Jump to content

Insertion reaction: Difference between revisions

From Wikipedia, the free encyclopedia
Content deleted Content added
→‎Organic chemistry: tweak language
(42 intermediate revisions by 3 users not shown)
Line 1: Line 1:
An '''insertion reaction''' is a [[chemical reaction]] where one chemical entity (a [[molecule]] or molecular fragment) interposes itself into an existing [[Chemical bond|bond]] of typically a second chemical entity ''e.g.'':
'''Insertion reactions''' are a type of [[chemical reaction]] in which a small molecule “inserts” itself into a metal-[[ligand]] bond. These reactions are typically [[organometallic chemistry|organometallic]] in nature and involve a bond between a [[transition metal]] and a [[carbon]] or [[hydrogen]].<ref name=DMA /> The term only refers to the result of the reaction and does not suggest a mechanism. It is usually reserved for the case where the [[coordination number]] and [[oxidation state]] of the metal remain unchanged.<ref name=Brown /> When these reactions are reversible, the removal of the small molecule from the metal-ligand bond is called extrusion or elimination.

[[File:Insertions_1'1&1'2.gif|thumb|center|300px|Examples of type 1,1 (a) and 1,2 (b) resulting geometries for insertion reactions]]
:<font color = "red">A</font> + <font color = "blue">B&ndash;C</font> → <font color = "blue">B&ndash;</font><font color = "red">A</font><font color = "blue">&ndash;C</font>

Insertion reactions are [[ ]] a metal-[[ligand]] bond reactions are typically organometallic in nature and involve a bond between a [[transition metal]] and a [[carbon]] or [[hydrogen]].<ref name=DMA /> It is usually reserved for the case where the [[coordination number]] and [[oxidation state]] of the metal remain unchanged.<ref name=Brown /> When these reactions are reversible, the removal of the small molecule from the metal-ligand bond is called extrusion or elimination.
[[File:'1&1'2.gif|thumb|center|300px|Examples of type 1,1 (a) and 1,2 (b) resulting geometries for insertion reactions]]

There are two common insertion geometries— 1,1 and 1,2 (pictured above). Additionally, the inserting molecule can act either as a [[nucleophile]] or as an [[electrophile]] to the metal complex.<ref name=Brown /> These behaviors will be discussed in more detail for [[Carbon monoxide|CO]], nucleophilic behavior, and [[Sulfur dioxide|SO<sub>2</sub>]], electrophilic behavior.

==Organic chemistry==
[[Homologation reaction]]s like the [[Kowalski ester homologation]]<ref>{{OrgSynth|author1 = Reddy, R. E.|author2 = Kowalski, C. J.|collvol = 9|collvolpages = 426|collvolyear = 1998|volume = 71|pages = 146|year = 1993|prep = cv9p0426|title = Ethyl 1-Naphthylacetate: Ester Homologation Via Ynolate Anions}}</ref> provide simple examples of insertion process in organic synthesis. In the [[Arndt-Eistert reaction]],<ref>{{cite journal|title = Ein Verfahren zur Überführung von Carbonsäuren in ihre höheren Homologen bzw. deren Derivate|last1 = Arndt|first1 = F.|authorlink1 = Fritz Arndt|last2 = Eistert|first2 = B.|authorlink2 = Bernd Eistert|journal = [[Ber. Dtsch. Chem. Ges.]]|volume = 1|issue = 68|pages = 200&ndash;208|year = 1935|language = German|doi = 10.1002/cber.19350680142}}</ref><ref>{{cite journal|last1 = Ye|first1 = T.|last2 = McKervey|first2 = M. A.|title = Organic Synthesis with &alpha;-Diazo Carbonyl Compounds|journal = [[Chem. Rev.]]|year = 1994|volume = 94|issue = 4|pages = 1091&ndash;1160|doi = 10.1021/cr00028a010}}</ref> a [[methylene]] unit is inserted into the [[carboxyl]]-carbon bond of [[carboxylic acid]] to form the next acid in the [[homologous series]]. ''[[Organic Syntheses]]'' provides the example of [[Di-tert-butyl dicarbonate|''t''-BOC protected]] (''S'')-[[phenylalanine]] (2-amino-3-phenylpropanoic acid) being reacted sequentially with [[triethylamine]], [[ethyl chloroformate]], and [[diazomethane]] to produce the [[diazo|&alpha;-diazoketone]], which is then reacted with silver trifluoroacetate / triethylamine in aqueous solution to generate the ''t''-BOC protected form of (''S'')-3-amino-4-phenylbutanoic acid.<ref name = "phenylalanine">{{OrgSynth|last1 = Linder|first1 = M. R.|last2 = Steurer|first2 = S.|last3 = Podlech|first3 = J.|title = (''S'')-3-(''tert''-Butyloxycarbonylamino)-4-phenylbutanoic acid|collvol = 10|collvolpages = 194|collvolyear = 2004|volume = 79|page = 154|year = 2002|prep = v79p0154}}</ref>

:[[Image:Homologation of N-boc-phenylalanine.png|700px|Homologation of N-boc-phenylalanine]]

Mechanistically,<ref>{{cite journal|title = The Mechanism of the Arndt-Eistert Reaction|first1 = C.|last1 = Huggett|first2 = R. T.|last2 = Arnold|first3 = T. I.|last3 = Taylor|journal = [[J. Amer. Chem. Soc.]]|year = 1942|volume = 64|issue = 12|page = 3043|doi = 10.1021/ja01264a505}}</ref> the α-diazoketone undergoes a [[Wolff rearrangement]]<ref>{{cite journal|year = 1975|journal = [[Angew. Chem. Int. Ed.]]|author1 = Meier, H.|author2 = Zeller, K.-P.|volume = 14|issue = 1|pages = 32&ndash;43|doi = 10.1002/anie.197500321|title = The Wolff Rearrangement of α-Diazo Carbonyl Compounds}}</ref><ref>{{cite journal|year = 2002|journal = [[Eur. J. Org. Chem.]]|author = Kirmse, W.|pages = 2193&ndash;2256|doi = 10.1002/1099-0690(200207)2002:14<2193::AID-EJOC2193>3.0.CO;2-D|title = 100 Years of the Wolff Rearrangement|volume = 2002|issue = 14}}</ref> to form a [[ketene]] in a [[1,2-rearrangement]]. Consequently, the methylene group α- to the carboxyl group in the product is the methylene group from the diazomethane reagant. The 1,2-rearrangement has been shown to conserve the stereochemisty of the chiral centre as the product formed from ''t''-BOC protected (''S'')-phenylalanine retains the (''S'') stereochemistry with a reported [[enantiomeric excess]] of at least 99%.<ref name = "phenylalanine" />

A related transformation is the [[Nierenstein reaction]] in which a diazomethane methylene group is inserted into the carbon-chlorine bond of an [[acid chloride]] to generate an α-chloromethyl ketone.<ref>{{cite journal|author1 = Clibbens, D. A.|author2 = Nierenstein, M.|title = The Action of Diazomethane on some Aromatic Acyl Chlorides|journal = [[J. Chem. Soc., Trans.]]|year = 1915|volume = 107|pages = 1491&ndash;1494|doi = 10.1039/CT9150701491}}</ref><ref>{{cite journal|last1 = Bachmann|first1 = W. E.|authorlink1 = Werner Emmanuel Bachmann|last2 = Struve|first2 = W. S.|title = The Arndt-Eistert Reaction|journal = [[Org. React.]]|year = 1942|volume = 1|pages = 38}}</ref> An example, published in 1924, illustrates the reaction in a substituted [[benzoyl chloride]] system:<ref>{{cite journal|title = The Action of Diazomethane on some Aromatic Acyl Chlorides II. Synthesis of Fisetol|author1 = Nierenstein, M.|author2 = Wang, D. G.|author3 = Warr, J. C.|journal = [[J. Amer. Chem. Soc.]]|volume = 46|issue = 11|pages = 2551&ndash;2555|year = 1924|doi = 10.1021/ja01676a028}}</ref>

:[[Image:Nierenstein1924.png|400px|Nierenstein 1924]]

Perhaps surprisingly, [[phenacyl bromide|&alpha;-bromoacetophenone]] is the minor product when this reaction is carried out with [[benzoyl|benzoyl bromide]], a [[dimer (chemistry)|dimeric]] [[dioxane]] being the major product.<ref>{{cite journal|title = The Action of Diazomethane on some Aromatic Acyl Chlorides III. The Mechanism of the Reaction|author1 = Lewis, H. H.|author2 = Nierenstein, M.|author3 = Rich, E. M.|journal = [[J. Amer. Chem. Soc.]]|volume = 47|issue = 6|pages = 1728&ndash;1732|year = 1925|doi = 10.1021/ja01683a036}}</ref> [[Azide#Organic azides|Organic azides]] also provide an example of an insertion reaction in organic synthesis and, like the above examples, the transformations proceed with loss of [[nitrogen|nitrogen gas]]. When [[tosyl azide]] reacts with [[norbornadiene]], a [[ring expansion and ring contraction|ring expansion]] reaction takes place in which a nitrogen atom is inserted into a carbon-carbon bond α- to the bridge head:<ref>{{cite journal|title = A Facile Synthesis of a Polyhydroxylated 2-Azabicyclo[3.2.1]octane|first1 = D. D.|last1 = Reed|first2 = S. C.|last2 = Bergmeier|journal = [[J. Org. Chem.]]|doi = 10.1021/jo0619231|year = 2007|volume = 72|pages = 1024&ndash;1026|pmid = 17253828|issue = 3}}</ref>

:[[Image:Azidenorbornadieneinsertion.png|400px|Norbornadiene reaction with tosyl azide]]

The [[Beckmann rearrangement]]<ref>{{cite journal|first = E.|last = Beckmann|authorlink = Ernst Otto Beckmann|title = Zur Kenntniss der Isonitrosoverbindungen|journal = [[Ber. Dtsch. Chem. Ges.]]|year = 1886|volume = 19|pages = 988&ndash;993|doi = 10.1002/cber.188601901222|language = German}}</ref><ref>{{cite journal|first = R. E.|last = Gawley|title = The Beckmann Reactions: Rearrangement, Elimination-Additions, Fragmentations, and Rearrangement-Cyclizations.|journal = [[Org. React.]]|year = 1988|volume = 35|pages = 14&ndash;24|doi = 10.1002/0471264180.or035.01}}</ref> is another example of a ring expanding reaction in which a heteroatom is inserted into a carbon-carbon bond. The most important application of this reaction is the conversion of [[cyclohexanone]] to its oxime, which is then rearranged under acidic conditions to provide ε-[[caprolactam]],<ref>{{OrgSynth|first1 = J. C.|last1 = Eck|first2 = C. S.|last2 = Marvel|title = &epsilon;-Benzoylaminocaproic acid|year = 1939|volume = 19|page = 20|collvol = 2|collvolpages = 76|collyear = 1943|prep = cv2p0076}}</ref> the feedstock for the manufacture of [[Nylon 6]]. Annual production of caprolactam exceeds 2 billion kilograms.<ref>{{cite book|first1 = J.|last1 = Ritz|first2 = H.|last2 = Fuchs|first3 = H.|last3 = Kieczka|first4 = W. C.|last4 = Moran|chapter = Caprolactam|title = Ullmann's Encyclopedia of Industrial Chemistry|publisher = Wiley-VCH|location = Weinheim|year = 2000|doi = 10.1002/14356007.a05_031}}</ref>

:[[Image:BeckmannRearrangement3.svg|400px|The Beckmann Rearrangement]]

[[Carbene]]s undergo both [[intermolecular]] and [[intramolecular]] insertion reactions. [[Cyclopentene]] moieties can be generated from sufficiently long-chain [[ketone]]s by reaction with [[trimethylsilyldiazomethane]], (CH<sub>3</sub>)<sub>3</sub>Si&ndash;CHN<sub>2</sub>:

:[[Image:alkylidene carbene.png|450px|Alkylidene carbene]]

Here, the carbene intermediate inserts into a carbon-hydrogen bond to form the carbon-carbon bond needed to close the cyclopentene ring. [[Carbene C-H insertion|Carbene insertions into carbon-hydrogen bonds]] can also occur intermolecularly:

:[[Image:carbene inter.png|500px|Carbene intermolecular reaction]]

[[Carbenoid]]s are [[reactive intermediate]]s that behave similarly to carbenes.<ref>{{cite book|title = Organic Chemistry|last = McMurry|first = J.|authorlink = John E. McMurry|publisher = [[Brooks/Cole]]|edition = 2nd|year = 1988|isbn = 0-534-07968-7}}</ref> One example is the chloroalkyllithium carbenoid reagent prepared ''in situ'' from a [[sulfoxide]] and [[organolithium|''t''-BuLi]] which inserts into the carbon-boron bond of a [[pinacol boronic ester]]:<ref>{{cite journal|title = Iterative Stereospecific Reagent-Controlled Homologation of Pinacol Boronates by Enantioenriched-Chloroalkyllithium Reagents|first1 = P. R.|last1 = Blakemore|authorlink1 = Paul R. Blakemore|first2 = M. S.|last2 = Burge|journal = [[J. Amer. Chem. Soc.]]|year = 2007|volume = 129|issue = 11|pages = 3068&ndash;3069|doi = 10.1021/ja068808s}}</ref>

:[[Image:CarbenoidApplication.png|400px|Insertion of carbenoid into carbon-boron bond]]


There are two common insertion geometries— 1,1 and 1,2 (pictured above). Additionally, the inserting molecule can act either as a [[nucleophile]] or as an [[electrophile]] to the metal complex.<ref name=Brown /> These behaviors will be discussed in more detail for [[Carbon monoxide|CO]], nucleophilic behavior, and [[Sulfur dioxide|SO<sub>2</sub>]], electrophilic behavior.
==CO Insertion==
==CO Insertion==
The insertion of carbon monoxide across a metal-carbon site to form an [[acetyl]] group is the oldest-known and most-studied metal-ligand insertion reaction. It proceeds by a 1,1 reaction coordinate, attaching the [[carbonyl]] carbon to both the metal and the ligand. The first CO insertion was discovered in 1957 by the reaction of CO with MnCO<sub>5</sub>CH<sub>3</sub>, forming Mn(CO)<sub>5</sub>COCH<sub>3</sub>.
The insertion of carbon monoxide across a metal-carbon site to form an [[acetyl]] group is the oldest-known and most-studied metal-ligand insertion reaction. It proceeds by a 1,1 reaction coordinate, attaching the [[carbonyl]] carbon to both the metal and the ligand. The first CO insertion was discovered in 1957 by the reaction of CO with <sub>5</sub>CH<sub>3</sub>, forming Mn(CO)<sub>5</sub>COCH<sub>3</sub>.


===Mechanism===
===Mechanism===
The mechanism for the apparent CO insertion into a metal-[[alkyl]] bond is actually a [[migratory insertion]], with a migration of the alkyl group to another bound CO, followed by addition of a free CO (see figure below). This can be demonstrated by [[Carbon-13|<sup>13</sup>C]]-labeling the incoming CO ligand, which results in 100% of the labeled CO residing [[Cis-trans isomerism|cis-]] to the acetyl group. [[File:Octahedral.png|thumb|center|606px|CO Insertion reaction pathway for an octahedral complex]]
The mechanism for the apparent CO insertion into a metal-[[alkyl]] bond is actually a [[migratory insertion]], with a migration of the alkyl group to another bound CO, followed by addition of a free CO (see figure below).<> in |
Yadav, M. S.|url = http://books.google.com.au/books?id=qlwlWsaoLwcC&pg=PA244&dq=Mn(CO)5CH3+insertion+reaction&hl=en&ei=F1ETTb2bBsa3cK_SzMAK&sa=X&oi=book_result&ct=result&resnum=1&ved=0CCQQ6AEwADgK#v=onepage&q=Mn(CO)5CH3%20insertion%20reaction&f=false|publisher = Anmol Publications|year = 2005|page = 244|isbn = 9788126118984}}</ref> This can be demonstrated by [[Carbon-13|<sup>13</sup>C]]-labeling the incoming CO ligand, which results in 100% of the labeled CO residing [[Cis-trans isomerism|cis-]] to the acetyl group. [[File:Octahedral.png|thumb|center|606px|CO Insertion reaction pathway for an octahedral complex]]


The CO insertion mechanism is not always a migration. The reaction of CO with (Cp)MeLFeCO, where L is a nucleophilic group such as [[Triphenylphosphine|PPh<sub>3</sub>]], yields a mix of both alkyl migration products and products formed by true insertion of bound carbonyls into the [[methyl]] group, which is controllable by the choice of solvent.<ref name="ACR">{{cite journal|last=Anderson|first=Gordon|title=Carbonyl-Insertion Reactions of Square Planar Complexes|journal=Acc. Chem. Res|year=1984|volume=17|issue=17|pages=67-74}}</ref>
The CO insertion mechanism is not always a migration. The reaction of CO with (Cp)MeLFeCO, where L is a nucleophilic group such as [[Triphenylphosphine|PPh<sub>3</sub>]], yields a mix of both alkyl migration products and products formed by true insertion of bound carbonyls into the [[methyl]] group, which is controllable by the choice of solvent.<ref name="ACR">{{cite journal|=Anderson|=|title=Carbonyl-Insertion Reactions of Square Planar Complexes|journal=Acc. Chem. Res|year=1984|volume=17|issue=|pages=6774}}</ref>


[[Square planar molecular geometry|Square planar complexes]] can also undergo CO insertions. Insertion reactions in square planar complexes are of particular interest because their structure allows additional reaction mechanisms to occur. While just like [[Octahedral molecular geometry|octahedral complexes]], square planar complexes can undergo in-plane migration, their lack of out-of-plane steric hindrance renders them much more open to nucleophilic attack of the metal by the CO. Since square planar groups usually form 16 electron species, the 5-coordinate intermediate that forms is stabilized by the [[18-Electron rule]], and undergoes migratory insertion readily. <ref name=ACR /> In most cases the in-plane migration pathway is preferred, but, unlike the nucleophilic pathway, it is inhibited by an excess of CO. <ref name="CCR">{{cite journal|last=Cavell|first=Kingsley|title=Recent Fundamental studies on migratory insertion into metal-carbon bonds|journal=Coordination Chemistry Reviews|year=1996|month=11|volume=155|pages=209-243}}</ref>
[[Square planar molecular geometry|Square planar complexes]] can also undergo CO insertions. Insertion reactions in square planar complexes are of particular interest because their structure allows additional reaction mechanisms to occur. While just like [[Octahedral molecular geometry|octahedral complexes]], square planar complexes can undergo in-plane migration, their lack of out-of-plane steric hindrance renders them much more open to nucleophilic attack of the metal by the CO. Since square planar groups usually form 16 electron species, the 5-coordinate intermediate that forms is stabilized by the [[18-Electron rule]], and undergoes migratory insertion readily.<ref name=ACR /> In most cases the in-plane migration pathway is preferred, but, unlike the nucleophilic pathway, it is inhibited by an excess of CO.<ref name="CCR">{{cite journal|last=Cavell|first=Kingsley|title=Recent Fundamental on into - |journal= |year=1996|=11|volume=155|pages=209243}}</ref>


[[File:Square_Planar.png|thumb|center|800px|Nucleophilic insertion and rearrangement of a square planar complex]]
[[File:.png|thumb|center|800px|Nucleophilic insertion and rearrangement of a square planar complex]]


===Effects on reaction rates===
===Effects on reaction rates===


*[[Steric effects|Steric]] strain - Increasing the steric strain of the [[chelation|chelate]] backbone in square planar complexes pushes the carbonyl and methyl groups closer together, increasing the reactivity of insertion reactions. <ref name=CCR />
*[[Steric effects|Steric]] strain Increasing the steric strain of the [[chelation|chelate]] backbone in square planar complexes pushes the carbonyl and methyl groups closer together, increasing the reactivity of insertion reactions.<ref name=CCR />


*Oxidation state - Oxidation of the metal tends to increase insertion reaction rates. As the main rate-limiting step in the reaction is the migration of CH<sub>3</sub>- to CO, oxidizing the metal gives a greater partial positive charge on the CO carbon, increasing the rate of reaction.<ref name=Brown />
*Oxidation state Oxidation of the metal tends to increase insertion reaction rates. main rate-limiting step in the is the migration of , oxidizing the metal a greater partial positive charge on the carbon, increasing the rate of reaction.<ref name=Brown />


*[[Lewis acids and bases| Lewis acids]] - Lewis acids also increase the reaction rates, for reasons similar to metal oxidation increasing the positive charge on the carbon. Lewis acids bind to the CO oxygen and remove charge, increasing the electrophilicity of the carbon. This can increase the reaction rate by a factor of up to 10<sup>8</sup>, and the complex formed is stable enough that the reaction proceeds even without additional CO to bind to the metal.<ref name=Brown />
*[[Lewis acids and bases|Lewis acids]] Lewis acids also increase the reaction rates, for reasons similar to metal oxidation increasing the positive charge on the carbon. Lewis acids bind to the CO oxygen and remove charge, increasing the electrophilicity of the carbon. This can increase the reaction rate by a factor of up to 10<sup>8</sup>, and the complex formed is stable enough that the reaction proceeds even without additional CO to bind to the metal.<ref name=Brown />


*[[Electronegativity]] of the leaving group - Increasing the electronegativity of the leaving alkyl group stabilizes the metal-carbon bond interaction and thus increases the [[activation energy]] required for migration, decreasing the reaction rate.<ref name="Shusterman">{{cite journal|last=Shusterman|first=Alan|coauthors=Idan Tamir, Addy Pross|title=The mechanism of organometallic micration reactions. A configuration mixing approach|journal=Journal of Organometallic Chemistry|year=1988|volume=340|pages=203-222| doi=10.1016/0022-328X(88)80076-7 }}</ref>
*[[Electronegativity]] of the leaving group - Increasing the electronegativity of the leaving alkyl group stabilizes the metal-carbon bond interaction and thus increases the [[activation energy]] required for migration, decreasing the reaction rate.<ref name="Shusterman">{{cite journal|=Shusterman|=|= Tamir Pross|title=The of . A |journal= |year=1988|volume=340|pages=203222|doi=10.1016/0022-328X(88)80076-7 }}</ref>


*[[Trans effect|Trans-effect]] Ligands in an octahedral or square planar complex are known to influence the reactivity of the group they are trans- to. This ligand influence is often referred to as the trans-influence, and it varies in intensity between ligands. A partial list of trans-influencing ligands is as follows, from highest trans-effect to lowest:<ref name=ACR /> [[aryl]], alkyl > NR<sub>3</sub> > PR<sub>3</sub> > AsR<sub>3</sub> > CO > [[Chlorine|Cl]]. Ligands with a greater trans-influence impart greater electrophilicity to the active site. Increasing the electrophilicity of the CO group has been shown experimentally to greatly increase the reaction rate, while decreasing the electrophilicity of the methyl group slightly increases the reaction rate. This can be demonstrated by reacting a square planar PNMCOCH<sub>3</sub> complex with CO, where PN is a [[bidentate]] P/N ligand bound to the metal. This reaction proceeds in much greater yield when the methyl group is trans-P and the CO trans-N, owing to nitrogen's higher trans-influence.<ref name=CCR />
*[[Trans effect|Trans-effect]] Ligands in an octahedral or square planar complex are known to influence the reactivity of the group they are trans-. This ligand influence is often referred to as the trans-influence, and it varies in intensity between ligands. A partial list of trans-influencing ligands is as follows, from highest trans-effect to lowest:<ref name=ACR /> [[aryl]], alkyl > NR<sub>3</sub> > PR<sub>3</sub> > AsR<sub>3</sub> > CO > [[Chlorine|Cl]]. Ligands with a greater trans-influence impart greater electrophilicity to the active site. Increasing the electrophilicity of the CO group has been shown experimentally to greatly increase the reaction rate, while decreasing the electrophilicity of the methyl group slightly increases the reaction rate. This can be demonstrated by reacting a square planar <sub>3</sub> complex with CO, where PN is a [[bidentate]] bound . This reaction proceeds in much greater yield when the methyl group is trans-P and the CO trans-N, owing to higher trans-influence.<ref name=CCR />


===Reverse reaction===
===Reverse reaction===
The reverse reaction of decarbonylation of [[aldehydes]] is much more difficult to perform than the associated carbonylation, as it requires both a cis- empty site and enough energy to break a carbon-carbon bond. Additionally, such reactions are usually not [[catalytic]], as the extruded CO binds too tightly to the metal to be freed even under the high heats required for this reaction.<ref>{{cite journal|last=Beck|doi=10.1021/om9905106}}</ref> Extrusion of CO from an organic aldehyde is most famously demonstrated using [[Wilkinson's catalyst|Wilkinson's catalyst]], which undergoes [[oxidative addition]] of the aldehyde C-H instead of H-H, and then performs the extrusion of CO to form an octahedral product.<ref name="OHNO">{{cite journal|last=Ohno|first=K|coauthors=J. Tsuji|journal=I. Am. Chem. Soc|year=1968|volume=90|pages=99|doi=10.1021/ja01003a018}}</ref>
of [[aldehydes]] is much more difficult to perform than the associated carbonylation as it requires both a cis- empty site and enough energy to break a carbon-carbon bond. Additionally, such reactions are usually not [[catalytic]], as the extruded CO binds too tightly to the metal to be even under the high heats required for this reaction.<ref>{{cite journal|=Beck|doi=10.1021/om9905106}}</ref> Extrusion of CO from an organic aldehyde is most famously demonstrated using [[Wilkinson's catalyst]], which undergoes [[oxidative addition]] of the aldehyde CH instead of HH, and then extrusion to form an octahedral product.<ref name="OHNO">{{cite journal|=Ohno|=K|=J. Tsuji|journal=. . Chem. Soc|year=1968|volume=90|pages=99|doi=10.1021/ja01003a018}}</ref>


===Industrial applications===
===Industrial applications===
The most widely known application of migratory insertion of carbonyl groups is the [[Monsanto process|Monsanto acetic acid process]], which uses an rhodium-iodine catalyst system to transform methanol into acetic acid under relatively mild, catalytic conditions. It has been more recently superceded by the [[Cativa process]] which uses the ''cis''-dicarbonyldiiodoiridate(I) anion [Ir(CO)<sub>2</sub>I<sub>2</sub>]<sup>&minus;</sup> ('''1''') as catalyst.<ref name = "Cativa">{{cite journal|title = The Cativa<sup>TM</sup> Process for the Manufacture of Acetic Acid|author = Jones, J. H.|journal = [[Platinum Metals Rev.]]|year = 2000|volume = 44|issue = 3|pages = 94&ndash;105|url = http://www.platinummetalsreview.com/pdf/pmr-v44-i3-094-105.pdf}}</ref><ref>{{cite journal|title = High Productivity Methanol Carbonylation Catalysis using Iridium - The Cativa<sup>TM</sup> Process for the Manufacture of Acetic Acid|author1 = Sunley, G. J.|author2 = Watson, D. J.|journal = Catalysis Today|year = 2000|volume = 58|issue = 4|pages = 293&ndash;307|doi = 10.1016/S0920-5861(00)00263-7}}</ref> By 2002, worldwide annual production of acetic acid stood at 6 million tons, of which approximately 60% is produced by the Cativa process.<ref name = "Cativa" />
The most widely-known and widely-used application of migratory insertion of carbonyl groups is the [[Monsanto process|Monsanto acetic acid process]]. This reaction, through an iodine intermediate, transforms methanol into acetic acid under relatively mild, catalytic conditions. It is used industrially to create over a million tons of acetic acid every year.

:[[File:Cativa-process-catalytic-cycle.png|center|400px|The catalytic cycle of the Cativa process]]

The [[catalytic cycle]] for the Cativa process, shown above, begins with the reaction of [[methyl iodide]] with the square planar active catalyst species ('''1''') to form the octahedral iridium(III) species ('''2'''), the [[Octahedral molecular geometry#Facial and meridional isomers|''fac''-isomer]] of [Ir(CO)<sub>2</sub>(CH<sub>3</sub>)I<sub>3</sub>]<sup>&minus;</sup>. This oxidative addition reaction involves the formal insertion of the iridium(I) centre into the carbon-iodine bond of methyl iodide. After ligand exchange, the migratory insertion of carbon monoxide into the iridium-carbon bond, step ('''3''') to ('''4'''), results in the formation of a [[square pyramidal molecular geometry|square pyramidal]] species with a bound [[acetyl]] ligand. The active catalyst species ('''1''') is regenerated by the [[reductive elimination]] of [[acetyl iodide]] from ('''4'''), a de-insertion reaction.<ref name = "Cativa" />


==SO<sub>2</sub> Insertion==
==SO<sub>2</sub> Insertion==
SO<sub>2</sub> is an electrophilic species. When SO<sub>2</sub> encounters a transition metal with alkyl ligands, it acts like a Lewis acid towards the alkyl ligand.<ref name="DMA">{{cite book|last=Douglas, McDaniel, and Alexander|title=Concepts and Models of Inorganic Chemistry 3rd Ed|year=1994|publisher=John Wiley & Sons, Inc.|isbn=9780471629788}}</ref>
SO<sub>2</sub> is an electrophilic species. When SO<sub>2</sub> encounters a transition metal with alkyl ligands, it acts like a Lewis acid towards the alkyl ligand.<ref name="DMA">{{cite book|last=Douglas, McDaniel, and Alexander|title=Concepts and Models of Inorganic Chemistry 3rd Ed|year=1994|publisher=John Wiley & Sons, Inc.|isbn=9780471629788}}</ref>

===Mechanism===
===Mechanism===
The mechanism of SO<sub>2</sub> insertion is shown in the figure below.<ref name=blue /> First, the sulfur coordinates to the ligand with the electronegative oxygens pointing away from the complex. Then, two different pathways can occur to “insert” the sulfur dioxide between the metal and the alkyl ligand. The top pathway shows the creation of an ionic complex, O,O’-sulphinate where the metal has a positive charge and the ligand has a negative charge. The bottom pathway shows the creation of an O-sulphinate, which can be a stable product. Also, either of these intermediates can form an S-sulphinate, the most common SO<sub>2</sub> insertion species and a 1,1 type geometry. [[IR spectroscopy]] and [[NMR]] can determine the type of compound.<ref name="blue">{{cite book|title=Advances in Organometallic Chemistry|year=1974|publisher=Academic Press|author=A. Wojcicki|authorlink=Sulfur Dioxide Insertion Reactions|editor=Stone and West}}</ref> S-sulphinate has sulfur-oxygen stretching frequencies from 1250-1000 cm<sup>-1</sup> and 1100-1000 cm<sup>-1</sup>. The O, O'-sulphinate and O-sulphinate are difficult to distinguish as they have stretching frequencies from 1085-1050 cm<sup>-1</sup> and 1000-820 cm<sup>-1</sup> or lower. The pathway involving the O, O' sulphinate can generally be ruled out if the original metal complex fulfilled the 18 electron rule because the two metal-oxygen bonds would exceed the 18 electron rule. <ref name="Brown">{{cite book|title=The chemistry of the metal-carbon bond, vol. 2|year=1985|publisher=John Wiley & Sons|author=J.J. Alexander|authorlink=Insertions into transition metal-carbon bonds|editor=Hartley and Patai}}</ref>
The mechanism of SO<sub>2</sub> insertion is shown in the figure below.<ref name=blue /> First, the sulfur coordinates to the ligand with the electronegative oxygens pointing away from the complex. Then, two different pathways can occur to “insert” the sulfur dioxide between the metal and the alkyl ligand. The top pathway shows the creation of an ionic complex, O,O’-sulphinate where the metal has a positive charge and the ligand has a negative charge. The bottom pathway shows the creation of an O-sulphinate, which can be a stable product. Also, either of these intermediates can form an S-sulphinate, the most common SO<sub>2</sub> insertion species and a 1,1 type geometry. [[IR spectroscopy]] and [[NMR]] can determine the type of compound.<ref name="blue">{{cite book|title=Advances in Organometallic Chemistry|year=1974|publisher=Academic Press|author=A. Wojcicki|authorlink=Sulfur Dioxide Insertion Reactions|editor=Stone and West}}</ref> S-sulphinate has sulfur-oxygen stretching frequencies from cm<sup></sup> and cm<sup></sup>. The O, O'-sulphinate and O-sulphinate are difficult to distinguish as they have stretching frequencies from cm<sup></sup> and cm<sup></sup> or lower. The pathway involving the O, O' sulphinate can generally be ruled out if the original metal complex fulfilled the 18 electron rule because the two metal-oxygen bonds would exceed the 18 electron rule.<ref name="Brown">{{cite book|title=The chemistry of the metal-carbon bond, vol. 2|year=1985|publisher=John Wiley & Sons|author=J.J. Alexander|authorlink=Insertions into transition metal-carbon bonds|editor=Hartley and Patai}}</ref>
[[File:Mechanism SO2 insertion labels.jpg|thumb|center|600px|Mechanism of SO<sub>2</sub> insertion]]
[[File:Mechanism SO2 insertion labels.jpg|thumb|center|600px|Mechanism of SO<sub>2</sub> insertion]]
When SO<sub>2</sub> is inserting into a square planar complex that does not have 18 electrons, it inserts slightly differently. Since the metal is [[saturation (chemistry)|coordinatively unsaturated]], the SO<sub>2</sub> can pre-coordinate with the metal. Then, it inserts itself between the metal and the alkyl group as shown in the example below.<ref name="Puddephatt">{{cite journal|last=Puddephatt|first=R.A.|coauthors=Stalteri, M.A.|title=Competition between Insertion of Sulfur Dioxide into the Methyl- or Phenyl- Transition Metal Bond|journal=Journal of Organometallic Chemistry|year=1980|volume=193|pages=C27|DOI=10.1016/S0022-328X(00)86091-X}}</ref>
When SO<sub>2</sub> is inserting into a square planar complex that does not have 18 electrons, it inserts slightly differently. Since the metal is [[saturation (chemistry)|coordinatively unsaturated]], the SO<sub>2</sub> can pre-coordinate with the metal. Then, it inserts itself between the metal and the alkyl group as shown in the example below.<ref name="Puddephatt">{{cite journal|last=Puddephatt|first=R.A.|coauthors=Stalteri, M.A.|title=Competition between Insertion of Sulfur Dioxide into the Methyl- or Phenyl- Transition Metal Bond|journal=Journal of Organometallic Chemistry|year=1980|volume=193|pages=C27|DOI=10.1016/S0022-328X(00)86091-X}}</ref>
Line 42: Line 84:


===Effects on reaction rates===
===Effects on reaction rates===
The ability of SO<sub>2</sub> to insert depends upon the electronic interactions between the metal and the ligand. If there is [[pi bond|pi bonding]] between the ligand and metal, the SO<sub>2</sub> will not insert. When steric effects are negligible, the ability of SO<sub>2</sub> to insert increases with an increase in the electron donating capacity of the ligand and decreases with the electron withdrawing capacity of the ligand. These trends highlight the electrophilic nature of SO<sub>2</sub> insertion.<ref name=blue />
The ability of SO<sub>2</sub> to insert depends upon the electronic interactions between the metal and the ligand. If there is [[pi bond]] between the ligand and metal, the SO<sub>2</sub> will not insert. When steric effects are negligible, the ability of SO<sub>2</sub> to insert increases with an increase in the electron donating capacity of the ligand and decreases with the electron withdrawing capacity of the ligand. These trends highlight the electrophilic nature of SO<sub>2</sub> insertion.<ref name=blue />


===Reverse reaction===
===Reverse reaction===
Line 49: Line 91:


===Industrial applications===
===Industrial applications===
There are few examples of SO<sub>2</sub> insertion in industry due to the low reactivity of SO<sub>2</sub> insertions and the low stability of the created products, but SO<sub>2</sub> insertion can be a step in production of [[sulfolane]]s. <ref name="Kurosawa">{{cite book|title=Fundamentals of Molecular Catalysis, Volume 3 (Current Methods in Inorganic Chemistry)|year=2003|publisher=Elsevier|isbn=0444509216|editor=Kurosawa, H.; Yamamoto, A.}}</ref><ref name="Dzhemilev">{{cite journal|last=Dzhemilev|first=K|coauthors=Kunakova, R.V.|title=Metal complex catalysis in the synthesis of organic sulfur compounds|journal=Journal of Organometallic Chemistry|year=1993|volume=455|doi=10.1016/0022-328X(93)80375-L}}</ref>
There are few examples of SO<sub>2</sub> insertion in industry due to the low reactivity of SO<sub>2</sub> insertions and the low stability of the created products, but SO<sub>2</sub> insertion can be a step in production of [[sulfolane]]s.<ref name="Kurosawa">{{cite book|title=Fundamentals of Molecular Catalysis, Volume 3 (Current Methods in Inorganic Chemistry)|year=2003|publisher=Elsevier|isbn=0444509216|editor=Kurosawa, H.; Yamamoto, A.}}</ref><ref name="Dzhemilev">{{cite journal|last=Dzhemilev|first=K|coauthors=Kunakova, R.V.|title=Metal complex catalysis in the synthesis of organic sulfur compounds|journal=Journal of Organometallic Chemistry|year=1993|volume=455|doi=10.1016/0022-328X(93)80375-L}}</ref>


==Olefin Insertion==
==Olefin Insertion==
[[Olefin]] insertions are insertion reactions involving [[alkene]]s and [[alkyne]]s and they can insert into metal-carbon as well as metal-hydrogen bonds. Many of olefin insertions' features are similar to those of carbonyl insertion in terms of the mechanism, however key differences are present. One of these is the potential to perform multiple insertions in tandem providing a mechanism for [[polymerization]]<ref name="Sinn & Kaminsky">{{cite journal|last=Sinn|first=H.|coauthors=Kaminsky, W.|journal=Advances in Organometallic Chemistry|year=1980|volume=18|pages=99}}</ref>.
[[Olefin]] insertions are insertion reactions involving [[alkene]]s and [[alkyne]]s and they can insert into metal-carbon as well as metal-hydrogen bonds. Many of olefin insertions' features are similar to those of carbonyl insertion in terms of the mechanism, however key differences are present. One of these is the potential to perform multiple insertions in tandem providing a mechanism for [[polymerization]]<ref name="Sinn & Kaminsky">{{cite journal|last=Sinn|first=H.|coauthors=Kaminsky, W.|journal=Advances in Organometallic Chemistry|year=1980|volume=18|pages=99}}</ref>

===Mechanism===
===Mechanism===
The mechanism for olefin insertions into metal-carbon or metal-hydrogen bonds converts an olefin complex into an alkyl complex, forming a bond between the carbon or hydrogen with the carbon at the β position. This insertion at the same time forms a bond between the metal and the carbon at the α position. When a metal has sufficient electrons for [[pi backbonding|pi-backbonding]], the olefin can be coordinated to the metal before insertion. Depending on the ligand density of the metal, ligand dissociation may be necessary to provide a coordination site for the olefin.
The mechanism for olefin insertions into metal-carbon or metal-hydrogen bonds converts an olefin complex into an alkyl complex, forming a bond between the carbon or hydrogen with the carbon at the β position. This insertion at the same time forms a bond between the metal and the carbon at the α position. When a metal has sufficient electrons for [[pi backbonding|pi-backbonding]], the olefin can be coordinated to the metal before insertion. Depending on the ligand density of the metal, ligand dissociation may be necessary to provide a coordination site for the olefin.
[[File:Olefin_Insertion_B-elimination)_revised.gif|thumb|center|600px|Olefin insertion (forward reaction) and beta-elimination (reverse reaction)]]
[[File:-elimination).gif|thumb|center|600px|Olefin insertion (forward reaction) and beta-elimination (reverse reaction)]]


Fundamental features of olefin insertions include the mechanism's dependance on the electron density of the metal complex it is inserting into. For instance, a coordinatively saturated 18-electron metal complex will likely require ligand dissociation to provide a vacant coordination site for the insertion to occur (this is also true for the β-elimination<ref name="Reger & Culbertson">{{cite journal|last=Reger|first=D.|coauthors=Culbertson, E.|title=Mechanism of the thermal decomposition of carbonyl(.eta.5-cyclopentadienyl)triphenylphosphineiron(II) alkyl derivatives into carbonyl(.eta.5-cyclopentadienyl)hydrido(triphenylphosphine)iron(II) and olefin|journal=Journal of the American Chemical Society|year=1976|month=May|volume=98|pages=2789|doi=10.1021/ja00426a020}}</ref>). In the case of coordinatively unsaturated metal complexes with lower electron counts (such as the 14-electron Cp*<sub>2</sub>ScCH<sub>3</sub> complex), there are already vacant coordination sites available for olefin insertion without the need for ligand dissociation.
Fundamental features of olefin insertions include the mechanism's dependance on the electron density of the metal complex it is inserting into. For instance, a coordinatively saturated 18-electron metal complex will likely require ligand dissociation to provide a vacant coordination site for the insertion to occur (this is also true for the β-elimination<ref name="Reger & Culbertson">{{cite journal|=Reger|=D.|=CulbertsonE.|title=Mechanism of the of (eta5-cyclopentadienyl)triphenylphosphineiron(II) into (eta5-cyclopentadienyl)hydrido(triphenylphosphine)iron(II) and |journal= |year=1976|volume=98|pages=2789|doi=10.1021/ja00426a020}}</ref>). In the case of coordinatively unsaturated metal complexes with lower electron counts (such as the 14-electron Cp*<sub>2</sub>ScCH<sub>3</sub> complex), there are already vacant coordination sites available for olefin insertion without the need for ligand dissociation.


===Effects on reaction rates===
===Effects on reaction rates===
Factors affecting the rate of olefin insertions include the formation of the cyclic, planar, four-center transition state involving incipient formation of a bond between the metal and an olefin carbon.
Factors affecting the rate of olefin insertions include the formation of the cyclic, planar, four-center transition state involving incipient formation of a bond between the metal and an olefin carbon.
[[File:Olefin_Insertion_Transition_State.gif|thumb|center|600px|Transition state of olefin insertion reaction]]
[[File:.gif|thumb|center|600px|Transition state of olefin insertion reaction]]

From this transition state, it can be seen that a partial positive charge forms on the β-position carbon with a partial negative charge formed on the carbon or hydrogen initially bonded to the metal. This explains the subsequently observed formation of the bond between the negatively-charged carbon/hydrogen and the positively-charged β-carbon as well as the simultaneously formation of the metal-α-carbon bond.


From this transition state, it can be seen that a partial positive charge forms on the β-position carbon with a partial negative charge formed on the carbon or hydrogen initially bonded to the metal. This explains the subsequently observed formation of the bond between the negativelycharged carbon/hydrogen and the positivelycharged β-carbon as well as the simultaneously formation of the metal-α-carbon bond.


This transition state also allows for consideration of the two most contributing factors to the rate of olefin insertion reactions. These include the orbital overlap of the alkyl group initially attached to the metal and the strength of the metal-alkyl bond. With greater orbital overlap between the partially positive β-carbon and the partially negative hydrogen/alkyl group carbon, the formation of a bond between them is facilitated and the insertion reaction rate is increased. With increasing strength of the metal-alkyl bond, the breaking of the bond between the metal and the hydrogen/alkyl carbon bond to form the two new bonds with the α-carbon and β-carbon (respectively) is slower, thus decreasing the rate of the insertion reaction<ref name="Burger et al.">{{cite journal|last=Burger|first=Barbara|coauthors=Thompson, Mark; Cotter, W; Bercaw, John|title=Ethylene insertion and .beta.-hydrogen elimination for permethylscandocene alkyl complexes. A study of the chain propagation and termination steps in Ziegler-Natta polymerization of ethylene|journal=Journal of the American Chemical Society|year=1990|month=February|volume=112|pages=1566|doi=10.1021/ja00160a041}}</ref>.
This transition state also allows for consideration of the two most contributing factors to the rate of olefin insertion reactions. These include the orbital overlap of the alkyl group initially attached to the metal and the strength of the metal-alkyl bond. With greater orbital overlap between the partially positive β-carbon and the partially negative hydrogen/alkyl group carbon, the formation of a bond between them is facilitated and the insertion reaction rate is increased. With increasing strength of the metal-alkyl bond, the breaking of the bond between the metal and the hydrogen/alkyl carbon bond to form the two new bonds with the α-carbon and β-carbon (respectively) is slower, thus decreasing the rate of the insertion reaction<ref name="Burger et al.">{{cite journal|=Burger|=|=Thompson CotterW Bercaw John|title=Ethylene and beta-hydrogen for . A of the and in Ziegler-Natta of |journal= |year=1990|=|volume=112|pages=1566|doi=10.1021/ja00160a041}}</ref>


===Reverse reaction===
===Reverse reaction===
The reverse mechanism for olefin insertion into a metal-hydrogen bond is [[Beta elimination|β-elimination]], which can only occur if a vacant orbital and a vacant coordination position are available in addition to the presence of a hydrogen at the β position. If these conditions exist, the decomposition of the metal alkyl complex can take place resulting in a hydride/olefin complex that can subsequently dissociate the olefin.
The reverse mechanism for olefin insertion into a metal-hydrogen bond is [[Beta elimination|β-elimination]], which can only occur if a vacant orbital and a vacant coordination position are available in addition to the presence of a hydrogen at the β position. If these conditions exist, the decomposition of the metal alkyl complex can take place resulting in a hydride/olefin complex that can subsequently dissociate the olefin.


===Industrial applications===
===Industrial applications===
As mentioned previously, a fundamental feature of olefin insertion is the opportunity for multiple insertions, providing a mechanism for polymerization. An industrial application of this is ethylene polymerization, which is catalyzed heterogeneously by titanium(III) salts and aluminum alkyls (also known as [[Ziegler-Natta catalyst]]s). In these reactions, ethylene is coordinated to a vacant site on the titanium metal followed by its insertion. This can be repeated multiple times, potentially leading to high molecular weight polymers.
As mentioned previously, a fundamental feature of olefin insertion is the opportunity for multiple insertions, providing a mechanism for polymerization. An industrial application of this is ethylene polymerization, which is catalyzed heterogeneously by titanium(III) salts and aluminum alkyls (also known as [[Ziegler-Natta catalyst]]s). In these reactions, ethylene is coordinated to a vacant site on the titanium metal followed by its insertion. This can be repeated multiple times, potentially leading to high molecular weight polymers.
[[File:OlefinInsertion_Polymerization.gif|thumb|center|600px|Synthetic pathway of ethylene polymerization via olefin insertion reaction]]
[[File:.gif|thumb|center|600px|Synthetic pathway of ethylene polymerization via olefin insertion reaction]]


==Other Insertion reactions==
==Other Insertion reactions==
Line 81: Line 123:
{{Reflist}}
{{Reflist}}


{{DEFAULTSORT:Insertion Reaction}}
[[Category:Organometallic chemistry]]
[[Category:Organometallic chemistry]]

Revision as of 06:30, 30 December 2010

An insertion reaction is a chemical reaction where one chemical entity (a molecule or molecular fragment) interposes itself into an existing bond of typically a second chemical entity e.g.:

A + B–CB–A–C

The term only refers to the result of the reaction and does not suggest a mechanism. Insertion reactions are observed in organic, inorganic, and organometallic chemistry. In cases where a metal-ligand bond in a coordination complex is involved, these reactions are typically organometallic in nature and involve a bond between a transition metal and a carbon or hydrogen.[1] It is usually reserved for the case where the coordination number and oxidation state of the metal remain unchanged.[2] When these reactions are reversible, the removal of the small molecule from the metal-ligand bond is called extrusion or elimination.

Examples of type 1,1 (a) and 1,2 (b) resulting geometries for insertion reactions

There are two common insertion geometries— 1,1 and 1,2 (pictured above). Additionally, the inserting molecule can act either as a nucleophile or as an electrophile to the metal complex.[2] These behaviors will be discussed in more detail for CO, nucleophilic behavior, and SO2, electrophilic behavior.

Organic chemistry

Homologation reactions like the Kowalski ester homologation[3] provide simple examples of insertion process in organic synthesis. In the Arndt-Eistert reaction,[4][5] a methylene unit is inserted into the carboxyl-carbon bond of carboxylic acid to form the next acid in the homologous series. Organic Syntheses provides the example of t-BOC protected (S)-phenylalanine (2-amino-3-phenylpropanoic acid) being reacted sequentially with triethylamine, ethyl chloroformate, and diazomethane to produce the α-diazoketone, which is then reacted with silver trifluoroacetate / triethylamine in aqueous solution to generate the t-BOC protected form of (S)-3-amino-4-phenylbutanoic acid.[6]

Homologation of N-boc-phenylalanine

Mechanistically,[7] the α-diazoketone undergoes a Wolff rearrangement[8][9] to form a ketene in a 1,2-rearrangement. Consequently, the methylene group α- to the carboxyl group in the product is the methylene group from the diazomethane reagant. The 1,2-rearrangement has been shown to conserve the stereochemisty of the chiral centre as the product formed from t-BOC protected (S)-phenylalanine retains the (S) stereochemistry with a reported enantiomeric excess of at least 99%.[6]

A related transformation is the Nierenstein reaction in which a diazomethane methylene group is inserted into the carbon-chlorine bond of an acid chloride to generate an α-chloromethyl ketone.[10][11] An example, published in 1924, illustrates the reaction in a substituted benzoyl chloride system:[12]

Nierenstein 1924

Perhaps surprisingly, α-bromoacetophenone is the minor product when this reaction is carried out with benzoyl bromide, a dimeric dioxane being the major product.[13] Organic azides also provide an example of an insertion reaction in organic synthesis and, like the above examples, the transformations proceed with loss of nitrogen gas. When tosyl azide reacts with norbornadiene, a ring expansion reaction takes place in which a nitrogen atom is inserted into a carbon-carbon bond α- to the bridge head:[14]

Norbornadiene reaction with tosyl azide

The Beckmann rearrangement[15][16] is another example of a ring expanding reaction in which a heteroatom is inserted into a carbon-carbon bond. The most important application of this reaction is the conversion of cyclohexanone to its oxime, which is then rearranged under acidic conditions to provide ε-caprolactam,[17] the feedstock for the manufacture of Nylon 6. Annual production of caprolactam exceeds 2 billion kilograms.[18]

The Beckmann Rearrangement

Carbenes undergo both intermolecular and intramolecular insertion reactions. Cyclopentene moieties can be generated from sufficiently long-chain ketones by reaction with trimethylsilyldiazomethane, (CH3)3Si–CHN2:

Alkylidene carbene

Here, the carbene intermediate inserts into a carbon-hydrogen bond to form the carbon-carbon bond needed to close the cyclopentene ring. Carbene insertions into carbon-hydrogen bonds can also occur intermolecularly:

Carbene intermolecular reaction

Carbenoids are reactive intermediates that behave similarly to carbenes.[19] One example is the chloroalkyllithium carbenoid reagent prepared in situ from a sulfoxide and t-BuLi which inserts into the carbon-boron bond of a pinacol boronic ester:[20]

Insertion of carbenoid into carbon-boron bond

CO Insertion

The insertion of carbon monoxide across a metal-carbon site to form an acetyl group is the oldest-known and most-studied metal-ligand insertion reaction. It proceeds by a 1,1 reaction coordinate, attaching the carbonyl carbon to both the metal and the ligand. The first CO insertion was discovered in 1957 by the reaction of CO with Mn(CO)5CH3, forming Mn(CO)5COCH3.

Mechanism

The mechanism for the apparent CO insertion into a metal-alkyl bond is actually a migratory insertion, with a migration of the alkyl group to another bound CO, followed by addition of a free CO (see figure below).[21] This can be demonstrated by 13C-labeling the incoming CO ligand, which results in 100% of the labeled CO residing cis- to the acetyl group.

CO Insertion reaction pathway for an octahedral complex

The CO insertion mechanism is not always a migration. The reaction of CO with (Cp)MeLFeCO, where L is a nucleophilic group such as PPh3, yields a mix of both alkyl migration products and products formed by true insertion of bound carbonyls into the methyl group, which is controllable by the choice of solvent.[22]

Square planar complexes can also undergo CO insertions. Insertion reactions in square planar complexes are of particular interest because their structure allows additional reaction mechanisms to occur. While just like octahedral complexes, square planar complexes can undergo in-plane migration, their lack of out-of-plane steric hindrance renders them much more open to nucleophilic attack of the metal by the CO. Since square planar groups usually form 16 electron species, the 5-coordinate intermediate that forms is stabilized by the 18-Electron rule, and undergoes migratory insertion readily.[22] In most cases the in-plane migration pathway is preferred, but, unlike the nucleophilic pathway, it is inhibited by an excess of CO.[23]

Nucleophilic insertion and rearrangement of a square planar complex

Effects on reaction rates

  • Steric strain – Increasing the steric strain of the chelate backbone in square planar complexes pushes the carbonyl and methyl groups closer together, increasing the reactivity of insertion reactions.[23]
  • Oxidation state – Oxidation of the metal tends to increase insertion reaction rates. The main rate-limiting step in the mechanism is the migration of the methyl group onto a carbonyl ligand, oxidizing the metal by imparting a greater partial positive charge on the acetyl carbon, and thus increasing the rate of reaction.[2]
  • Lewis acids – Lewis acids also increase the reaction rates, for reasons similar to metal oxidation increasing the positive charge on the carbon. Lewis acids bind to the CO oxygen and remove charge, increasing the electrophilicity of the carbon. This can increase the reaction rate by a factor of up to 108, and the complex formed is stable enough that the reaction proceeds even without additional CO to bind to the metal.[2]
  • Electronegativity of the leaving group - Increasing the electronegativity of the leaving alkyl group stabilizes the metal-carbon bond interaction and thus increases the activation energy required for migration, decreasing the reaction rate.[24]
  • Trans-effect – Ligands in an octahedral or square planar complex are known to influence the reactivity of the group to which they are trans-. This ligand influence is often referred to as the trans-influence, and it varies in intensity between ligands. A partial list of trans-influencing ligands is as follows, from highest trans-effect to lowest:[22] aryl, alkyl > NR3 > PR3 > AsR3 > CO > Cl. Ligands with a greater trans-influence impart greater electrophilicity to the active site. Increasing the electrophilicity of the CO group has been shown experimentally to greatly increase the reaction rate, while decreasing the electrophilicity of the methyl group slightly increases the reaction rate. This can be demonstrated by reacting a square planar [(PN)M(CO)(CH3)] complex with CO, where PN is a bidentate phosphorus- or nitrogen-bound ligand. This reaction proceeds in much greater yield when the methyl group is trans-P and the CO trans-N, owing to the higher trans-influence of the more electronegative nitrogen.[23]

Reverse reaction

Decarbonylation of aldehydes, the reverse of CO insertion, is much more difficult to perform than the associated carbonylation as it requires both a cis- empty site and enough energy to break a carbon-carbon bond. Additionally, such reactions are usually not catalytic, as the extruded CO binds too tightly to the metal centre to be dissociated even under the high heats required for this reaction.[25] Extrusion of CO from an organic aldehyde is most famously demonstrated using Wilkinson's catalyst, which undergoes oxidative addition of the aldehyde C–H instead of H–H, and then undergoes CO extrusion to form an octahedral product.[26]

Industrial applications

The most widely known application of migratory insertion of carbonyl groups is the Monsanto acetic acid process, which uses an rhodium-iodine catalyst system to transform methanol into acetic acid under relatively mild, catalytic conditions. It has been more recently superceded by the Cativa process which uses the cis-dicarbonyldiiodoiridate(I) anion [Ir(CO)2I2] (1) as catalyst.[27][28] By 2002, worldwide annual production of acetic acid stood at 6 million tons, of which approximately 60% is produced by the Cativa process.[27]

The catalytic cycle of the Cativa process
The catalytic cycle of the Cativa process

The catalytic cycle for the Cativa process, shown above, begins with the reaction of methyl iodide with the square planar active catalyst species (1) to form the octahedral iridium(III) species (2), the fac-isomer of [Ir(CO)2(CH3)I3]. This oxidative addition reaction involves the formal insertion of the iridium(I) centre into the carbon-iodine bond of methyl iodide. After ligand exchange, the migratory insertion of carbon monoxide into the iridium-carbon bond, step (3) to (4), results in the formation of a square pyramidal species with a bound acetyl ligand. The active catalyst species (1) is regenerated by the reductive elimination of acetyl iodide from (4), a de-insertion reaction.[27]

SO2 Insertion

SO2 is an electrophilic species. When SO2 encounters a transition metal with alkyl ligands, it acts like a Lewis acid towards the alkyl ligand.[1]

Mechanism

The mechanism of SO2 insertion is shown in the figure below.[29] First, the sulfur coordinates to the ligand with the electronegative oxygens pointing away from the complex. Then, two different pathways can occur to “insert” the sulfur dioxide between the metal and the alkyl ligand. The top pathway shows the creation of an ionic complex, O,O’-sulphinate where the metal has a positive charge and the ligand has a negative charge. The bottom pathway shows the creation of an O-sulphinate, which can be a stable product. Also, either of these intermediates can form an S-sulphinate, the most common SO2 insertion species and a 1,1 type geometry. IR spectroscopy and NMR can determine the type of compound.[29] S-sulphinate has sulfur-oxygen stretching frequencies from 1250–1000 cm−1 and 1100–1000 cm−1. The O, O'-sulphinate and O-sulphinate are difficult to distinguish as they have stretching frequencies from 1085–1050 cm−1 and 1000–820 cm−1 or lower. The pathway involving the O, O' sulphinate can generally be ruled out if the original metal complex fulfilled the 18 electron rule because the two metal-oxygen bonds would exceed the 18 electron rule.[2]

Mechanism of SO2 insertion

When SO2 is inserting into a square planar complex that does not have 18 electrons, it inserts slightly differently. Since the metal is coordinatively unsaturated, the SO2 can pre-coordinate with the metal. Then, it inserts itself between the metal and the alkyl group as shown in the example below.[30]

Mechanism of SO2 Insertion Into a Square Planar Complex

Effects on reaction rates

The ability of SO2 to insert depends upon the electronic interactions between the metal and the ligand. If there is pi bonding between the ligand and metal, the SO2 will not insert. When steric effects are negligible, the ability of SO2 to insert increases with an increase in the electron donating capacity of the ligand and decreases with the electron withdrawing capacity of the ligand. These trends highlight the electrophilic nature of SO2 insertion.[29]

Reverse reaction

The insertion of SO2 is rarely reversible.[1] Often, in trying to desulfinate a compound, the whole compound is destroyed.[29] Some exceptions include: thermal/photochemical extrusion in [CpFe(CO2{S(O)2C6F5}] in toluene and thermal extrusion from [CpFe(CO){P(OPh)3}{S(O)2CH(Ph)(SiMe3)}] under vacuum.[2] The reason that desulfination does not easily occur may be that when there is ligand expulsion, instead of the alkyl group migrating as happens in the case of CO, the SO2 bonds again with the metal to form an O,O’ sulphinate.[2]

Industrial applications

There are few examples of SO2 insertion in industry due to the low reactivity of SO2 insertions and the low stability of the created products, but SO2 insertion can be a step in production of sulfolanes.[31][32]

Olefin Insertion

Olefin insertions are insertion reactions involving alkenes and alkynes and they can insert into metal-carbon as well as metal-hydrogen bonds. Many of olefin insertions' features are similar to those of carbonyl insertion in terms of the mechanism, however key differences are present. One of these is the potential to perform multiple insertions in tandem providing a mechanism for polymerization.[33]

Mechanism

The mechanism for olefin insertions into metal-carbon or metal-hydrogen bonds converts an olefin complex into an alkyl complex, forming a bond between the carbon or hydrogen with the carbon at the β position. This insertion at the same time forms a bond between the metal and the carbon at the α position. When a metal has sufficient electrons for π-backbonding,[34] the olefin can be coordinated to the metal before insertion. Depending on the ligand density of the metal, ligand dissociation may be necessary to provide a coordination site for the olefin.

Olefin insertion (forward reaction) and beta-elimination (reverse reaction)

Fundamental features of olefin insertions include the mechanism's dependance on the electron density of the metal complex it is inserting into. For instance, a coordinatively saturated 18-electron metal complex will likely require ligand dissociation to provide a vacant coordination site for the insertion to occur (this is also true for the β-elimination[35]). In the case of coordinatively unsaturated metal complexes with lower electron counts (such as the 14-electron Cp*2ScCH3 complex), there are already vacant coordination sites available for olefin insertion without the need for ligand dissociation.

Effects on reaction rates

Factors affecting the rate of olefin insertions include the formation of the cyclic, planar, four-center transition state involving incipient formation of a bond between the metal and an olefin carbon.

Transition state of olefin insertion reaction

From this transition state, it can be seen that a partial positive charge forms on the β-position carbon with a partial negative charge formed on the carbon or hydrogen initially bonded to the metal. This explains the subsequently observed formation of the bond between the negatively charged carbon/hydrogen and the positively charged β-carbon as well as the simultaneously formation of the metal-α-carbon bond.

This transition state also allows for consideration of the two most contributing factors to the rate of olefin insertion reactions. These include the orbital overlap of the alkyl group initially attached to the metal and the strength of the metal-alkyl bond. With greater orbital overlap between the partially positive β-carbon and the partially negative hydrogen/alkyl group carbon, the formation of a bond between them is facilitated and the insertion reaction rate is increased. With increasing strength of the metal-alkyl bond, the breaking of the bond between the metal and the hydrogen/alkyl carbon bond to form the two new bonds with the α-carbon and β-carbon (respectively) is slower, thus decreasing the rate of the insertion reaction.[36]

Reverse reaction

The reverse mechanism for olefin insertion into a metal-hydrogen bond is β-elimination, which can only occur if a vacant orbital and a vacant coordination position are available in addition to the presence of a hydrogen at the β position.[37] If these conditions exist, the decomposition of the metal alkyl complex can take place resulting in a hydride/olefin complex that can subsequently dissociate the olefin.

Industrial applications

As mentioned previously, a fundamental feature of olefin insertion is the opportunity for multiple insertions, providing a mechanism for polymerization. An industrial application of this is ethylene polymerization, which is catalyzed heterogeneously by titanium(III) salts and aluminum alkyls (also known as Ziegler-Natta catalysts).[38] In these reactions, ethylene is coordinated to a vacant site on the titanium metal followed by its insertion. This can be repeated multiple times, potentially leading to high molecular weight polymers.

Synthetic pathway of ethylene polymerization via olefin insertion reaction

Other Insertion reactions

CO2 insertion could be important for use in feedstocks. CO2 prefers insertion into metal-nitrogen bonds rather than metal-carbon bonds. Most CO2 extrusion reactions can occur thermally. Isocyanides (RNC) and nitrosyls (NO) insert similarly to CO in many ways including the same reaction mechanism for insertion into a square planar complex. Insertions of unsaturated species such as O2, S, carbenes, and nitrenes have also been observed.[2]

References

  1. ^ a b c Douglas, McDaniel, and Alexander (1994). Concepts and Models of Inorganic Chemistry 3rd Ed. John Wiley & Sons, Inc. ISBN 9780471629788.{{cite book}}: CS1 maint: multiple names: authors list (link)
  2. ^ a b c d e f g h J.J. Alexander (1985). Hartley and Patai (ed.). The chemistry of the metal-carbon bond, vol. 2. John Wiley & Sons.
  3. ^ Reddy, R. E.; Kowalski, C. J. (1993). "Ethyl 1-Naphthylacetate: Ester Homologation Via Ynolate Anions". Organic Syntheses. 71: 146; Collected Volumes, vol. 9, p. 426.
  4. ^ Arndt, F.; Eistert, B. (1935). "Ein Verfahren zur Überführung von Carbonsäuren in ihre höheren Homologen bzw. deren Derivate". Ber. Dtsch. Chem. Ges. (in German). 1 (68): 200–208. doi:10.1002/cber.19350680142.
  5. ^ Ye, T.; McKervey, M. A. (1994). "Organic Synthesis with α-Diazo Carbonyl Compounds". Chem. Rev. 94 (4): 1091–1160. doi:10.1021/cr00028a010.
  6. ^ a b Linder, M. R.; Steurer, S.; Podlech, J. (2002). "(S)-3-(tert-Butyloxycarbonylamino)-4-phenylbutanoic acid". Organic Syntheses. 79: 154; Collected Volumes, vol. 10, p. 194.
  7. ^ Huggett, C.; Arnold, R. T.; Taylor, T. I. (1942). "The Mechanism of the Arndt-Eistert Reaction". J. Amer. Chem. Soc. 64 (12): 3043. doi:10.1021/ja01264a505.
  8. ^ Meier, H.; Zeller, K.-P. (1975). "The Wolff Rearrangement of α-Diazo Carbonyl Compounds". Angew. Chem. Int. Ed. 14 (1): 32–43. doi:10.1002/anie.197500321.
  9. ^ Kirmse, W. (2002). "100 Years of the Wolff Rearrangement". Eur. J. Org. Chem. 2002 (14): 2193–2256. doi:10.1002/1099-0690(200207)2002:14<2193::AID-EJOC2193>3.0.CO;2-D.
  10. ^ Clibbens, D. A.; Nierenstein, M. (1915). "The Action of Diazomethane on some Aromatic Acyl Chlorides". J. Chem. Soc., Trans. 107: 1491–1494. doi:10.1039/CT9150701491.
  11. ^ Bachmann, W. E.; Struve, W. S. (1942). "The Arndt-Eistert Reaction". Org. React. 1: 38.
  12. ^ Nierenstein, M.; Wang, D. G.; Warr, J. C. (1924). "The Action of Diazomethane on some Aromatic Acyl Chlorides II. Synthesis of Fisetol". J. Amer. Chem. Soc. 46 (11): 2551–2555. doi:10.1021/ja01676a028.
  13. ^ Lewis, H. H.; Nierenstein, M.; Rich, E. M. (1925). "The Action of Diazomethane on some Aromatic Acyl Chlorides III. The Mechanism of the Reaction". J. Amer. Chem. Soc. 47 (6): 1728–1732. doi:10.1021/ja01683a036.
  14. ^ Reed, D. D.; Bergmeier, S. C. (2007). "A Facile Synthesis of a Polyhydroxylated 2-Azabicyclo[3.2.1]octane". J. Org. Chem. 72 (3): 1024–1026. doi:10.1021/jo0619231. PMID 17253828.
  15. ^ Beckmann, E. (1886). "Zur Kenntniss der Isonitrosoverbindungen". Ber. Dtsch. Chem. Ges. (in German). 19: 988–993. doi:10.1002/cber.188601901222.
  16. ^ Gawley, R. E. (1988). "The Beckmann Reactions: Rearrangement, Elimination-Additions, Fragmentations, and Rearrangement-Cyclizations". Org. React. 35: 14–24. doi:10.1002/0471264180.or035.01.
  17. ^ Eck, J. C.; Marvel, C. S. (1939). "ε-Benzoylaminocaproic acid". Organic Syntheses. 19: 20; Collected Volumes, vol. 2, 1943, p. 76.
  18. ^ Ritz, J.; Fuchs, H.; Kieczka, H.; Moran, W. C. (2000). "Caprolactam". Ullmann's Encyclopedia of Industrial Chemistry. Weinheim: Wiley-VCH. doi:10.1002/14356007.a05_031.
  19. ^ McMurry, J. (1988). Organic Chemistry (2nd ed.). Brooks/Cole. ISBN 0-534-07968-7.
  20. ^ Blakemore, P. R.; Burge, M. S. (2007). "Iterative Stereospecific Reagent-Controlled Homologation of Pinacol Boronates by Enantioenriched-Chloroalkyllithium Reagents". J. Amer. Chem. Soc. 129 (11): 3068–3069. doi:10.1021/ja068808s.
  21. ^ Yadav, M. S. (2005). Quick Review in Inorganic Chemistry. Anmol Publications. p. 244. ISBN 9788126118984.
  22. ^ a b c Anderson, G. K.; Cross, R. J. (1984). "Carbonyl-Insertion Reactions of Square Planar Complexes". Acc. Chem. Res. 17 (2): 67–74. doi:10.1021/ar00098a005.
  23. ^ a b c Cavell, K. J. (1996). "Recent Fundamental Studies on Migratory Insertion into Metal-Carbon Bonds". Coord. Chem. Rev. 155 (11): 209–243. doi:10.1016/S0010-8545(96)90182-4.
  24. ^ Shusterman, A. J.; Tamir, I.; Pross, A. (1988). "The Mechanism of Organometallic Migration Reactions. A Configuration Mixing (CM)Approach". J. Organomet. Chem. 340 (2): 203–222. doi:10.1016/0022-328X(88)80076-7.
  25. ^ Beck, C. M.; Rathmill, S. E.; Park, Y. J.; Chen, J.; Crabtree, R. H.; Liable-Sands, L. M.; Rheingold, A. L. (1999). "Aldehyde Decarbonylation Catalysis under Mild Conditions". Organometallics. 18 (25): 5311–5317. doi:10.1021/om9905106.
  26. ^ Ohno, K.; Tsuji, J. (1968). "Organic Synthesis by means of Noble Metal Compounds. XXXV. Novel Decarbonylation Reactions of Aldehydes and Acyl Halides using Rhodium Complexes". J. Amer. Chem. Soc. 90 (1): 99–107. doi:10.1021/ja01003a018.
  27. ^ a b c Jones, J. H. (2000). "The CativaTM Process for the Manufacture of Acetic Acid" (PDF). Platinum Metals Rev. 44 (3): 94–105.
  28. ^ Sunley, G. J.; Watson, D. J. (2000). "High Productivity Methanol Carbonylation Catalysis using Iridium - The CativaTM Process for the Manufacture of Acetic Acid". Catalysis Today. 58 (4): 293–307. doi:10.1016/S0920-5861(00)00263-7.
  29. ^ a b c d A. Wojcicki (1974). Stone and West (ed.). Advances in Organometallic Chemistry. Academic Press.
  30. ^ Puddephatt, R.A. (1980). "Competition between Insertion of Sulfur Dioxide into the Methyl- or Phenyl- Transition Metal Bond". Journal of Organometallic Chemistry. 193: C27. doi:10.1016/S0022-328X(00)86091-X. {{cite journal}}: Unknown parameter |coauthors= ignored (|author= suggested) (help)
  31. ^ Kurosawa, H.; Yamamoto, A., ed. (2003). Fundamentals of Molecular Catalysis, Volume 3 (Current Methods in Inorganic Chemistry). Elsevier. ISBN 0444509216.{{cite book}}: CS1 maint: multiple names: editors list (link)
  32. ^ Dzhemilev, K (1993). "Metal complex catalysis in the synthesis of organic sulfur compounds". Journal of Organometallic Chemistry. 455. doi:10.1016/0022-328X(93)80375-L. {{cite journal}}: Unknown parameter |coauthors= ignored (|author= suggested) (help)
  33. ^ Sinn, H. (1980). Advances in Organometallic Chemistry. 18: 99. {{cite journal}}: Missing or empty |title= (help); Unknown parameter |coauthors= ignored (|author= suggested) (help)
  34. ^ Crabtree, R. H. (2009). The Organometallic Chemistry of the Transition Metals. John Wiley and Sons. p. 19–25. ISBN 9780470257623.
  35. ^ Reger, D. L.; Culbertson, E. C. (1976). "Mechanism of the Thermal Decomposition of Carbonyl(η5-cyclopentadienyl)triphenylphosphineiron(II) Alkyl Derivatives into Carbonyl(η5-cyclopentadienyl)hydrido(triphenylphosphine)iron(II) and Olefin". J. Amer. Chem. Soc. 98 (10): 2789–2794. doi:10.1021/ja00426a020.
  36. ^ Burger, B. J.; Thompson, M. E.; Cotter, W. D.; Bercaw, J. E. (1990). "Ethylene Insertion and β-hydrogen Elimination for Permethylscandocene Alkyl Complexes. A Study of the Chain Propagation and Termination Steps in Ziegler-Natta Polymerization of Ethylene". J. Amer. Chem. Soc. 112 (4): 1566–1577. doi:10.1021/ja00160a041.
  37. ^ Crabtree, R. H. (2009). The Organometallic Chemistry of the Transition Metals. John Wiley and Sons. p. 192. ISBN 9780470257623.
  38. ^ Kissin, Y. V. (2008). "Synthesis, Chemical Composition, and Structure of Transition Metal Components and Cocatalysts in Catalyst Systems for Alkene Polymerization". Alkene Polymerization Reactions with Transition Metal Catalysts. Amsterdam: Elsevier. pp. 207–290. ISBN 9780444532152.