Iminophosphoranes (also known as phosphine imides, phosphinimide, phosphinimines, λ5-phosphazenes, acyclic phosphazenes) are a class of organophosphorus compounds and an acyclic subclass of phosphazenes with the general formula R3P=NR’.[1] First reported by Staudinger and Meyer in 1919[2], these isoelectronic analogues of phosphine oxides and phosphonium ylides (also known as Wittig reagents)[3][4] are most commonly synthesized via the Staudinger reaction or Kirsanov reaction, though alternative synthetic routes have also been developed. [5][6][7]
The P=N bond is best described as a highly polarized single bond consistent with ylidic resonance structure R3P+=N-R', and the steric and electronic character may be tuned by varying the substituents on either the phosphorus or nitrogen.[8][9] These properties allowing for interesting reactivity of iminophosphoranes as a Brønsted (super)base[10] or coordinating ligand at the nitrogen[8], or for [2+2] cycloadditions with the P=N bond.[11][12] As such, iminophosphoranes have found diverse applications as ligands for homogeneous catalysis (i.e. cross coupling, polymerization, etc.)[8], superbasic or bifunctional organocatalysts[13], probes for chemical biology [14][15], and more.
Synthesis
Staudinger Reaction
The earliest synthesis of an iminophosphorane was reported by Staudinger and Meyer in 1919[2] from the reaction of an azide and a phosphine with nitrogen extrusion. This reaction proceeds via nucleophilic attack of the phosphine on the azide through a cis-transition state to form a phosphazide intermediate which undergoes a four-membered ring closure to expel N2 (Figure 1). [16] Over 100 years later, the Staudinger reaction remains one of the most general and widely used methods. [5][6] A variation of the Staudinger reaction, Staudinger ligation, is a highly selective bioorthogonal reaction that is prominent in chemical biology in labelling or modifying cellular environments, proteins, DNA, etc. [14][15]

Kirsanov Reaction

A second notable method to synthesize iminophosphoranes is the Kirsanov reaction first reported in 1950 in which P-halogenated iminophosphoranes are accessed from phosphorus pentachloride and amine starting materials (Figure 2).[17] In 1959, Horner and Oediger published a modified Kirsanov reaction in which halogenation of a tertiary phosphine produces a phosphonium salt, which is treated in situ with a primary amine to yield the iminophosphorane (Figure 3).[18]

Other Synthetic Routes
Staudinger and Kirsanov reactions are considered the two most common synthetic routes, but many other potential synthetic strategies to access iminophosphoranes have also been developed to address potential limitations. For instance, the Staudinger reaction involves high energy and potentially explosive azides, and there are some instances when azide may not be readily available. The phosphorus pentachloride and bromine reagents involved in the Kirsanov and modified Kirsanov reaction are also toxic.[7] Examples of alternative routes include the synthesis of N-acyliminophosphoranes via iron-catalyzed imidzation of phosphines with N-acyloxyamides (Figure 4, bottom)[19] or by an iron-photocatalyzed nitrene transfer reaction with dioxazolones (Figure 4, top).[20] Another example is the electrochemical nickel-catalyzed synthesis of N-cyanoiminophosphoranes from treating phosphines with bis(trimethylsilyl)carbodiimide (Figure 5).[7]


Structure
The precise nature of the iminophosphorane P=N bond (and more generally, P=E bonds, where E = C, N, O) has historically been the subject of much discussion. [4][21] Initially, from around 1950 to 1970, the P=N bond was thought to have significant contribution from a π-type interaction between low lying d orbitals on phosphorus and the p orbitals on nitrogen due to a shortened P=N bond length relative to a single P-N bond indicated by spectroscopic studies, dipole measurements, and X-ray analyses. [4][5][22] However, further computations in the 1980s would clearly demonstrate that d-orbitals are not significantly involved in such bonds with key contributions from Magnusson, Reed, Weinhold, and Schleyer. [4] This model was also inconsistent with observed reactivity (i.e. cleavage of the P-N bond by metal organyls in polar solvents or susceptibility to hydrolysis, which would be difficult with a true P=N double bond).[23]

Over time, the accepted bond model was revised to be that there is no true P-N π bond and the interaction is instead best described as a strongly polarized P+-N- bond.[21][9] The shortened strengthened bond and stabilization of the phosphorus can be justified by negative hyperconjugation, in which electron density on the nitrogen p-orbital delocalizes into the σ* (P-C) orbital. [9][4] This was first proposed and supported by theoretical studies from Sun, Liu, and coworkers.[24][25] In 2004, Stalke and coworkers provided experiment evidence of this model with experimental charge-density studies and topological analysis of alkali-metal coordinated iminophosphoranes showing a polar P+-N- "augmented by electrostatic contributions." [23] Therefore, although the iminophosphorane P=N bond is typically drawn as a double bond, it is highly polarized and is most accurately described as a hybrid of resonance contributors between the ylene and ylidic forms (Figure 6).[8][9]
Reactivity
Brønsted Basicity
Iminophosphoranes are strong Brønsted bases with high affinity for protons. In the proton adduct (R3P=NR’H+), the nitrogen now favors a pyramidal sp3 hybridization and electron density still resides primarily on the nitrogen with partial positive charge on the phosphorus. Therefore, this adduct is best described as an aminophosphonium ion instead of an iminium ion. [3]
Certain iminophosphorane derivatives are also known to behave as nonionic superbases. Generally, the study and design of strong neutral (super)bases is of great interest in organic synthesis applications due to their increased solubility in common solvents, allowance for milder conditions, and greater stability relative to inorganic ionic bases.[26] Schwesinger first reported the outstanding Brønsted basicity of amino-substituted iminophosphorane derivatives with high stability and kinetic activity in 1985, which would lay the foundation for future advances in the realm of iminophosphorane superbases.[27][10] Computational studies by Maksić and coworkers in 2004 would later developed a theoretical framework for estimating pKBH+ values of iminophosphoranes and rationalize their basicity from effective resonance stabilization of the conjugate acid. [26] This family of iminophosphorane (super)bases can be classified by Schwesinger's nomenclature (i.e. Pn, where n denotes the number of P=N units). The basicity can be increased by a number of factors including higher n, electron donating substituents, and branching. [10]
Coordination to Metal Centers
As ligands, iminophosphoranes can coordinate to other centers (i.e. transition metals, main group elements) via lone pair(s) on the nitrogen.[8] Iminophosphorane are predominantly hard σ- donors, with possible π-donor capabilities due to the two lone pairs on nitrogen.[9] Simple iminophosphoranes are monodentate neutral L-type donors,[8] but additional donor sites can be incorporated in the iminophosphorane substituents to create polydentate ligands which allow for increased stability by the chelating effect.[8] These polydentate mixed ligands are also highly useful for enabling select catalysis, as the donor sites and linkers in these polydentate mixed ligands can be precisely tuned to vary reactivity.[8]
In electron rich centers (i.e. late transition metals), iminophosphoranes tend to coordinate relatively weakly and are easily displaced by other ligands, which was first demonstrated by Fukui and coworkers in 1975 by the rapid displacement of iminophosphorane ligands by triphenylphosphine, triphenylphosphite, and 2,2'-bipyridine under mild conditions at Pd(II) complex centers. [28][8] However, iminophosphoranes, particularly polydentate mixed ligands, can bind most strongly to electron deficient centers (i.e. rare earth elements, early transition metals, and highly oxidized centers).[9]
[2+2] Addition to P=N Bond
In addition to the nitrogen-centered reactivity, the phosphorus has also been shown to play a role in some reactions enabled by the vicinal ambiphilic character of the P=N bond. [29] Despite the absence of the d-orbitals in the P=N bonds (compared to an analogous M=N transition metal imide), iminophosphoranes of the class X3P=NR (X = Cl, pyrrolyl; R = alkyl, aryl) were shown by Geib and coworkers to be engage in catalytic double-bond metathesis in a Chauvin-type pathway[30], proceeding through a phosphetidine intermediate formed from [2+2] addition of carbodiimides to the iminophosphorane. [11]
Another prominent representative example is the aza-Wittig reaction, a widely used method of making C=N double bonds in which an aldehyde or ketone reacts with a stoichiometric amount of iminophosphorane to yield an imine.[12] The exact mechanism has historically been difficult to determine owing to the instability of the intermediates, but many theoretical studies including those by Koketsu and coworkers in 1997,[31] Lu and coworkers in 1999[32], and Palacios and coworkers in 2006,[33] support a two-step mechanism proceeding via a four-membered ring intermediate from the first tandem [2+2] cycloaddition-cycloreversion step in a thermally allowed mechanism [12], followed by the second [2+2] cycloreversion step (Figure 7).

Selected applications
Ligands for Homogeneous Catalysis

One notable example of the use of iminophosphorane as ligands in catalyst complexes is the highly active yttrium phosphasalen initiator for rac-Lactide polymerization with excellent polymerization control, stereoselectivity, and rate of reaction reported by Auffrant and Williams in 2012. This effective initiator (Figure 8) is enabled by the use of a “phosphasalen” ligand, an iminophosphorane analog of the well-known “salen” ligand. The authors suggest that the strong σ and π donation to the electron deficient yttrium by the iminophosphorane (with greater electron donation relative to the "salen" ligand) allowed for higher rates of reaction by acclerating the rate of the lactide inversion step . The phosphasalene structure was also advantageous in that it could be modified to allow for high iso-selectivity or hetero-selectivity, eliminating the need for chiral ancillary ligands or complexes for stereocontrol.[34]
There is are many additional instances in literature of employing iminophosphoranes as ligands for various reactions including cross-coupling reactions (i.e. Sonogashira, Suzuki-Miyaura, Kumada-Corriu, and Negishi couplings), hydrogenation of alkenes, cyclopropanation, and other polymerization reactions.[8][7] One example of a Kumada coupling catalyzed by a nickel complex with an iminophosphorine P, N, N-amido pincer ligand is show below in Figure 9.

Bifunctional Iminophosphorane Catalysts
Bi- or multi-functional Brønsted base catalysts, with a separate Brønsted base and H-bond donor moiety, are widely used for mediating enantioselective polar addition reactions with hundreds of reports to date.[13] Such catalysts commonly employ teritary amine as the Brønsted base, which imposes challenges in catalytic design due to the high acidity of the pronucleophile and moderate nucleophilicity of the conjugate base. [13] In recent years, it was found that these limitations could be addressed by incorporating a superbasic iminophosphorane moiety, allowing for controlled tunability for good enantiocontrol and sufficient Brønsted basicity for good pronucleophile activation. [13][35]
One of the earliest instances of such bifunctional iminophosphorane (BIMP) catalysts was reported by Dixon and coworkers in 2013, when a BIMP catalyst with a triaryliminophosphorane Brønsted base site was employed as a catalyst to achieve first general enantioselective nitro-Mannich reaction of nitromethane with N-DPP-protected ketimines (Figure 10), yielding the desired β-nitroamine in 3 hours whereas other bifunctional Brønsted tertiary amine-based catalysts did not yield product. The BIMP catalyst was highly tunable, with the substituted aryl groups on the iminophosphorane phosphorus being easily synthetically modified to control the pKBH+ and electronics of the superbase moiety. [35]

In recent years, BIMP catalysts have been employed across a number of diverse and valuable reactions, with notable examples including the first enantioselective sulfa-Michael addition of alkyl thiols to unactivated α-substituted acrylate esters by Dixon and coworkers in 2015[36], ring-opening polymerization (ROP) of cyclic esters l-lactide (LA), δ-valerolactone (VL), and ε-caprolactone (CL) by Dixon and coworkers in 2014,[37] and in total synthesis of complex alkaloid natural product (-)-Himalensine A. [38]
References
- ↑ Chakraborty, Amit; Ahmed, Naushad; Chandrasekhar, Vadapalli (2022-09-05), Higham, Lee J; Allen, David W; Tebby, John C (eds.), "Phosphazenes", Organophosphorus Chemistry (1 ed.), The Royal Society of Chemistry, pp. 355–397, doi:10.1039/9781839166198-00355, ISBN 978-1-83916-522-1, retrieved 2025-11-21
- 1 2 Staudinger, H.; Meyer, Jules (1919). "Über neue organische Phosphorverbindungen III. Phosphinmethylenderivate und Phosphinimine". Helvetica Chimica Acta. 2 (1): 635–646. Bibcode:1919HChAc...2..635S. doi:10.1002/hlca.19190020164. ISSN 1522-2675.
- 1 2 Starzewski, K. A. Ostoja; Tom Dieck, H. (1979-12-01). "Electronic structure and reactivity. 8. Iminophosphoranes - simple ylide analogs? An investigation of quantitative and phenomenological differences". Inorganic Chemistry. 18 (12): 3307–3316. doi:10.1021/ic50202a005. ISSN 0020-1669.
- 1 2 3 4 5 Gilheany, Declan G. (1994-07-01). "No d Orbitals but Walsh Diagrams and Maybe Banana Bonds: Chemical Bonding in Phosphines, Phosphine Oxides, and Phosphonium Ylides". Chemical Reviews. 94 (5): 1339–1374. Bibcode:1994ChRv...94.1339G. doi:10.1021/cr00029a008. ISSN 0009-2665. PMID 27704785.
- 1 2 3 Wamhoff, Heinrich; Richardt, Gabriele; Stölben, Stephan (1995-01-01), Iminophosphoranes: Versatile Tools in Heterocyclic Synthesis, Advances in Heterocyclic Chemistry, vol. 64, Academic Press, pp. 159–249, doi:10.1016/s0065-2725(08)60172-5, ISBN 978-0-12-020764-0, retrieved 2025-11-20
- 1 2 Steiner, Alexander; Zacchini, Stefano; Richards, Philip I (2002-04-15). "From neutral iminophosphoranes to multianionic phosphazenates. The coordination chemistry of imino–aza-P(V) ligands". Coordination Chemistry Reviews. 227 (2): 193–216. doi:10.1016/S0010-8545(02)00012-7. ISSN 0010-8545.
- 1 2 3 4 Mdluli, Velabo; Lehnherr, Dan; Lam, Yu-hong; Chaudhry, Mohammad T.; Newman, Justin A.; DaSilva, Jimmy O.; Regalado, Erik L. (2024). "Electrosynthesis of iminophosphoranes and applications in nickel catalysis". Chemical Science. 15 (16): 5980–5992. doi:10.1039/D3SC05357A. ISSN 2041-6520. PMC 11041257. PMID 38665537.
- 1 2 3 4 5 6 7 8 9 10 García-Álvarez, Joaquín; García-Garrido, Sergio E.; Cadierno, Victorio (2014-02-01). "Iminophosphorane–phosphines: Versatile ligands for homogeneous catalysis". Journal of Organometallic Chemistry. 50th anniversary special issue. 751: 792–808. doi:10.1016/j.jorganchem.2013.07.009. hdl:10651/25304. ISSN 0022-328X.
- 1 2 3 4 5 6 Tannoux, Thibault; Auffrant, Audrey (2023-01-01). "Complexes featuring tridentate iminophosphorane ligands: Synthesis, reactivity, and catalysis". Coordination Chemistry Reviews. 474 214845. doi:10.1016/j.ccr.2022.214845. ISSN 0010-8545.
- 1 2 3 Ishikawa, Tsutomu (2009). Superbases for organic synthesis: guanidines, amidines and phosphazenes and related organocatalysts. Chichester, UK: John Wiley & Sons. ISBN 978-0-470-51800-7.
- 1 2 Bell, Stephen A.; Meyer, Tara Y.; Geib, Steven J. (2002-08-17). "Catalytic Double-Bond Metathesis without the Transition Metal". Journal of the American Chemical Society. 124 (36): 10698–10705. doi:10.1021/ja020494v. ISSN 0002-7863.
- 1 2 3 Palacios, Francisco; Alonso, Concepción; Aparicio, Domitila; Rubiales, Gloria; de los Santos, Jesús M. (January 2007). "The aza-Wittig reaction: an efficient tool for the construction of carbon–nitrogen double bonds". Tetrahedron. 63 (3): 523–575. doi:10.1016/j.tet.2006.09.048. ISSN 0040-4020.
- 1 2 3 4 Formica, Michele; Rozsar, Daniel; Su, Guanglong; Farley, Alistair J. M.; Dixon, Darren J. (2020-10-20). "Bifunctional Iminophosphorane Superbase Catalysis: Applications in Organic Synthesis". Accounts of Chemical Research. 53 (10): 2235–2247. doi:10.1021/acs.accounts.0c00369. ISSN 0001-4842.
- 1 2 Saxon, Eliana; Bertozzi, Carolyn R. (2000-03-17). "Cell Surface Engineering by a Modified Staudinger Reaction". Science. 287 (5460): 2007–2010. doi:10.1126/science.287.5460.2007.
- 1 2 Lin, Fiona L.; Hoyt, Helen M.; van Halbeek, Herman; Bergman, Robert G.; Bertozzi, Carolyn R. (2005-03-01). "Mechanistic Investigation of the Staudinger Ligation". Journal of the American Chemical Society. 127 (8): 2686–2695. doi:10.1021/ja044461m. ISSN 0002-7863.
- ↑ Tian, Wei Quan; Wang, Yan Alexander (2004-06-01). "Mechanisms of Staudinger Reactions within Density Functional Theory". The Journal of Organic Chemistry. 69 (13): 4299–4308. doi:10.1021/jo049702n. ISSN 0022-3263.
- ↑ Keat, R.; Miller, M. C.; Shaw, R. A. (1967-01-01). "Phosphorus–nitrogen compounds. Part XXV. The reactions of aminocyclotriphosphazatrienes with dihalogenotriphenyl- and trihalogenodiphenyl-phosphoranes. Phosphazen-1′-ylcyclotriphosphazatrienes". Journal of the Chemical Society A: Inorganic, Physical, Theoretical (0): 1404–1408. doi:10.1039/J19670001404. ISSN 0022-4944.
- ↑ Horner, Leopold; Oediger, Hermann (1959). "Phosphororganische Verbindungen, XVIII Phosphinimino-Verbindungen aus Phosphindihalogeniden und primären Aminen". Justus Liebigs Annalen der Chemie (in German). 627 (1): 142–162. doi:10.1002/jlac.19596270115. ISSN 1099-0690.
- ↑ Lin, Sen; Lin, Bo; Zhang, Zongtao; Chen, Jianhui; Luo, Yanshu; Xia, Yuanzhi (2022-05-06). "Construction of N-Acyliminophosphoranes via Iron(II)-Catalyzed Imidization of Phosphines with N-Acyloxyamides". Organic Letters. 24 (17): 3302–3306. doi:10.1021/acs.orglett.2c01238. ISSN 1523-7060.
- ↑ Tang, Jing-Jing; Yu, Xiaoqiang; Wang, Yi; Yamamoto, Yoshinori; Bao, Ming (2021). "Interweaving Visible-Light and Iron Catalysis for Nitrene Formation and Transformation with Dioxazolones". Angewandte Chemie International Edition. 60 (30): 16426–16435. doi:10.1002/anie.202016234. ISSN 1521-3773.
- 1 2 Kocher, Nikolaus; Leusser, Dirk; Murso, Alexander; Stalke, Dietmar (2004). "Metal Coordination to the Formal PN Bond of an Iminophosphorane and Charge-Density Evidence against Hypervalent Phosphorus(V)". Chemistry – A European Journal. 10 (15): 3622–3631. Bibcode:2004ChEuJ..10.3622K. doi:10.1002/chem.200400163. ISSN 1521-3765. PMID 15281145.
- ↑ Horner, Leopold; Oediger, Hermann (1959). "Phosphororganische Verbindungen, XVIII Phosphinimino-Verbindungen aus Phosphindihalogeniden und primären Aminen". Justus Liebigs Annalen der Chemie (in German). 627 (1): 142–162. doi:10.1002/jlac.19596270115. ISSN 1099-0690.
- 1 2 Kocher, Nikolaus; Leusser, Dirk; Murso, Alexander; Stalke, Dietmar (2004-07-22). "Metal Coordination to the Formal PN Bond of an Iminophosphorane and Charge-Density Evidence against Hypervalent Phosphorus(<scp>V</scp>)". Chemistry – A European Journal. 10 (15): 3622–3631. Bibcode:2004ChEuJ..10.3622K. doi:10.1002/chem.200400163. ISSN 0947-6539. PMID 15281145.
- ↑ Lu, Wen Cai; Liu, Cheng Bu; Sun, Chia Chung (1999-02-01). "Theoretical Study of the H3PNH + H2CO Reaction Mechanism via Five Reaction Channels". The Journal of Physical Chemistry A. 103 (8): 1078–1083. Bibcode:1999JPCA..103.1078L. doi:10.1021/jp982124s. ISSN 1089-5639.
- ↑ Lu, Wen Cai; Sun, Chia Chung; Zang, Qing Jun; Liu, Cheng Bu (October 1999). "Theoretical study of the aza-Wittig reaction X3P=NH+O=CHCOOH→X3P=O+HN=CHCOOH for X=Cl, H and CH3". Chemical Physics Letters. 311 (6): 491–498. doi:10.1016/s0009-2614(99)00887-8. ISSN 0009-2614.
- 1 2 Kovačević, Borislav; Barić, Danijela; Maksić, Zvonimir B. (2004-02-03). "Basicity of exceedingly strong non-ionic organic bases in acetonitrile —Verkade's superbase and some related phosphazenes". New Journal of Chemistry. 28 (2): 284–288. doi:10.1039/B309506A. ISSN 1369-9261.
- ↑ Schwesinger, Reinhard (1985-09-30). "Extremely Strong, Non-ionic Bases: Syntheses and Applications". CHIMIA. 39 (9): 269. doi:10.2533/chimia.1985.269. ISSN 2673-2424.
- ↑ Fukui, Mitsuyo; Itoh, Kenji; Ishii, Yoshio (1975-07-01). "Iminophosphorane Complexes of Palladium(II)". Bulletin of the Chemical Society of Japan. 48 (7): 2044–2046. doi:10.1246/bcsj.48.2044. ISSN 0009-2673.
- ↑ Lin, Yi-Chun; Gilhula, James C.; Radosevich, Alexander T. (2018-05-09). "Nontrigonal constraint enhances 1,2-addition reactivity of phosphazenes". Chemical Science. 9 (18): 4338–4347. doi:10.1039/C8SC00929E. ISSN 2041-6539. PMC 5944378. PMID 29780566.
- ↑ Jean‐Louis Hérisson, Par; Chauvin, Yves (1971-02-09). "Catalyse de transformation des oléfines par les complexes du tungstène. II. Télomérisation des oléfines cycliques en présence d'oléfines acycliques". Die Makromolekulare Chemie. 141 (1): 161–176. doi:10.1002/macp.1971.021410112. ISSN 0025-116X.
- ↑ Koketsu, Jugo; Ninomiya, Yoshihiko; Suzuki, Yoshizo; Koga, Nobuaki (1997-02-12). "Theoretical Study on the Structures of Iminopnictoranes and Their Reactions with Formaldehyde". Inorganic Chemistry. 36 (4): 694–702. doi:10.1021/ic951220u. ISSN 0020-1669.
- ↑ Lu, Wen Cai; Sun, Chia Chung; Zang, Qing Jun; Liu, Cheng Bu (October 1999). "Theoretical study of the aza-Wittig reaction X3P=NH+O=CHCOOH→X3P=O+HN=CHCOOH for X=Cl, H and CH3". Chemical Physics Letters. 311 (6): 491–498. doi:10.1016/s0009-2614(99)00887-8. ISSN 0009-2614.
- ↑ Cossío, Fernando P.; Alonso, Concepción; Lecea, Begoña; Ayerbe, Mirari; Rubiales, Gloria; Palacios, Francisco (2006-03-02). "Mechanism and Stereoselectivity of the Aza-Wittig Reaction between Phosphazenes and Aldehydes". The Journal of Organic Chemistry. 71 (7): 2839–2847. doi:10.1021/jo0525884. ISSN 0022-3263.
- ↑ Bakewell, Clare; Cao, Thi-Phuong-Anh; Long, Nicholas; Le Goff, Xavier F.; Auffrant, Audrey; Williams, Charlotte K. (2012-12-26). "Yttrium Phosphasalen Initiators for rac -Lactide Polymerization: Excellent Rates and High Iso-Selectivities". Journal of the American Chemical Society. 134 (51): 20577–20580. doi:10.1021/ja310003v. ISSN 0002-7863.
- 1 2 Núñez, Marta G.; Farley, Alistair J. M.; Dixon, Darren J. (2013-10-24). "Bifunctional Iminophosphorane Organocatalysts for Enantioselective Synthesis: Application to the Ketimine Nitro-Mannich Reaction". Journal of the American Chemical Society. 135 (44): 16348–16351. doi:10.1021/ja409121s. ISSN 0002-7863.
- ↑ Farley, Alistair J. M.; Sandford, Christopher; Dixon, Darren J. (2015-12-30). "Bifunctional Iminophosphorane Catalyzed Enantioselective Sulfa-Michael Addition to Unactivated α-Substituted Acrylate Esters". Journal of the American Chemical Society. 137 (51): 15992–15995. doi:10.1021/jacs.5b10226. ISSN 0002-7863.
- ↑ Goldys, Anna M.; Dixon, Darren J. (2014-02-25). "Organocatalytic Ring-Opening Polymerization of Cyclic Esters Mediated by Highly Active Bifunctional Iminophosphorane Catalysts". Macromolecules. 47 (4): 1277–1284. doi:10.1021/ma402258y. ISSN 0024-9297.
- ↑ Shi, Heyao; Michaelides, Iacovos N.; Darses, Benjamin; Jakubec, Pavol; Nguyen, Quynh Nhu N.; Paton, Robert S.; Dixon, Darren J. (2017-12-13). "Total Synthesis of (−)-Himalensine A". Journal of the American Chemical Society. 139 (49): 17755–17758. doi:10.1021/jacs.7b10956. ISSN 0002-7863.