Itaconic anhydride

Last updated
Itaconic anhydride
Itaconsaureanhydrid Struktur.svg
Names
Preferred IUPAC name
3-Methylideneoxolane-2,5-dione
Other names
Methylenesuccinic anhydride
Identifiers
3D model (JSmol)
ChEMBL
ChemSpider
ECHA InfoCard 100.016.835 OOjs UI icon edit-ltr-progressive.svg
EC Number
  • 218-518-2
PubChem CID
UNII
  • InChI=1S/C5H4O3/c1-3-2-4(6)8-5(3)7/h1-2H2
    Key: OFNISBHGPNMTMS-UHFFFAOYSA-N
  • C=C1CC(=O)OC1=O
Properties
C5H4O3
Molar mass 112,09 g·[mol −1
Appearancecolorless crystalline solid [1]
Melting point 70–72 °C (158–162 °F; 343–345 K) [2]
soluble in acetone and chloroform, only slightly soluble in Diethylether, [3] reacts with water
Hazards
GHS labelling:
GHS-pictogram-exclam.svg
Warning
H302, H315, H319, H335
P261, P264, P270, P271, P280, P301+P312, P302+P352, P304+P340, P305+P351+P338, P312, P321, P330, P332+P313, P337+P313, P362, P403+P233, P405, P501
Except where otherwise noted, data are given for materials in their standard state (at 25 °C [77 °F], 100 kPa).

Itaconic anhydride is the cyclic anhydride of itaconic acid (an unsaturated, dicarboxylic acid) and is obtained by the pyrolysis of citric acid. It is a colourless, crystalline solid, which dissolves in many polar organic solvents and hydrolyzes forming itaconic acid. [4] Itaconic anhydride and its derivative itaconic acid have been promoted as biobased "platform chemicals" and bio- building blocks. [5] [6] ) These expectations, however, have not been fulfilled. [7]

Contents

Production

As discovered as early as 1836, attempted distillation of citric acid gives the so-called "pyrocitric acid" ("Brenzcitronensäure"), now known as itaconic anhydride. [8]

Synthese von Itaconsaureanhydrid aus Citronensaure Itaconsaureanhydrid aus Citronensaure.svg
Synthese von Itaconsäureanhydrid aus Citronensäure

According to an organic synthesis protocol, [4] itaconic anhydride is obtained from the rapid heating of citric acid monohydrate in a modest yield (37-47 %). The by-product is the thermodynamically more stable citraconic anhydride. [9]

Also when heating anhydrous citric acid to 260 °C in a vacuum, a mixture of itaconic and citraconic anhydride is achieved "in good yield". [10]

More productive are processes based on the biotechnologically accessible itaconic acid, [11] which produces exclusively itaconic anhydride in yields of up to 98% at temperatures of 165-180 °C and pressures of 10-30 mmHg in the presence of catalytic quantities of strong acids, such as concentrated sulphuric acid. [12]

Synthese von Itaconsaureanhydrid aus Itaconsaure Itaconsaureanhydrid aus Itaconsaure.svg
Synthese von Itaconsäureanhydrid aus Itaconsäure

In order to avoid overheating and thus higher proportions of citraconic anhydride, the dehydration reaction can also be carried out in higher boiling aromatic solvents such as toluene or xylene in the presence of acidic montmorillonite [13] or in cumene in the presence of methanesulfonic acid. [14] In both variants yields of 95-97 % of itaconic anhydride are achieved.

Another process of cyclizing dicarboxylic acids with diethyl carbonate in the presence of a chromium-salen complex with µ-nitrido-bis(triphenylphosphane) chloride as cocatalyst quantitatively provides itaconic anhydride contaminated with citraconic anhydride already at 40 °C in 1 millimolar preparations. However, the reaction is technically uninteresting because of its expensive catalysts. [15]

Reactions

At temperatures above its melting point, itaconic anhydride converts to citraconic anhydride. [12] Even at significantly lower temperatures, such as in boiling chloroform, isomerization can take place in the presence of tertiary amines. [16]

Umlagerung von Itaconsaureanhydrid in Citraconsaureanhydrid Citraconsaureanhydrid aus Itaconsaureanhydrid.svg
Umlagerung von Itaconsäureanhydrid in Citraconsäureanhydrid

By treating itaconic anhydride with phosphorus pentachloride (PCl5), itaconic acid dichloride (itaconyl chloride) is obtained: [17]

Synthese von Itaconylchlorid Itaconylchlorid Synthese.svg
Synthese von Itaconylchlorid

from which polyamides with reactive vinylidene groups can be formed with diamines. [18]

Bromination of itaconic anhydride at – 20 °C and subsequent dehydrobromination produces 2-bromomethylmaleic anhydride in 70% yield by shifting the double bond into the five-membered ring. [19]

Bromierung-Dehydrobromierung von Itaconsaureanhydrid Bromierung von Itaconsaureanhydrid.svg
Bromierung-Dehydrobromierung von Itaconsäureanhydrid

Otto Diels and Kurt Alder already described the addition (Diels-Alder reaction) of the dienophile itaconic anhydride to the diene cyclopentadiene in 1928. [20] Also furfuryl alcohol, which is accessible from renewable raw materials, reacts as a diene to form the Diels-Alder adduct, in which the reaction of the alcohol group with the cyclic anhydride structure forms a lactone and a carboxylic acid group, i.e. the cyclic half ester of itaconic acid. [21]

Diels-Alder-Reaktionen mit Itaconsaureanhydrid Diels-Alder-Reaktionen mit Itaconsaureanhydrid.svg
Diels-Alder-Reaktionen mit Itaconsäureanhydrid

Itaconic anhydride can react with aromatics such as benzene via Friedel-Crafts acylation. This always happens in such a way that the ring opening occurs at the carbonyl group, which is further away from the methylene group (3-position). [22]

Friedel-Crafts-Acylierung mit Itaconsaureanhydrid Friedel-Crafts-Acylierung mit Itaconsaureanhydrid.svg
Friedel-Crafts-Acylierung mit Itaconsäureanhydrid

Nucleophiles such as thiols can easily be added to the methylene group. With other nucleophiles, such as alcohols, ammonia, [23] amines and hydroxylamine, itaconic anhydride reacts regioselectively in position 3 to the corresponding esters, amides and hydroxamic acids.

Reaktionen von Itaconsaureanhydrid mit Nukleophilen Itaconsaureanhydrid Reaktionen mit Nukleophilen.svg
Reaktionen von Itaconsäureanhydrid mit Nukleophilen

The hydroxamic acid formed with O-benzylhydroxylamine can be cyclized in high yields with dicyclohexylcarbodiimide (DCC) to five-membered isoimides (iminofuranones) or with acetanhydride (Ac2O) to imides. [24]

ildung von Itaconimiden und Iminofuranonen Cyclisierungsreaktionen von Itaconhydroxamsauren.svg
ildung von Itaconimiden und Iminofuranonen

A number of five-, six- and seven-membered heterocycles (such as benzothiazepines) are obtainable from itaconic anhydride in useful yields. [25]

Bildung von Benzothiazepinessigsaure aus Itaconsaureanhydrid Benzothiazepinessigsaure aus Itaconsaureanhydrid.svg
Bildung von Benzothiazepinessigsäure aus Itaconsäureanhydrid

Polymers of itaconic anhydride

As an unsaturated cyclic anhydride, itaconic anhydride undergoes radical polymerization [26] and via polycondensation with diols or diamines. The two reactions can also be carried out sequentially – first radical polymerization, then polycondensation or vice versa. [27] [28]

Radically produced itaconic anhydride polymers and copolymers can be alkaline hydrolyzed to polyitaconic acids under ring opening or converted into polymeric acid amides or esters subsequent to polymerization. [29]

Copolymere von Itaconsaureanhydrid mit Stearylmethacrylat + Hydrolyse ITA-SMA-Copolymer.svg
Copolymere von Itaconsäureanhydrid mit Stearylmethacrylat + Hydrolyse

The resulting copolymers show properties that suggest a potential use as biomaterials for therapeutic systems and prostheses. [30]

Functional polymers exclusively from biogenic monomers involves the ring-opening metathesis polymerisation of an oxanorbornene ester produced from itaconic anhydride and furfuryl alcohol by Diels-Alder lactonisation using a Grubbs II catalyst. [31]

ROMP-Reaktion mit Itaconsaureanhydrid-Cyclopentadien/Furfurylalkohol-Diels-Alder-Addukt ROMP-Reaktion mit ITA-Diels-Alder-Addukt.svg
ROMP-Reaktion mit Itaconsäureanhydrid-Cyclopentadien/Furfurylalkohol-Diels-Alder-Addukt

Related Research Articles

<span class="mw-page-title-main">Diels–Alder reaction</span> Chemical reaction

In organic chemistry, the Diels–Alder reaction is a chemical reaction between a conjugated diene and a substituted alkene, commonly termed the dienophile, to form a substituted cyclohexene derivative. It is the prototypical example of a pericyclic reaction with a concerted mechanism. More specifically, it is classified as a thermally-allowed [4+2] cycloaddition with Woodward–Hoffmann symbol [π4s + π2s]. It was first described by Otto Diels and Kurt Alder in 1928. For the discovery of this reaction, they were awarded the Nobel Prize in Chemistry in 1950. Through the simultaneous construction of two new carbon–carbon bonds, the Diels–Alder reaction provides a reliable way to form six-membered rings with good control over the regio- and stereochemical outcomes. Consequently, it has served as a powerful and widely applied tool for the introduction of chemical complexity in the synthesis of natural products and new materials. The underlying concept has also been applied to π-systems involving heteroatoms, such as carbonyls and imines, which furnish the corresponding heterocycles; this variant is known as the hetero-Diels–Alder reaction. The reaction has also been generalized to other ring sizes, although none of these generalizations have matched the formation of six-membered rings in terms of scope or versatility. Because of the negative values of ΔH° and ΔS° for a typical Diels–Alder reaction, the microscopic reverse of a Diels–Alder reaction becomes favorable at high temperatures, although this is of synthetic importance for only a limited range of Diels-Alder adducts, generally with some special structural features; this reverse reaction is known as the retro-Diels–Alder reaction.

In organic chemistry, the Swern oxidation, named after Daniel Swern, is a chemical reaction whereby a primary or secondary alcohol is oxidized to an aldehyde or ketone using oxalyl chloride, dimethyl sulfoxide (DMSO) and an organic base, such as triethylamine. It is one of the many oxidation reactions commonly referred to as 'activated DMSO' oxidations. The reaction is known for its mild character and wide tolerance of functional groups.

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

The Henry reaction is a classic carbon–carbon bond formation reaction in organic chemistry. Discovered in 1895 by the Belgian chemist Louis Henry (1834–1913), it is the combination of a nitroalkane and an aldehyde or ketone in the presence of a base to form β-nitro alcohols. This type of reaction is also referred to as a nitroaldol reaction. It is nearly analogous to the aldol reaction that had been discovered 23 years prior that couples two carbonyl compounds to form β-hydroxy carbonyl compounds known as "aldols". The Henry reaction is a useful technique in the area of organic chemistry due to the synthetic utility of its corresponding products, as they can be easily converted to other useful synthetic intermediates. These conversions include subsequent dehydration to yield nitroalkenes, oxidation of the secondary alcohol to yield α-nitro ketones, or reduction of the nitro group to yield β-amino alcohols.

A dendralene is a discrete acyclic cross-conjugated polyene. The simplest dendralene is buta-1,3-diene (1) or [2]dendralene followed by [3]dendralene (2), [4]dendralene (3) and [5]dendralene (4) and so forth. [2]dendralene (butadiene) is the only one not cross-conjugated.

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

Norbornene or norbornylene or norcamphene is a highly strained bridged cyclic hydrocarbon. It is a white solid with a pungent sour odor. The molecule consists of a cyclohexene ring with a methylene bridge between carbons 1 and 4. The molecule carries a double bond which induces significant ring strain and significant reactivity.

<span class="mw-page-title-main">Chiral auxiliary</span> Stereogenic group placed on a molecule to encourage stereoselectivity in reactions

In stereochemistry, a chiral auxiliary is a stereogenic group or unit that is temporarily incorporated into an organic compound in order to control the stereochemical outcome of the synthesis. The chirality present in the auxiliary can bias the stereoselectivity of one or more subsequent reactions. The auxiliary can then be typically recovered for future use.

In organic chemistry, the Arndt–Eistert reaction is the conversion of a carboxylic acid to its homologue. Named for the German chemists Fritz Arndt (1885–1969) and Bernd Eistert (1902–1978), the method entails treating an acid chlorides with diazomethane. It is a popular method of producing β-amino acids from α-amino acids.

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

Itaconic acid, or methylidenesuccinic acid, is an organic compound. This dicarboxylic acid is a white solid that is soluble in water, ethanol, and acetone. Historically, itaconic acid was obtained by the distillation of citric acid, but currently it is produced by fermentation. The name itaconic acid was devised as an anagram of aconitic acid, another derivative of citric acid.

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

Furfuryl alcohol is an organic compound containing a furan substituted with a hydroxymethyl group. It is a colorless liquid, but aged samples appear amber. It possesses a faint odor of burning and a bitter taste. It is miscible with but unstable in water. It is soluble in common organic solvents.

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

Methacrylic acid, abbreviated MAA, is an organic compound with the formula CH2=C(CH3)COOH. This colorless, viscous liquid is a carboxylic acid with an acrid unpleasant odor. It is soluble in warm water and miscible with most organic solvents. Methacrylic acid is produced industrially on a large scale as a precursor to its esters, especially methyl methacrylate (MMA), and to poly(methyl methacrylate) (PMMA).

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

The Wolff rearrangement is a reaction in organic chemistry in which an α-diazocarbonyl compound is converted into a ketene by loss of dinitrogen with accompanying 1,2-rearrangement. The Wolff rearrangement yields a ketene as an intermediate product, which can undergo nucleophilic attack with weakly acidic nucleophiles such as water, alcohols, and amines, to generate carboxylic acid derivatives or undergo [2+2] cycloaddition reactions to form four-membered rings. The mechanism of the Wolff rearrangement has been the subject of debate since its first use. No single mechanism sufficiently describes the reaction, and there are often competing concerted and carbene-mediated pathways; for simplicity, only the textbook, concerted mechanism is shown below. The reaction was discovered by Ludwig Wolff in 1902. The Wolff rearrangement has great synthetic utility due to the accessibility of α-diazocarbonyl compounds, variety of reactions from the ketene intermediate, and stereochemical retention of the migrating group. However, the Wolff rearrangement has limitations due to the highly reactive nature of α-diazocarbonyl compounds, which can undergo a variety of competing reactions.

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

Diphenylketene is a chemical substance of the ketene family. Diphenylketene, like most stable disubstituted ketenes, is a red-orange oil at room temperature and pressure. Due to the successive double bonds in the ketene structure R1R2C=C=O, diphenyl ketene is a heterocumulene. The most important reaction of diphenyl ketene is the [2+2] cycloaddition at C-C, C-N, C-O, and C-S multiple bonds.

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

Sulfolene, or butadiene sulfone is a cyclic organic chemical with a sulfone functional group. It is a white, odorless, crystalline, indefinitely storable solid, which dissolves in water and many organic solvents. The compound is used as a source of butadiene.

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

Diglycolic acid is an aliphatic dicarboxylic acid, its acidity is between the one of acetic acid and oxalic acid. It is formed in the oxidation of diethylene glycol in the body and can lead to severe complications with fatal outcome.

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

Citraconic acid is an organic compound with the formula CH3C2H(CO2H)2. It is a white solid. The alkene is cis. The related trans alkene is called mesaconic acid. It is one of the pyrocitric acids formed upon the heating of citric acid. Citraconic acid can be produced, albeit inefficiently, by oxidation of xylene and methylbutanols. The acid displays the unusual property of spontaneously forming the anhydride, which, unlike maleic anhydride, is a liquid at room temperature.

<span class="mw-page-title-main">2-Ethyl-2-oxazoline</span> Chemical compound

2-Ethyl-2-oxazoline (EtOx) is an oxazoline which is used particularly as a monomer for the cationic ring-opening polymerization to poly(2-alkyloxazoline)s. This type of polymers are under investigation as readily water-soluble and biocompatible materials for biomedical applications.

<i>N</i>-Hydroxyphthalimide Chemical compound

N-Hydroxyphthalimide is the N-hydroxy derivative of phthalimide. The compound can be utilized as a catalyst for oxidation reactions, in particular for the selective oxidation with molecular oxygen under mild conditions.

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

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

α,β-Unsaturated carbonyl compound Functional group of organic compounds

α,β-Unsaturated carbonyl compounds are organic compounds with the general structure (O=CR)−Cα=Cβ-R. Such compounds include enones and enals, but also carboxylic acids and the corresponding esters and amides. In these compounds the carbonyl group is conjugated with an alkene. Unlike the case for carbonyls without a flanking alkene group, α,β-unsaturated carbonyl compounds are susceptible to attack by nucleophiles at the β-carbon. This pattern of reactivity is called vinylogous. Examples of unsaturated carbonyls are acrolein (propenal), mesityl oxide, acrylic acid, and maleic acid. Unsaturated carbonyls can be prepared in the laboratory in an aldol reaction and in the Perkin reaction.

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

Diethyl acetamidomalonate (DEAM) is a derivative of malonic acid diethyl ester. Formally, it is derived through the acetylation of ester from the unstable aminomalonic acid. DEAM serves as a starting material for racemates including both, natural and unnatural α-amino acids or hydroxycarboxylic acids. It is also usable as a precursor in pharmaceutical formulations, particularly in the cases of active ingredients like fingolimod, which is used to treat multiple sclerosis.

References

  1. Sigma-Aldrich Co., product no. {{{id}}}.
  2. J.L. Belletire, R.J. Rauh (2001), "Itaconic Anhydride", E-EROS Encyclopedia of Reagents for Organic Synthesis, doi:10.1002/047084289X.ri086, ISBN   0-471-93623-5
  3. Entry from Itaconic Anhydride from TCI Europe, retrieved on {{{Date}}}
  4. 1 2 R.L. Shriner, S.G. Ford, L.J. Roll (1931). "Itaconic Anhydride and Itaconic Acid". Organic Syntheses. 11: 70. doi:10.15227/orgsyn.011.0070.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  5. B. Kamm (2007), "Produktion von Plattformchemikalien und Synthesegas aus Biomasse", Angew. Chem. (in German), vol. 119, no. 27, pp. 5146–5149, Bibcode:2007AngCh.119.5146K, doi:10.1002/ange.200604514
  6. Birgit Kamm (2008-04-14). "Das Konzept der Bioraffinerie – Schlüssel für Ressourceneffizienz". GDCh Aktuelle Wochenschau. Retrieved 2018-10-01.
  7. Jim Lane (2015-04-30). "The DOE's 12 top biobased molecules – what became of them?". BiofuelsDigest. Retrieved 2018-10-01.
  8. S. Baup (1836), "Ueber eine neue Pyrogen-Citronensäure, und ueber Benennung der Pyrogen-Säuren überhaupt", Justus Liebigs Ann. Chem. (in German), vol. 19, no. 1, pp. 29–38, doi:10.1002/jlac.18360190107
  9. R.L. Shriner, S. G. Ford, L. J. Roll (1931). "Citraconic Anhydride and Citraconic Acid". Organic Syntheses. 11: 28. doi:10.15227/orgsyn.011.0028.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  10. US 2258947,J.H. Crowell,"Production of itaconic and citraconic anhydrides",issued 1941-10-14, assigned to National Aniline & Chemical Co.
  11. Novamont SpA (2018-07-26). "Final Report Summary – BIO-QED (Quod Erat Demonstrandum: Large scale demonstration for the bio-based bulk chemicals BDO and IA aiming at cost reduction and improved sustainability)". CORDIS. Retrieved 2018-10-01.
  12. 1 2 GB 854999,"A process for the production of itaconic anhydride",issued 1960-11-23, assigned to Chas. Pfizer & Co., Inc.
  13. US 5260456,M. Alas, M. Gubelmann, J.-M. Popa,"Process for producing itaconic anhydride",issued 1993-11-9, assigned to Rhone-Poulenc Chimie
  14. WO 9506026,A.G. Talma, A.G. Bovenkamp-Bouwman, H.P. Verlaanhooft,"Dehydration of itaconic acid",issued 1995-3-2, assigned to Akzo Nobel N.V.
  15. C. Robert, F. de Montigny, C.M. Thomas (2011), "Tandem synthesis of alternating polyesters from renewable resources", Nature Communications , vol. 2, pp. 1–6, Bibcode:2011NatCo...2..586R, doi:10.1038/ncomms1596, PMC   3247812 , PMID   22158441 {{citation}}: CS1 maint: multiple names: authors list (link)
  16. M.C. Galanti, A.V. Galanti (1982), "Kinetic study of the isomerization of itaconic anhydride to citraconic anhydride", J. Org. Chem. , vol. 47, no. 8, pp. 1572–1574, doi:10.1021/jo00347a041
  17. W. Petri (1881), "Beiträge zur Kenntnis der Itaconsäure, Mesaconsäure und Citraconsäure", Ber. Dtsch. Chem. Ges. (in German), vol. 14, no. 2, pp. 1634–1637, doi:10.1002/cber.18810140213
  18. C. Wang, X. Wang, Z. Wie, X. Zeng (2018), "Synthesis and characterization of poly(p-phenyleneitaconamide)", Polym. Mater. Sci. Eng., vol. 34, no. 6, pp. 9–15, doi:10.16865/j.cnki.1000-7555.2018.06.002 {{citation}}: CS1 maint: multiple names: authors list (link)
  19. J. Nokami, T. Tamaoka, H. Ogawa, S. Wakabayashi (1986), "Facile synthesis of 2-methylene-4-butyrolactones", Chem. Lett. , vol. 15, no. 4, pp. 541–544, doi:10.1246/cl.1986.541 {{citation}}: CS1 maint: multiple names: authors list (link)
  20. O. Diels, K. Alder (1928), "Synthesen in der hydroaromatischen Reihe", Justus Liebigs Ann. Chem. (in German), vol. 460, no. 1, pp. 98–122, doi:10.1002/jlac.19284600106
  21. A.D. Pehere, S. Xu, S.K. Thompson, M.A. Hillmyer, T.R. Hoye (2016), "Diels-Alder reactions of furans with itaconic anhydride: Overcoming unfavorable thermodynamics", Org. Lett. , vol. 18, no. 11, pp. 2584–2587, doi:10.1021/acs.orglett.6b00929, PMC   5136459 , PMID   27214494 {{citation}}: CS1 maint: multiple names: authors list (link)
  22. K. Kameo, K. Ogawa, K. Takeshita, S. Nakaike, K. Tomisawa, K. Sato (1988), "Studies on antirheumatic agents: 3-benzoylpropionic acid derivatives", Chem. Pharm. Bull. , vol. 36, no. 6, pp. 2050–2060, doi: 10.1248/cpb.36.2050 , PMID   3240440 {{citation}}: CS1 maint: multiple names: authors list (link)
  23. H. Takeda, T.Tachinami, S. Hosokawa, M. Aburatani, K. Inoguchi, K. Achiwa (1991), "Efficient Preparation of Optically Active (S)-(-)-3-Methyl-γ-butyrolactone by Catalytic Asymmetric Hydrogenation Using Chiral N-Substituted Pyrrolidinebisphosphine Rhodium Complexes", Chem. Pharm. Bull. , vol. 39, no. 10, pp. 2706–2708, doi: 10.1248/cpb.39.2706 {{citation}}: CS1 maint: multiple names: authors list (link)
  24. M. Akiyama, K. Shimizu, S. Aiba, F. Banba (1980), "Synthesis of N-Hydroxymaleimide and N-Hydroxyitaconimide and their related derivatives", J. Chem. Soc. Perkin I , pp. 2122–2125, doi:10.1039/P19800002122 {{citation}}: CS1 maint: multiple names: authors list (link)
  25. A.M. Medway, J. Sperry (2014), "Heterocycle construction using the biomass-derived building block itaconic acid", Green Chem. , vol. 16, no. 4, pp. 2084–2101, doi:10.1039/c4gc00014e
  26. D. Stawski, S. Polowinski (2005), "Polymerization of itaconic acid", Polimery, vol. 50, no. 2, pp. 118–122
  27. F.H. Isikgor, C.R. Becer (2015), "Lignocellulosic biomass: a sustainable platform for the production of bio-based chemicals and polymers", Polym. Chem. , vol. 6, no. 25, pp. 4497–4559, arXiv: 1602.01684 , doi:10.1039/c5py00263j, S2CID   51812213
  28. T. Okuda, K. Ishimoto, H. Ohara, S. Kobayashi (2012), "Renewable biobased polymeric materials: Facile synthesis of itaconic anhydride-based copolymers with poly(L-lactic acid) grafts", Macromolecules , vol. 45, no. 10, pp. 4166–4174, Bibcode:2012MaMol..45.4166O, doi:10.1021/ma300387j {{citation}}: CS1 maint: multiple names: authors list (link)
  29. T. Otsu, J.-Z. Yang (1991), "Radical polymerization of itaconic anhydride and reactions of the resulting polymers with amines and alcohols", Polymer Int., vol. 25, no. 4, pp. 245–251, doi:10.1002/pi.4990250408
  30. S. Shang, S.J. Huang, R.A. Weiss (2011), "Comb-like ionomers from sustainable resources: Copolymers of itaconic anhydride-co-stearyl methacrylate", Polymer , vol. 52, no. 13, pp. 2764–2771, doi:10.1016/j.polymer.2011.04.025 {{citation}}: CS1 maint: multiple names: authors list (link)
  31. Y. Bai, J.H. Clark, T.J. Farmer, I.D.V. Ingram, M. North (2017), "Wholly biomass derivable sustainable polymers by ring-opening metathesis polymerization of monomers obtained from furfuryl alcohol and itaconic anhydride", Polymer Chem. , vol. 8, no. 20, pp. 3074–3081, doi:10.1039/C7PY00486A {{citation}}: CS1 maint: multiple names: authors list (link)