Truxillic acid

Last updated
Truxillic acid
Truxillic acid.svg
Names
IUPAC name
7,8′-Cyclo-8,7′-neolignane-9,9′-dioic acid
Systematic IUPAC name
2,4-Diphenylcyclobutane-1,3-dicarboxylic acid
Identifiers
3D model (JSmol)
ChemSpider
ECHA InfoCard 100.022.478 OOjs UI icon edit-ltr-progressive.svg
PubChem CID
UNII
  • OC(C1C(C2=CC=CC=C2)C(C(O)=O)C1C3=CC=CC=C3)=O
Properties
C18H16O4
Molar mass 296.322 g·mol−1
Except where otherwise noted, data are given for materials in their standard state (at 25 °C [77 °F], 100 kPa).

Truxillic acids are any of several crystalline stereoisomeric cyclic dicarboxylic acids with the formula (C6H5C2H2(CO2H)2. They are colorless solids. These compounds are obtained by the [2 + 2] photocycloadditions of cinnamic acid where the two trans alkenes react head-to-tail. The isolated stereoisomers are called truxillic acids. [1] The preparation of truxillic acids provided an early example of organic photochemistry. [2]

Contents

Cinnamic Acid CycloAddition CinnamicAcidCycloAddition.png
Cinnamic Acid CycloAddition

Occurrence and reactions

These compounds are found in a variety of plants, for example in coca. [3] [4] Incarvillateine, an alkaloid from the plant Incarvillea sinensis , is a derivative of α-truxillic acid.

Upon heating, truxillic acids undergo cracking to give cinnamic acid. [5]

Isomers

Truxillic acid can exist in five stereoisomers. [6] [7]

Truxillic and truxinic acid stereo.svg
Truxillic acid isomers
Isomerabcdef
α-truxillic acid
(cocaic acid [8] )
COOHHHC6H5HCOOH
γ-truxillic acidCOOHHHC6H5COOHH
ε-truxillic acidHCOOHC6H5HHCOOH
peri-truxillic acidCOOHHC6H5HCOOHH
epi-truxillic acidCOOHHC6H5HHCOOH

Below are the five stereoisomers of truxillic acid, called alpha, gamma, epsilon, peri, and epi. These are shown both in a 2D skeletal diagram with stereocenters indicated and a 3D rendering of the structural geometry of the isomers themselves.

Truxillic acids.svg
Truxillic acids 3D BS.png

See also

Related Research Articles

<span class="mw-page-title-main">Pinner reaction</span> Reaction of cyanide and alcohol to give imino ester salt

The Pinner reaction refers to the acid catalysed reaction of a nitrile with an alcohol to form an imino ester salt ; this is sometimes referred to as a Pinner salt. The reaction is named after Adolf Pinner, who first described it in 1877. Pinner salts are themselves reactive and undergo additional nucleophilic additions to give various useful products:

Pyrrole is a heterocyclic, aromatic, organic compound, a five-membered ring with the formula C4H4NH. It is a colorless volatile liquid that darkens readily upon exposure to air. Substituted derivatives are also called pyrroles, e.g., N-methylpyrrole, C4H4NCH3. Porphobilinogen, a trisubstituted pyrrole, is the biosynthetic precursor to many natural products such as heme.

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

Cinnamic acid is an organic compound with the formula C6H5-CH=CH-COOH. It is a white crystalline compound that is slightly soluble in water, and freely soluble in many organic solvents. Classified as an unsaturated carboxylic acid, it occurs naturally in a number of plants. It exists as both a cis and a trans isomer, although the latter is more common.

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

Resorcinol (or resorcin) is a phenolic compound. It is an organic compound with the formula C6H4(OH)2. It is one of three isomeric benzenediols, the 1,3-isomer (or meta-isomer). Resorcinol crystallizes from benzene as colorless needles that are readily soluble in water, alcohol, and ether, but insoluble in chloroform and carbon disulfide.

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

Erythorbic acid is a stereoisomer of ascorbic acid. It is synthesized by a reaction between methyl 2-keto-D-gluconate and sodium methoxide. It can also be synthesized from sucrose or by strains of Penicillium that have been selected for this feature. It is denoted by E number E315, and is widely used as an antioxidant in processed foods.

In organic chemistry, the Knoevenagel condensation reaction is a type of chemical reaction named after German chemist Emil Knoevenagel. It is a modification of the aldol condensation.

The Japp–Klingemann reaction is a chemical reaction used to synthesize hydrazones from β-keto-acids and aryl diazonium salts. The reaction is named after the chemists Francis Robert Japp and Felix Klingemann.

The Bischler–Möhlau indole synthesis, also often referred to as the Bischler indole synthesis, is a chemical reaction that forms a 2-aryl-indole from an α-bromo-acetophenone and excess aniline; it is named after August Bischler and Richard Möhlau .

<span class="mw-page-title-main">Reimer–Tiemann reaction</span> Chemical reaction for ortho-formylation of phenols

The Reimer–Tiemann reaction is a chemical reaction used for the ortho-formylation of phenols. with the simplest example being the conversion of phenol to salicylaldehyde. The reaction was first reported by Karl Reimer and Ferdinand Tiemann.

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

'DibenzylideneacetoneDibenzylideneacetone' or dibenzalacetone, often abbreviated dba, is an organic compound with the formula C17H14O. It is a pale-yellow solid insoluble in water, but soluble in ethanol.

<span class="mw-page-title-main">Hofmann–Martius rearrangement</span>

The Hofmann–Martius rearrangement in organic chemistry is a rearrangement reaction converting an N-alkylated aniline to the corresponding ortho and / or para aryl-alkylated aniline. The reaction requires heat, and the catalyst is an acid like hydrochloric acid.

The Koenigs–Knorr reaction in organic chemistry is the substitution reaction of a glycosyl halide with an alcohol to give a glycoside. It is one of the oldest glycosylation reactions. It is named after Wilhelm Koenigs (1851–1906), a student of von Baeyer and fellow student with Hermann Emil Fischer, and Edward Knorr, a student of Koenigs.

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

Fischer glycosidation refers to the formation of a glycoside by the reaction of an aldose or ketose with an alcohol in the presence of an acid catalyst. The reaction is named after the German chemist, Emil Fischer, winner of the Nobel Prize in chemistry, 1902, who developed this method between 1893 and 1895.

<span class="mw-page-title-main">Wolffenstein–Böters reaction</span> Organic reaction converting benzene to picric acid

The Wolffenstein–Böters reaction is an organic reaction converting benzene to picric acid by a mixture of aqueous nitric acid and mercury(II) nitrate.

The Willgerodt rearrangement or Willgerodt reaction is an organic reaction converting an aryl alkyl ketone, alkyne, or alkene to the corresponding amide by reaction with ammonium polysulfide, named after Conrad Willgerodt. The formation of the corresponding carboxylic acid is a side reaction resulting from hydrolysis of the amide. When the alkyl group is an aliphatic chain, multiple reactions take place with the amide group always ending up at the terminal end. The net effect is thus migration of the carbonyl group to the end of the chain and oxidation.

The Schotten–Baumann reaction is a method to synthesize amides from amines and acid chlorides:

The Glaser coupling is a type of coupling reaction. It is by far the oldest acetylenic coupling and is based on cuprous salts like copper(I) chloride or copper(I) bromide and an additional oxidant like oxygen. The base in its original scope is ammonia. The solvent is water or an alcohol. The reaction was first reported by Carl Andreas Glaser in 1869. He suggested the following process for his way to diphenylbutadiyne:

In organic chemistry, the Claisen–Schmidt condensation is the reaction between an aldehyde or ketone having an α-hydrogen with an aromatic carbonyl compound lacking an α-hydrogen. It can be considered as a specific variation of the aldol condensation. This reaction is named after two of its pioneering investigators Rainer Ludwig Claisen and J. Gustav Schmidt, who independently published on this topic in 1880 and 1881. An example is the synthesis of dibenzylideneacetone ( -1,5-diphenylpenta-1,4-dien-3-one).

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

Spiropentane is a hydrocarbon with formula C5H8. It is the simplest spiro-connected cycloalkane, a triangulane. It took several years after the discovery in 1887 until the structure of the molecule was determined. According to the nomenclature rules for spiro compounds, the systematic name is spiro[2.2]pentane. However, there can be no constitutive isomeric spiropentanes, hence the name is unique without brackets and numbers.

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

Truxinic acids are any of several stereoisomeric cyclic dicarboxylic acids with the formula (C6H5)2C4H4(COOH)2, found in various plants. They are obtained by a photochemical cycloaddition from cinnamic acid, where the two trans alkenes react head-to-head.

References

  1. Cohen, M. D.; Schmidt, G. M. J.; Sonntag, F. I. (1964). "Topochemistry. II. The photochemistry of trans-cinnamic acids". J. Chem. Soc.: 2000–2013. doi:10.1039/jr9640002000.
  2. Roth, Heinz D. (1989). "The Beginnings of Organic Photochemistry". Angewandte Chemie International Edition in English. 28 (9): 1193–1207. doi:10.1002/anie.198911931.
  3. Liebermann, C. (1888). "Ueber Cinnamylcocaïn". Berichte der Deutschen Chemischen Gesellschaft. 21 (2): 3372–3376. doi:10.1002/cber.188802102223.
  4. Krauze-Baranowska, Miroslawa (2002). "Truxillic and truxinic acids-occurrence in plant kingdom". Acta Poliniae Pharmaceutica-Drug Research. 59 (5): 403–410. PMID   12602803.
  5. Hein, Sara M. (2006). "An Exploration of a Photochemical Pericyclic Reaction Using NMR Data". Journal of Chemical Education. 83 (6): 940–942. Bibcode:2006JChEd..83..940H. doi:10.1021/ed083p940.
  6. Stoermer, R.; Bachér, F. (1924). "Zur Stereoisomerie der Truxillsäuren und über die Auffindung der letzten Säure dieser Gruppe (VIII.)". Berichte der Deutschen Chemischen Gesellschaft (A and B Series). 57: 15–23. doi:10.1002/cber.19240570105.
  7. Agarwai, O. P. (2011). Organic Chemistry Reactions and Reagents. Krishna Prakashan Media. ISBN   978-8187224655.
  8. "ChemSpider ID 10218892". ChemSpider . Retrieved 15 October 2016.