Michaelis-Arbuzov Synthesis diagrams

Reaction of alkyl halides and trialkyl phosphites to phosphonates.
View examples on Reaxys(opens in new window)

Synthesis: Carbon with Two Heteroatoms, Each Attached by a Single Bond

Peter D. Kennewell, ... Nicholas J. Westwood, in Comprehensive Organic Functional Group Transformations, 1995

4.20.2.1.7 Enolisation of α-carbonylphosphorus compounds

The standard Michaelis–Arbuzov reaction results from the alkylation of trialkyl phosphites to give dialkyl alkylphosphonates (Equation (24)). With α-halo ketones, the reaction proceeds differently to give a product which was shown by Perkow to be an enol phosphate (256) (Equation (25)), whilst with α-halo acid halides both reactions take place and the product is an α-phosphorylated enol phosphate (257) (Equation (26) and Table 32) <61CRV607>.

(24)
(25)
(26)

Table 32. The Perkow reaction of α-halo acyl halides.

X1X2X3X4Yield (%)Ref.
MeHHCl8657CA(51)10366g, 60CA(54)18356h
EtHHCl8657ZOB2161, 60CA(54)18356h
EtHMeBr7957ZOB2161, 60CA(54)18356h
EtHClCl8659CA(53)10040f
MeClClCl7857ZOB2161, 60CA(54)18356h, 90JOC5982
EtClClCl7959CA(53)10040f, 90JOC5982
MeMeMeBr9257ZOB2161
EtMeMeBr9057ZOB2161
PriHHClquantitative60CA(54)18356h
MeMeMeClquantitative60CA(54)18356h
CH2CH2C1HHClquantitative60CA(54)18356h
MeHClClquantitative60CA(54)18356h
EtMeMeClquantitative60CA(54)18356h
CH2CH2C1ClClClquantitative60CA(54)18356h
CH2CH2C1HClClquantitative60CA(54)18356h

Baldwin and Swallow showed that triethyl phosphite reacts with diphenylacetyl chloride in refluxing diethyl ether to give the stable, isolable enol, diethyl 1-hydroxy-2,2-diphenyl-ethenylphosphonate (258) <70JOC3583>.

The reaction in Equation (26) presumably goes via an acylphosphonate (259), and preformed derivatives (259; X = H) react with acylating agents under acidic conditions at 130–140 °C to give, for example, the enol acetate (260) (Table 33) <73CA(78)85856s>. Zinc and phosgene convert the 2-bromo-2-methylpropanoylphosphonate (261) into the unstable chloroformate (262) (Scheme 25) <89FP2610926, 90JOC5982>.

Table 33. O-Acylation of acyl phosphonates.

R1R2R3X1X2Yield (%)Ref.
EtHHMeCOCl3673CA85856s, 76CA122034u, 84ZOB1324
EtHH(EtO)2PCl4284ZOB1324
EtHHTMSCl4384ZOB1324
MeHHTMSCl5084ZOB1324
EtHHEtCOOCOEt8376CA122034u
EtHMeMeCOOCOMe7776CA122034u
EtHMeEtCOOCOEt7276CA122034u

Scheme 25.

The Perkow reaction also proceeds with other phosphorus compounds, for example triphenylphosphine (Equation (27)) <74BSF2263>, methyl diphenylphosphinate (Equation (28)) <83CB3141, 83ZN(B)726> and ethyl tetraethyldiamidophosphite (Equation (29)) <90ZOB1940>.

(27)
(28)
(29)

Tyryshkin and co-workers have shown that dialkyl phosphites react with acetic anhydride in refluxing acetonitrile to give mixtures of the acetyl phosphonates (263) and 1-acetoxyvinylphosphonates (264) and that the use of metal catalysts, particularly iron(II), iron(III) and cobalt(II), greatly increases the yield of (264) (Table 34) <92HAC127>.

Table 34. Reactions of dialkyl phosphites with acetic anhydride in the presence of metal catalysts to give acetyl phosphonates (RO)2P(O)COCH3(263)-ref and their enol acetates (264).

RCatalystTime (h)Yield (%)
Empty CellEmpty CellEmpty Cell263264
MeCoCl241075
FeCl321088
FeCl3·6H2O2560
EtCoCl241080
FeCl221085
FeCl3·6H2O1.8560
PrCoCl241580
FeCl321090
FeCl3·6H2O21060
PriCoCl241085
FeCl221090
FeCl3·6H2O21080
BuCoCl231575
FeCl221085
FeCl3·6H2O1.5570

With unsaturated acid chlorides, the reaction can take a different course. Thus, trans-but-2-enoyl chloride reacts with trimethyl phosphite to give, inter alia, (265) in about 27% yield (Scheme 26) <81JCS(P1)1363>. p-Chlorocinnamoyl chloride, however, reacts with trimethyl phosphite in the absence of solvent to give the ‘trimer’ (266) in 37% yield <80HCA402>.

Scheme 26.

Read full chapter
URL: https://www.sciencedirect.com/science/article/pii/B0080447058002715

ABIKO–MASAMUNE Asymmetric Aldol Reaction to AUWERS–INHOFFEN Dienone-Phenol Rearrangement

A. Hassner, I. Namboothiri, in Organic Syntheses Based on Name Reactions (Third Edition), 2012

ARBUZOV Phosphonate Synthesis

Also known as Michaelis–Arbuzov. Synthesis of phosphonates 8 by heating of alkyl halides 5 with trialkyl phosphites. Ni catalyzed conversion of aryl halides 3 to aryl phosphonates 4 by reaction with phosphites 1, via phosphite-Ni complex 2.

Diethyl phenylphosphonate (4).3 To 2 (2 mg) and PhI 3 (1 g, 4.9 mmol) was slowly added 1 (0.93 g, 5.64 mmol) at 160 °C. The solution (red upon each addition of 1) faded to yellow and Et–I was distilled. Vacuum distillation afforded 4 (94%), bp 94–101 °C, 0.1 mm.

Dimethyl (N-benzyl-N-(1-phenylethyl)carbamoyl)methylphosphonate (8).10 Phosphite 6 (0.725 g, 0.69 mL, 5.85 mmol) and bromide 5 (0.65 g, 1.95 mmol) were heated at 110 °C for 5 h. Volatile impurities were removed in a Kugelrohr in vacuum and the residue was purified by flash chromatography (silica gel, EA/hexanes/MeOH) to afford 8 (98%).

1Michaelis AChem Ber1898311048
2Arbuzov AEJ Russ Phys Chem Soc190638687
3Balthazor TMJ Org Chem1980455425
4Lebeau LTet Lett1995365183
5Hudson HRARKIVOC2004ix19
6Klausmeyer KKInorg Chem Comm20069418
7Reddy CSARKIVOC2006xvi128
8Matveeva EVTet Lett2006477645
9Pakulski ZTet Lett2007488482
10*Ordonez MTet Asymm2007182427
11Michalski JChem Eur J2009151747
12Mohanakrishnan AKOrg Lett2011131270
Read full chapter
URL: https://www.sciencedirect.com/science/article/pii/B9780080966304001276

PRINCIPLES OF STABILIZATION

George Wypych, in PVC Degradation and Stabilization (Third Edition), 2015

11.3.5.2.4 Research findings

The so-called Michaelis-Arbuzov reaction is frequently quoted in literature regarding phosphites.56,295,296 The reaction was first described in 1898.300–302 It outlines transitional steps of reaction of phosphite with either HCl or ZnCl2 to obtain phosphonate and alkyl chloride:

Internal rearrangement from transitional state to final products usually requires heat and in many cases also catalysis.

Using spin measurement (see polaron mechanism of PVC degradation) it was determined that the stabilization effect of alkyl phosphites improves with alkyl chain length increasing.290

Trinonylphenyl phosphite (considered non-toxic costabilizer) was used to stabilize PVC for medical applications.293 It performed its functions during processing but did not affect stability of PVC during γ-sterilization.293

In restabilization of PVC wastes, a system of costabilizers has been proposed and it performed well.294 The system contained phenolic antioxidant, phosphite, and HALS.294

Complex phosphite esters such as poly(dipropylene glycol) phosphites and alkyl bisphenol phosphites outperform simple phosphite esters as replacements in Ca/Zn stabilized systems.295 These and other similar phosphites are proposed now for partial (only a very small amount of zinc carboxylate is needed for betterment of initial color) or complete replacement of heavy metal stabilizers.296

Liquid organic phosphites, based on pentaerythritol, were patented303 for PVC stabilization. They are used in combination with Ba/Zn stabilizers to give phenol-free formulations.

PVC membranes for swimming pool liners were obtained by coating with PVC containing Ba/Zn carboxylate, alkyl phosphite, and benzophenone UV stabilizer.304 An older invention305 shows a combination of tin stabilizer with zinc carboxylate and phosphite – a direction which is exploited in order to reduce amount of tin stabilizers.

A composition comprising a mixture of two alkylaryl phosphites was patented. A mixture of two phosphites is a liquid at ambient conditions.306 Phosphites were stabilized against hydrolysis by the use of amines.307

Read full chapter
URL: https://www.sciencedirect.com/science/article/pii/B9781895198850500133

Synthesis: Carbon with One Heteroatom Attached by a Single Bond

Toru Minami, Kentaro Okuma, in Comprehensive Organic Functional Group Transformations, 1995

2.16.3.3.3 Michaelis–Arbuzov reaction of phosphinic acid esters with alkyl halides (Method C)

Tertiary phosphine oxides are also obtained by the Michaelis–Arbuzov reaction, which involves the reaction of phosphinic acid esters with alkyl halides to give intermediate phosphonium salts, followed by rapid decomposition (Scheme 18) <55JA3526>.

Scheme 18.

The use of the Michaelis–Arbuzov reaction to prepare tertiary phosphine oxides has the disadvantage that the alkyl halides (R2X) generated frequently react with the starting esters as alkylating agents.

An optically active phosphine oxide has been prepared by the Michaelis–Arbuzov reaction of asymmetric methyl ethylphenylphosphinite with methyl iodide, which proceeds with retention of configuration at phosphorus (Equation (114)) <75CC382>.

(114)
Read full chapter
URL: https://www.sciencedirect.com/science/article/pii/B008044705800159X

Fused Five- and Six-membered Rings with Ring Junction Heteroatoms

David J. Calderwood, Colin G.P. Jones, in Comprehensive Heterocyclic Chemistry II, 1996

8.03.6.1 Reactions at Phosphorus

The oxazaphospholidine (41) undergoes a Michaelis–Arbuzov reaction to give the phosphine oxide (46). This reaction was completely stereospecific with retention of configuration 〈89CC805〉. Reaction of (41) with borane-THF gave a stable complex (47) which has been used as a catalyst in enantioselective borane reductions of ketones 〈92CC287〉. Base-catalyzed methanolysis of oxazaphospholidine oxides and sulfides (32) and (33) gave the corresponding ring-opened products (48) resulting from Psingle bondO bond fission. This reaction proceeds with inversion of configuration at phosphorus 〈81TL477〉. Similarly, (32) and (33) react with Grignard reagents to give (49;R = alkyl, aryl) in almost quantitative yield 〈81TL571〉 (Scheme 1).

Scheme 1.

The dimethylamino group of (38) can be displaced by phenols to give mixtures of diastereomeric products (39). By controlling the reaction conditions, mixtures containing predominantly the kinetic or the thermodynamic diastereomer can be obtained 〈88OM59〉. The chlorodiazaphosphole oxide (31) on treatment with N,N-dimethylethylenediamine gave the phosphoramide (50) in good yield. Again, this reaction occurs with inversion of configuration at phosphorus (Scheme 1) 〈87JOC5320〉.

Read full chapter
URL: https://www.sciencedirect.com/science/article/pii/B9780080965185001702

Additions to C–X Π-Bonds, Part 1

P.R. Blakemore, in Comprehensive Organic Synthesis (Second Edition), 2014

1.15.3.2.2 Synthesis of phosphonates for the Horner–Wadsworth–Emmons reaction

Simple phosphonates are readily accessed by the Michaelis–Arbuzov reaction of trialkyl phosphites with alkyl halides (Scheme 47).258,259 The reaction works well for 1° alkyl halides, including α-bromoacetates, but it is less effective for 2° alkyl halides and branched 1° alkyl halides. The classical method involves strong heating of the alkyl halide and phosphite ester without solvent but more recently Mohanakrishnan and coworkers have demonstrated efficient room temperature Michaelis–Arbuzov reactions of benzylic (and heterobenzylic) halides, α-bromoacetates, and α-bromoketones, mediated by Lewis acid additives such as ZnBr2 and InBr3 (e.g., equation 91).260 This method is extendable to benzylic alcohols but fails to convert 1° alkyl bromides to phosphonates which in any event are easily made via the traditional pyrolysis route.

Scheme 47. The Michaelis–Arbuzov reaction of triethylphosphite with alkyl halides.

(91)

At low temperatures simple α-metallated alkyl phosphonates are well-behaved nucleophiles that can be readily derivatized by reactions with alkyl electrophiles (e.g., alkyl halides, alkyl sulfonate esters, epoxides, etc.) or via acylation with various types of acyl donors (e.g., esters, dialkylcarbonates, etc.).261 The latter chemistry provides for a synthesis of β-ketophosphonates which are not generally available by the conventional Michaelis–Arbuzov reaction from α-haloketones due to competing Perkow reactions (enol phosphate generation).258 For example, Pandey and coworkers achieved a high yielding synthesis of a complex β-ketophosphonate (431) during the total synthesis of pancracine by adding an excess of the lithiated form of diethyl methylphosphonate to ester 430 (equation 92).262 An alternative synthesis of the same phosphonate via the addition of a methyl ketone enolate to diethyl chlorophosphate (ClP(=O)(OEt)2) resulted in only a 12% yield. Both phosphonate alkylation and acylation chemistry were employed to good effect by Heathcock and Ott in their synthesis of a complex phosphonoacetate (435) representing the side-chain of spongistatin 2 (Scheme 48).263 This sequence aptly demonstrates that phosphonates are tolerant of many standard functional group interconversions and that their anion chemistry can be used to build up complex fragments before HWE olefination reactions.

Scheme 48. Illustrative synthetic manipulations of phosphonate anions en route to a spongistatin 2 fragment (435).

(92)

Many of the more desirable small α-phosphonocarbonyl building blocks (e.g., 429) are commericially available and these present a convenient starting point for the elaboration of HWE reagents of greater complexity or examples that contain unusual ligands on phosphorus. Net transesterification of alkoxy groups on phosphorus can be achieved via alcoholysis of dichlorophosphonates generated by treatment of (dialkylphosphono)acetates with PCl5. For example, Ando prepared the modified HWE reagent ethyl (diphenylphosphono)acetate (422) from inexpensive triethyl phosphonoacetate (429) in this manner by quenching phosphoryl dichloride 438 with phenol (Scheme 49).264 The same reagent is also available via acylation of the phosphonate anion derived from 437 under Barbier conditions, or via the Michaelis–Becker reaction between ethyl bromoacetate (440) and diphenylphosphite (439).265 Incorporating a phosphonoacetate moiety into a complex molecule is perhaps most easily achieved by ester formation between an alcoholic substrate and diethylphosphonoacetic acid (442) or a related compound. Acylation proceeds under standard conditions with carbodiimide reagents providing arguably the most straightforward means to effect the desired coupling. Using this method Kerr and Carson adorned hindered 3° alcohol 441 with a phosphonoacetate moiety to prime the substrate for a subsequent intramolecular HWE reaction en route to the Securinega alkaloid phyllantidine (equation 93).266

Scheme 49. Three syntheses of ethyl (diphenylphosphono)acetate (422).

(93)

Conjugate addition to vinyl phosphonates provides another derivatization tactic potentially amenable to the synthesis of complex substrates for HWE reactions. The concept can be applied via conventional anionic chemistry or via the addition of radical species,267 but for the product phosphonates to be of direct use in olefination the usual electron-withdrawing substituent must be present. A Michael acceptor such as ethyl α-(diethylphosphono)acrylate (445) is particularly useful in this regard and was used by Minami et al. as a linchpin in a multicomponent coupling process involving carbon-centered nucleophiles and aldehyde electrophiles (e.g., Scheme 50).268 The same kind of process can also be terminated by an intramolecular HWE reaction. For example, Hayashi and coworkers recently achieved a concise synthesis of the anti-influenza drug oseltamivir (Tamiflu) by intermolecular addition of nitroaldehyde 448 to vinyl phosphonate 445 followed by spontaneous intramolecular HWE reaction of the resulting phosphonate anion with the pendant aldehyde to yield cyclohexene 449 in high yield (equation 94).269 Heteroatom nucleophiles can be similarly employed to access other useful phosphonoacetate derivatives for HWE reactions. This tactic was utilized by Janecki and coworkers for the preparation of phosphonates (e.g., 452) needed for the synthesis of α-alkylidene-β-amino acids (equation 95).270

Scheme 50. Phosphonate anions generated in situ via conjugate addition can be used directly in HWE reactions.

(94)
(95)

Finally, we note an important direct synthesis of β-ketophosphonates from aldehydes using the Gilbert–Seyferth reagent (454). Thus, Roskamp and coworker disclose that aldehydes are converted into HWE olefination ready phosphonates by treatment with dimethyl diazomethyl-phosphonate (454) in the presence of stannous chloride.271 The illustrated synthesis of an advanced norhalichondrin B fragment (455) by Phillips and coworkers exemplifies the method (equation 96).272 An enantioselective variant of the Roskamp phosphonate synthesis was recently reported.273

(96)
Read full chapter
URL: https://www.sciencedirect.com/science/article/pii/B9780080977423001208

Bicyclic 5-5 and 5-6 Fused Ring Systems with at least One Bridgehead (Ring Junction) N

L. Micouin, in Comprehensive Heterocyclic Chemistry III, 2008

11.03.6.1 Reactions at Phosphorus

The oxazaphospholidine 72 can undergo an enantiospecific Michaelis–Arbuzov reaction with activated electrophiles, leading to phosphinamide 71 with complete retention of configuration at the phosphorus atom (Scheme 3) <1989CC805, 1997T11577>. Under the same reaction conditions, E-crotyl bromide led to the corresponding phosphinamide with retention of the double bound configuration. It is interesting to note that a Perkow reaction is observed with α-chloroacetophenone, leading to compounds 73 and 74 in a 7/3 diastereomeric ratio. The mechanism of this transformation has been discussed (Scheme 3).

Scheme 3.

The ring expansion of diazaphospholidine oxide 75 involves a stereospecific migration of phosphorus atom from N to a Csp2 center (Equation 1). The overall retention of configuration at the phosphorus center has been explained by a sequential intramolecular apical addition, Berry pseudo-rotation and apical elimination pathway <1999AGE1479>. A P–O to P–C stereospecific rearrangement occurs when only 2 equiv of lithium diisopropylamide (LDA) are used for the deprotonation step, leading to diazaphospholidine oxide 78 from compound 77 (Equation 2) <1999EJO1099, 1998CEJ1061>.

(1)
(2)

The ring opening of various phosphinamides occurs with inversion of the phosphorus configuration <1996TL39, 1996TL2853, 1998JP11027, 2005TL8677>, leading to interesting ligands such as 81 or 82 (Scheme 4). The chlorodiazaphosphole oxide 83 leads to phosphoramide 84 when treated with pinane 2,3-diol in the presence of NaH <2004TA47>. The same reaction with diamines has been reported <2001TA685, 2004TA1881>. A similar reactivity is observed with chlorodiazaphospholidines <2005ASC61, 2002ASC868>. Among various oxidation reactions described at the phosphorus center of diazaphospholidines <2002TL4025>, the sterospecific reaction of compound 85 with phenylazide leading to iminophosphorane 86 is noteworthy <1999JA5807>.

Scheme 4.

Read full chapter
URL: https://www.sciencedirect.com/science/article/pii/B9780080449920010038

Synthesis: Carbon With Three or Four Attached Heteroatoms

F. Sączewski, in Comprehensive Organic Functional Group Transformations II, 2005

6.21.1.2.1 N-Alkylimino derivatives with one P or As function

Title compounds 186 (Figure 5) were usually obtained by Michaelis–Arbuzov reaction of the carbamimidic chlorides with trimethyl phosphite (P(OMe)3). Another method for preparation of 187 consists in the attack of alkali metal organoarsenides at N,N′-dialkyl carbodiimides followed by hydrolysis or alkylation of the intermediary formed (lithioamidino)arsines <1995COFGT(6)639>.

Figure 5.

Recently, the phospha(III)guanidine compounds of the general formula Ph2PC(NR) (NHR);R = Pri, cyclohexyl) have been prepared in good yields as described in Scheme 55. Lithium diphenylphosphide obtained by treatment of Ph2PH with BuLi is allowed to react with suitable carbodiimide giving the corresponding lithium phospha(III)guanidinate. Quenching the reaction with triethylamine hydrochloride yields the neutral phospha(III)guanidines 188 <2002CC2794, 2003JCS(D)2573>.

Scheme 55.

Read full chapter
URL: https://www.sciencedirect.com/science/article/pii/B0080446558001422

Synthesis: Carbon with Three or Four Attached Heteroatoms

Angela Marinetti, Philippe Savignac, in Comprehensive Organic Functional Group Transformations, 1995

6.09.3.1.3 One halogen, one chalcogen, and two phosphorus functions

The first approach to this kind of compound is a traditional Michaelis–Arbuzov reaction between triethyl phosphite and trichloromethyl phenyl ether (Scheme 13) <70JPR475>. Analogous tetraethyl halomethoxymethylene bisphosphonates are obtained by radical halogenation of the corresponding methoxymethylene bisphosphonates (Equation (21)) <84ZOB2504>.

Scheme 13.

(21)

Unexpectedly, two products were obtained during the reaction of the methylene bisphosphonate salts, [((RO)2PO)2CH]M+ (M = Li, Na, K), with CF3SCl. The expected sulfenyl derivative is obtained in a mixture with chloro(trifluoromethylthio)methylene bisphosphonate, which results from a chlorination process (Equation (22)). The use of aluminum trichloride reduces the proportion of chlorinated by-product (16%) with respect to the sulfonylated product (74%) <85JCS(P1)1935>. CClF2SCl and CCl2FSCl have also been used in analogous reactions.

(22)
Read full chapter
URL: https://www.sciencedirect.com/science/article/pii/B0080447058002235

Synthetic Methods II – Chiral Auxiliaries

S. Jugé, ... A. Tessier, in Comprehensive Chirality, 2012

3.17.4.2 Use of Tartrate Derivatives

The asymmetric synthesis of P-stereogenic phosphine oxides, involving a Michaelis–Arbuzov process, was also achieved starting from the cyclic phosphonite 64, prepared from the diol (XX) derived from tartaric acid.64 After reaction with benzyl halides, the phosphonite 64 afforded the corresponding phosphinate by P–O bond cleavage. The diastereoselectivities ranged from 23 to 100% depending on the stereoelectronic nature of the benzyl halide used. The second P–O bond cleavage was then performed by addition of a Grignard reagent affording the corresponding tertiary phosphine oxides in good yield but with only low or moderate enantiomeric excesses (<45% ee).

The use of the diol (XXI), also derived from tartaric acid, was reported by Hoppe and Eggert for the preparation of enantiomerically pure cyclic H-phosphite 65, which is useful for the synthesis of α-aminophosphonic acid derivatives 68 and 71 (Scheme 16).65 The BF3-catalyzed addition of 65 to dihydrothiazole 66 afforded the phosphonates 67 in a diastereomeric ratio 2:1 favoring the (R)-diastereomer. After separation, the hydrolysis of 67 gave rise to enantiopure α-aminophosphonic acid 68 (Scheme 16(a)).

Scheme 16. Stereoselective synthesis of α-amino phosphonic acids using tartaric acid (XVIII).

Rico–Lattes and coworkers have described another route for the synthesis of amphiphilic α-aminophosphonic acids 71 using the chiral H-spirophosphorane 69 derived from the tartaric acid (XVIII) (Scheme 16(b)).66 The Pudovik addition of the H-spirophosphoranes 69 to the long-chain aldimines 70 (n=12, 16, or 18), followed by the separation of diastereomers and their acidolysis, afforded the corresponding enantiopure α-aminophosphonic acids 71 (Scheme 16(b)).

In the late 1980s, Giordano's group published the first asymmetric synthesis of fosfomycin 74 (Scheme 17).67 The first step is based on the use of various tartaric acid derivatives (XVIII) as chiral auxiliaries for the condensation with the (Z)-1-propenyl phosphonic dichloride 72. After hydrolysis, the phosphono monoester 73 was obtained in 70% yield. After epoxydation via a bromhydrin intermediate, which was purified by crystallization, the enantiopure (1R,2S)-1,2-epoxypropylphosphonic acid (fosfomycin) 74 was obtained in 35% yield after hydrolysis (Scheme 17).

Scheme 17. Stereoselective synthesis of fosfomycin 74 using tartaric acid (XVIII).

Read full chapter
URL: https://www.sciencedirect.com/science/article/pii/B9780080951676003177