effects in the E2 Reaction1". Journal of the American Chemical Society. 78 (4): 790–6. doi:10.1021/ja01585a024.
  28. ^ Bach, R.D.; Andrzejewski, Denis; Dusold, Laurence R. (1973). "Mechanism of the Cope elimination". J. Org. Chem. 38 (9): 1742–3. doi:10.1021/jo00949a029.
  29. ^ Crowell, Thomas I. (1964). "Alkene-Forming Condensation Reactions". In Patai, Saul (ed.). The Chemistry of Alkenes. PATAI'S Chemistry of Functional Groups. Wiley Interscience. pp. 241–270. doi:10.1002/9780470771044.ch4. ISBN 978-0-470-77104-4.
  30. ^ Delaude, Lionel; Noels, Alfred F. (2005). "Metathesis". Kirk-Othmer Encyclopedia of Chemical Technology. Weinheim: Wiley-VCH. pp. metanoel.a01. doi:10.1002/0471238961.metanoel.a01. ISBN 978-0-471-23896-6.
  31. ^ Vogt, D. (2010). "Cobalt-Catalyzed Asymmetric Hydrovinylation". Angew. Chem. Int. Ed. 49 (40): 7166–8. doi:10.1002/anie.201003133. PMID 20672269.
  32. ^ Zweifel, George S.; Nantz, Michael H. (2007). Modern Organic Synthesis: An Introduction. W. H. Freeman. pp. 366. ISBN 978-0-7167-7266-8.
  33. ^ Ninkuu, Vincent; Zhang, Lin; Yan, Jianpei; et al. (June 2021). "Biochemistry of Terpenes and Recent Advances in Plant Protection". International Journal of Molecular Sciences. 22 (11): 5710. doi:10.3390/ijms22115710. PMC 8199371. PMID 34071919.
  34. ^ McMurry, John E. (2014). Organic Chemistry with Biological Applications (3rd ed.). Cengage Learning. p. 189. ISBN 978-1-285-84291-2.