<<

8 CHIMIA 2016, 70, No. 1/2 The French connecTion

doi:10.2533/chimia.2016.8 Chimia 70 (2016) 8–19 © Swiss Chemical Society Advances in Homogeneous Catalysis Using Secondary (SPOs): Pre-ligands for Metal Complexes

Thierry Achard*

Abstract: The secondary phosphine oxides are known to exist in equilibrium between the pentavalent phos- phine oxides (SPO) and the trivalent phosphinous acids (PA). This equilibrium can be displaced in favour of the trivalent tautomeric form upon coordination to late transition metals. This tutorial review provides the state of the art of the use of secondary phosphine oxides as pre-ligands in transition metal-catalysed reactions. Using a combination of SPOs and several metals such as Pd, Pt, Ru, Rh and Au, a series of effective and original transformations have been obtained and will be discussed here. Keywords: Metal-catalysed reactions · Phosphinito complexes · Phosphinous acids (PA) · Secondary phosphine oxides (SPO)

Phosphorus is one of the most impor- For trialkyl phosphine (TPO) the presence of a base the weak PA acids[9] can tant atoms for as it is involved in nu- pentavalent phosphorus loses its coordi- be easily deprotonated into anionic phos- merous biological processes.[1] Humans nating ability in favour of the at- phinito ligands that generate P-coordinat- mainly use phosphorus in its phosphate om. However, secondary phosphine oxides ed type (b) complexes or O-metalated spe- form as fertilizer to feed and grow plants. (SPOs) are in equilibrium with pentavalent cies of type (d) (Scheme 2). Such anionic Furthermore, our cells require pentavalent phosphine oxide and trivalent phosphinous ligands combine with Mo, W, Mn, Re, Fe, phosphorus to build membranes (phos- acids.[5] It is well established that in pres- Pt and Au to form preferentially type (b) pholipids), transport energy (ATP) and ence of a transition metal, the equilibrium complexes. In contrast, with Fe, Rh, Ir replicate (DNA, RNA). Chemists prefer is displaced in favour of the phosphinous and Ag, the O-coordination (type d) is fa- the use of the trivalent phosphorus and tri- acid (PA) through coordination of the voured. Representative X-ray structures of alkyl/aryl in particular. These phosphorus atom (Scheme 1).[6] type (a) and (c) complexes are highlighted are involved in several standard The secondary phosphine oxides pres- in Fig. 1.[10] organic transformations such as the Mitsu- ent a remarkable dichotomy in their coor- Phosphinito ligands can adopt the nobu-, Vilsmeier-, and Wittig reactions and dination mode (monodentate versus biden- bridging mode (e) on several transition many more. But the most exploited facet tate). In addition, the P=O group displays metals such as Cr, Mn, Re, Fe, Co, Ni, Ir, of phosphine derivatives is their ability to an interesting ambident character that can Pd, Pt and Au.[6a,7] In this case, the sys- bind transition metals.[2] In organometallic chelate either late-transition metals (M1) tem often displays a six-membered ring chemistry and homogenous catalysis they through the ‘soft’ phosphorus atom or unit [M (µ-SPO )] involving two metals 2 2 represent the ligand of choice due to their early transition metals (M2) through the bridged by two SPO ligands. Outstand- ability to stabilize an exceptionally wide ‘hard’ oxygen atom (Scheme 2).[6a,7] ingly, a few examples with Mo, W, Rh and variety of metal centres.[3] However, one Neutral ligands generate complexes of Ir show a rare dihapto coordination mode major issue still remains for most trivalent type (a) and (c): the P-coordination mode of type (f). This monomeric unit was often phosphorus compounds: their sensitivity to (a) was found to accommodate Fe, Ir, Pd, postulated in equilibrium with the dinucle- oxidation. So, their preparation often re- Pt, Rh and Ru transition metals, whereas ar complex (e).[7a] Representative X-ray quires an inert gas atmosphere, not exactly mode (c) was only described with Cr, Mo, structures of type (e) complexes are high- an advantage for routine use and scale-up. W, Re, Ti and observed for Rh.[6a,7,8] In the lighted in Fig. 2.[10] Protection through BH groups can be 3 achieved, but the in situ release of the free Scheme 1. SPO-PA phosphine needs the presence of an amine SPO tautomeric equilib- that must be done prior to any coordina- O rium and chelate [4] formation. tion to transition metals. Unfortunately, 1 P the P-BH adduct is often unstable towards R H 3 2 a series of acids and Lewis bases. R

R2 R1 O P OH [M] [M] H [M] R1 P [M] P R2 Base OH *Correspondence: Dr. T. Achard O P R2 R1 n University of Strasbourg, IPCMS: UMR 7504 CNRS R2 R1 Département des Matériaux Organiques (DMO) PA 23 rue du Loess BP 43, Strasbourg 67034 bidentate monodentate E-mail: [email protected] The French connecTion CHIMIA 2016, 70, No. 1/2 9

Scheme 2. for this favoured coordination mode. How- Coordination mode of hard binding site ever, steric hindrance at the phosphorus secondary phosphine atom can prevent the formation of such a oxides. R2 OH complex by preferring a trans configura- Phosphinous acid P tion between the two SPO units.[15] This (σ3λ3) R1 soft binding site bidentate behaviour has been observed for numerous transition metals such as Cr, Mo, W, Ru, Rh, Ir, Pd and Pt.[6a,7] Representa- monodentate tive X-ray structures of type (g) complexes are highlighted in Fig. 3.[10] P-coordinated O-coordinated

OH O H [M] O P R2 [M] P R2 [M] P R2 [M] O P R2 R1 R1 R1 R1 (a) (b) (c) (d)

bridge bidentate

1 R1 R2 R1 R2 R1 R 2 R2 P O R P O P O P O [M] [M1] [M2] [M1] [M2] [M] H P O P O 1 2 1 2 2 [15] µ-(P,O) η2-(P,O) R R R R Pt(κ -OAc){[(t-Bu)2-PO]2H} (e) (f) (g) (h) Fig. 3. PLUTO representation of the X-ray crystal structure type (g) coordination. Most Fig. 1. PLUTO rep- of the atoms are omitted for clarity. resentation of the Adapted from ref. [15]. X-ray crystal structure types (a) and (c) co- Captivatingly, the acidic bridging hy- ordination. Most of drogen can be replaced by a wide variety the hydrogen atoms of metals (Al, Co, Cu, Ir, Mn, Na, Ni, Rh, are omitted for clar- ity. Adapted from refs Th, T1, U, V and Zn) providing bimetal- [11] and [12]. lic species of type (h). Other coordina- tion compounds have been reported in the literature in which the is replaced by non-metal atoms such as B, Si and P. Representative X-ray structures of type (h) complexes are highlighted in Fig. 4.[10] Electronic properties of SPO ligands [11] [12] [(Ph,t-Bu)P-(OH)AuCl] [RuCl2(CO3)(Ph,t-BuP-OH)] have been investigated by Martin et al. on various metal carbonyl SPO-complexes by the measurement and/or calculation of the IR frequencies of CO.[18] It appears that generally the phosphinous acid form (a) is less donating than the correspond- ing trisubstituted phosphine ligands. How- ever, abstraction of the hydrogen makes the generated phosphinito (b and e) an extremely strong donating ligand (Tol- man’s electronic parameter (TEP) ranging from 2006 to 2028 cm–1). According to both their TEPs and υ˜ frequencies, bi- CO dentate anionic ligands of type (g) display the strongest net-donation (close to phos- phonium ylides) outclassing the N-hetero- cyclic carbenes. Finally, the net-donation decreases in the presence of an alkali ion [Pt (µ-OPPh ) (MePh ) ][13] [Pd (µ-OPPh ) {(N,N-Ph P)N(Me)CO(Me)NPPh }][14] 2 2 2 2 2 2 2 2 2 2 M2 (type e and h) and the magnitude of this Fig. 2. PLUTO representation of the X-ray crystal structure type (e) coordination. Most of the hy- effect directly correlates to the hardness of drogen atoms are omitted for clarity. Adapted from refs [13] and [14]. the counter ion. Electronic modulation can be easily achieved by varying the nature of the group bound to the phosphorus atom. Interestingly, two SPOs ligands at- chelate of type (g). This cis arrangement Interestingly electron-poor SPOs (depict- tached on the same metal centre can form results from a combination of a neutral ed in Scheme 1) shift the equilibrium in an intramolecular hydrogen bond into the ligand of type (a) with an anionic ligand favour of phosphinous acid form making backbone and so ‘mimic’ a bidentate li- of type (b). The formation of a strong hy- them ideal candidates to form metal com- gand to give a monoanionic diphosphorus drogen bond is definitely the driving force plexes.[10] 10 CHIMIA 2016, 70, No. 1/2 The French connecTion

acyl chlorides with boronic acids[33] or organostannanes,[34] direct C(3) arylation of indoles,[35] conjugate addition of arylsi- loxanes,[36] intramolecular α-arylations of amides,[37] and oxidative esterification of O [38] Ti . These catalysts allow cross- P coupling reactions to proceed at room tem- Pd Rh perature even with deactivated aryl chlo- P P [24a] F P rides and fluorides. However, the development of ‘greener’ O OO reaction conditions such as replacement B of toxic organic solvents by is still F an issue. Hu and co-workers reported an efficient Suzuki cross-coupling reaction of aryl halides in water (Scheme 3).[39] POPd Fig. 4. PLUTO representation of the X-ray crystal structure type (h) coordination. Most of the (R = t-Bu) catalyst (0.2 mol%) in presence hydrogen atoms are omitted for clarity. Adapted from refs [16] and [17]. of phenyl boronic acid, appropriate phe- nyl halides 1 and a phase transfer agent Steric properties have to be taken into Secondary phosphine oxides bind sev- (n-Bu )NBr in refluxing water gave biaryl 4 account as they play a crucial role on both eral transition metals giving a rich family 2 in good yields.[39a] The authors reported the formation of metal complexes and of potential catalysts. The following sec- that substrates with electron-donating or their coordination mode. Indeed, in square tion will discuss their catalytic activities -withdrawing substituents were converted planar complexes such as PtCl (cod) and classified by the metal centre involved in to functionalized biphenyls. Using these 2 ([{Rh(CO) Cl} ], the bulky and electron- these processes. heterogeneous conditions the authors were 2 2 rich SPOs such as (t-Bu) P-OH react to able to promote the reaction also using het- 2 form exclusively trans-platinum com- eroaryl bromides 3 and 5.[39b,c] This system plexes [ML (t-Bu POH) ].[15,17] Moreover, Palladium–Phosphinous Acid allows an easy separation of the catalyst 2 2 2 the same ligands do not react with bulkier Complexes generating the desired compound in high [Rh(cod) ][BF ] precursors mainly due purity. 2 4 to steric hindrance of the t-Bu groups.[17] Coupling Reactions The combination of imidazole-based Furthermore, because of that it is diffi- SPO complexes with a palladium centre secondary phosphine oxide[40] and pal- cult to locate the two phosphorus groups have remained the most widely investigat- ladium acetate (under mild conditions: in the required cis geometry to form the ed since the report by Li and co-workers [25] 60 °C), promoted a highly regioselective ‘pseudo-bidentate’ complex and only rare in 2001 on applications in cross-coupling Heck coupling reaction as described by examples have been described in the litera- reactions. Complexes such as POPd and Shaikh and Hong.[40] They demonstrated ture.[15,18,19] POPd1 (Fig. 5) have been successfully ap- that the coupling of aryl halides and olefins Finally, it is important to note that the plied in a broad range of cross-coupling in the presence of K CO in DMF solvent 2 3 electronic donation of such a ligand can be reactions such as the Heck reaction,[26] was promoted using 1.5 mol% of complex dramatically modified during the course of Stille coupling,[27] Sonogashira cou- 7 resulting in the formation of exclusive the reaction. Indeed, phosphinous acids are pling,[28] Hiyama coupling,[29] Kumada trans-stilbene derivatives in good yields able to reversibly form strong electron-do- coupling,[30] Negishi coupling,[31] Suzuki- (Scheme 4).[41] These reaction conditions nating phosphinito species upon deproton- Miyaura coupling,[32] cross coupling of could be applied to the synthesis of α,β- ation. This facet is quite unique and could be the source of their exceptional behav- iour as catalysts. R R R R R R Since 1986 and the first application re- R R O OH ported by van Leeuwen et al. on the (SPO) HO P Cl O P Cl P O O P O P Pt-catalysed hydroformylation of ole- Pd H Pd Pd H H Pd O O O fins,[20] the use of SPO-catalysts has gained Cl P OH O P Cl P O O P Me Me R R R R R R Me Me great interest. The main reasons for their R R use as pre-ligands could be summarised as 2 follows: i) more stable to oxidation than POPd POPd1 (κ -OAc)Pd(PA)2 HASPO the parent phosphine;[21] ii) fairly simple syntheses and ease to modulate both ste- Fig. 5. Selected phosphinous acid-coordinated palladium complexes. ric and electronic factors;[15,21,22] iii) strong electron-donating groups[18] between π-accepting phosphite-like compounds and beyond N-heterocyclic carbenes.[21,23] All these properties are indeed suitable for both organometallic chemistry and ho- mogenous catalysis. In this review, recent advances in this area will be discussed. We will confine mostly our discussion to the past five years on catalytic reactions; cross-coupling reactions will be briefly mentioned as reviews were recently pub- lished on this topic.[24] Scheme 3. Suzuki coupling reaction in water catalysed by SPO-Pd. The French connecTion CHIMIA 2016, 70, No. 1/2 11

unsaturated esters from vinyl ester sub- strates. Various functional groups such as hydroxyl, carbonyl, nitrile etc. were toler- ated using this mild procedure. The Hong group also reported the prep- aration of a quinolyl-substituted secondary phosphine oxide hemilabile pre-ligand 8. In combination with palladium this modi- fied SPO 8 catalysed C–C bond formation in a three-component reaction of olefins 9, aryl halides 10 and to form polycyclic compounds 11 (Scheme 5).[42]

Cycloaddition Reactions SPO-Pd complexes displayed very in- teresting cycloaddition reactions leading to Scheme 4. SPO-Pd catalysed Heck coupling reaction. bicyclic compounds. Buono and co-work- ers recently revealed unprecedented SPO- Pd 14 (SPO-Pt 18 is described in the next section) catalysed [2+1] cycloadditions of alkynes with norbornadiene derivatives, which afford only the vinylidene cyclopro- panes 12 rather than the expected standard alkynyl products (Scheme 6). This strategy was extended to afford bicyclo[3.2.1]octa- dienes by a tandem cycloaddition-ring ex- pansion reaction.[43] This time Scheme 5. SPO-Pd catalysed mutli-component reaction (MCR). in presence of propargylic acetate gives a mixture of an unusual ring expanded major product 13 and cycloadduct 12 as minor base-sensitive functional groups contain- due to the partial decomposition of the products (Scheme 6).[44] ing nitriles under quite mild (70–100 °C) substrate which generated HCN that defi- and neutral reaction conditions in water.[45] nitely poisoned the catalyst. Attempts to Since that time, a number of other studies modify the substituents on the phosphorus Platinum–Phosphinous Acid have been conducted using commonly a atom to increase the catalytic activity were Complexes water/organic solvent mixture as reaction unfruitful. According to the authors, the medium.[46] formation of the catalyst PtCl(PMe OH) 2 In comparison with SPO-palladium The catalytic hydration of cyanohydrins {(PMe O) H} is governed by steric effects, 2 2 complexes, only a few reactions have been 16 remains an important and challenging generating (when P-alkyl ≠ Me) the inac- described in the literature using a SPO- transformation due to the easy decompo- tive dimer species instead of the expected platinum combination. Since the hydro- sition of cyanohydrins under basic con- active monomer complex.[49b] formylation depicted by Van Leeuwen et ditions and high temperature, producing al.,[20] only nitrile hydration and cycload- undesired HCN.[47] The cyanide released Cycloaddition Reactions dition reactions have been reported so far. in solution binds irreversibly the metallic The first carbon–carbon bond forma- centre of the catalyst, annihilating its cata- tion using SPO-Pt complexes was re- Nitrile Hydration lytic activity.[48] Latterly, PtCl(PMe OH) ported using acetate Pt(κ2-acetato)-{[(R) 2 In 1995, Ghaffar and Parkins reported {(PMe O) H} was used in a pure aque- (Ph)PO] H} as catalyst. These acetate- 2 2 2 nitrile hydration using PtH(PMe OH) ous medium to catalyse the hydration of complexes 18 promoted benzylidenecy- 2 {(PMe O) H} and PtCl(PMe OH) cyanohydrins (Scheme 7).[49] It showed clopropanation of bicyclic alkene through 2 2 2 {(PMe O) H} (15) under homogenous moderate activity compared to classic ni- an intermolecular [2+1] cycloaddition 2 2 conditions.[45] The hydrido-platinum(ii) trile substrates, the reaction was slow and (Scheme 6).[19a] The reaction proceeds un- complex selectively catalyses the hydra- only few turnovers could be achieved.[49a] der mild reaction conditions (20 h at 55 °C) tion of a wide range of bulky, unsaturated, This lack of reactivity was assumed to be and the was found crucial to

Scheme 6. Pd or Pt/SPOs-catalysed [2 + 1] cycloaddition and rearrangement reactions. 12 CHIMIA 2016, 70, No. 1/2 The French connecTion

obtain methylenecyclopropane adducts in good yield (up to 80%). Interestingly, even hetero olefins such as diaza- and oxyaza- bicyclic substrates resulted in the formation of the products in excellent yields. More recently, Achard et al. described an unprecedented platinum-SPO 18 cata- lysed regio- and diastereoselective tandem Scheme 7. Hydration of cyanohydrins catalysed by SPO-platinum complexes. [2+1]/[3+2] intermolecular cycloaddi- tion (Scheme 8).[50] This sequence allows the formation of tricyclic compounds 19 in good overall yields by the reaction of norbornadiene with several alkynes. In- terestingly, using simple alkynes, such as phenyl , only provides a methy- lenecyclopropane (MCP) compound 12 as already observed for palladium catalysis (Scheme 6). This highlighted the key role Scheme 8. Pt/SPO-catalysed tandem [2+1]/[3+2] cycloaddition reaction. of the homopropargylic oxygen atom in this transformation where the MCP compound acid displayed better activity for this reac- nium complex is active in the hydration of is the intermediate of this tandem reaction. tion.[54] various nitriles including several cyanohy- The authors later expanded this method drins (α-hydroxynitriles). As observed for to the intermolecular cycloaddition of Nitrile Hydration other catalysts, this SPO-Ru complex is alkylidenecyclopropane and terminal As described in the platinum section, sensitive to cyanide poisoning and so was alkynes.[50] Tricyclic architecture com- hydration of cyanohydrin could be ca- not good enough to transform the pounds 20 were generated bearing various talysed by transition metal complexes (Pt, cyanohydrin to σ-hydroxyamide. Despite functional groups such as hydroxyl, ether, this, [RuCl (p-cymene){PMe (OH)}] 25 Mo, Ru). Ruthenium catalysts are particu- 2 2 2 trimethylsilyl ether, trimethylsilylethyne, larly effective in nitrile hydration reactions (Fig. 6) remains the best catalyst, having sulfonamide, phthalimide and sulfonyl. and promising results have been described the highest reaction rate described in the using phosphine[55] and SPO ligands.[48,56] literature for such transformations. Tyler and co-workers used the ability of Inspired by these results, Cadierno Ruthenium–Phosphinous Acid SPOs to form hydrogen bonds in the sec- and co-workers, described the easy prepa- Complexes ration and the use of [RuCl (p-cymene) ondary coordination sphere to activate 2 {PMe (OH)}] complex 26 as catalyst water molecules in order to increase the 2 2 Since the seminal paper of Acker- reaction rate of nitrile hydration.[56] The in the hydration of nitrile.[57] This SPO mann et al. on the directed ortho-ary- [Ru(p-cymene)Cl ] metal precursor, in the Ru(iv) complex displayed a remarkably 2 2 lation of catalysed by a ruthe- presence of appropriate SPOs, generates high catalytic activity (TOF = 100 h–1) in nium/SPO combination,[51] other C–H the desired [RuCl (p-cymene){PR (OH)}] pure water without the assistance of any 2 2 2 functionalizations[23b,52] have been de- species. Good results were obtained with acidic or basic co-catalyst. The transfor- scribed using this base-assisted cata- PMe (OH) derivatives (TOF up to 32 h−1 at mation proceeded cleanly under very mild 2 lytic system.[53] The use of SPOs in such 100 °C) which displayed the highest hydro- conditions (60 °C) with low catalyst load- transformations shows many advantages: gen bond accepting characters of the SPOs ing (1 mol%, see Scheme 10a) generating i) accelerated reaction, ii) use of a broad evaluated by the authors. Thereby, the ac- desired amides 29. However, catalyst 26 is range of organic (challenging tivation of water molecules by the ligand able to hydrate demanding cyanohydrin 30 aryl chlorides or tosylates), iii) tolerance would favour its nucleophilic addition to and can be easily recovered at the end of towards a lot of functional groups (ester, the metal-coordinated nitrile. The ruthe- the reaction (Scheme 10b). ketones, cyano, etc.), iv) the robustness of their rutheno-complexes that allows the re- action to proceed in water.[53]

Direct C–H Bond Arylations Recently, Ackermann et al. showed that sterically hindered SPO 21 combined with ruthenium was even able to promote the first direct arylation of heteroarenes Scheme 9. Direct CH bond arylation catalysed by SPO-Pt complexes. such as 2-pyrimidyl-substituted indoles 22 (Scheme 9).[52c] This base-assisted cata- lytic system allows the use of cheaper aryl Fig. 6. Selected chlorides 23 giving satisfactory yields of phosphinous acid-co- ordinated ruthenium arylated products 24. For aryl bromide complexes. substrates, even if SPOs 21 were able to catalyse the reaction, the most efficient catalysis was performed with a bulky car- boxylic acid. Finally, bis-adamanthyl SPO 21 cataly- ses the direct arylation of phenol such as 2-phenoxypyridine but again carboxylic The French connecTion CHIMIA 2016, 70, No. 1/2 13

The authors extended their studies us- In the presence of SPO-Ru complexes ing a cationic ruthenium complex derived and silver triflate salts, electron-rich are- from a Binap core. The triflate complex nynes 37 (R = OMe) reacted to give bi- [RuOTf{(η6:κ1(P)-PPh -Binaphthyl} cyclic compounds 38. Ru complexes with 2 {PPh (OH)}][OTf] (27b) is an efficient electron-rich SPOs such as Cy POH were 2 2 catalyst for the hydration of nitrile in pure found to be the most active catalysts giving water.[58] However, this metal complex was full conversion after 8 h with only 2 mol% found to be much more sensitive to steric of ruthenium (Scheme 14). Less active are- hindrance as 24 h were required to achieve nynes 37 (R = H) need thermal activation full conversion of bulkier nitriles. Finally, up to 80 °C and provide mostly the iso- cationic SPO-Ru 27b remains less active mer 39a. The author shows that there is no compared to neutral cymene 25 and bisal- relation between the shape of the catalyst Scheme 10. Catalytic hydration of nitriles to lyl-ruthenium 26 complexes. Noteworthy, amide in water catalysed by SPOs-Ru com- and the observed selectivity; the selectivity this ruthenium complex 27b catalyses the plexes. seems only related to the steric effect of the rearrangement of aldoximines 32 to am- .[12] However, in this case ides via a dehydration/hydration sequence combination of the less donating and more involving the nitrile intermediate. Good re- sterically demanding t-BuPhPOH ligand sults were obtained (conversions up to 99% and ruthenium efficiently catalyses this and yields up to 87%) using the same reac- reaction contrasting with the previous sub- tion conditions (5 mol% of [Ru] in water at strate. This suggests a strong relationship 100 °C, Scheme 11). between ligand and substrate structures and so different rate-determining steps Hydrogenation Using Ru-Nano- Scheme 11. Catalytic rearrangement of al- might be considered for the two different particles doximines to amides catalysed by SPOs-Ru substrates. According to the author, elec- In a relevant observation, Wang and complexes. tron-poor complexes seem to activate ef- Buhro reported that the presence of dioc- tylphosphine oxide (impurities in Cy P) 3 Scheme 12. Hydro- strongly altered the morphology of CdSe genation of benzene [59] nanocrystals during their formation. derivatives. Inspired by this work, Van Leeuwen and co-workers investigated SPOs for the stabilisation of ruthenium nanoparticles (RuNPs).[60] SPOs possessing different electronic and steric properties were com- bined with the nanoparticles in a SPO/Ru: action conditions used for the synthesis of ficiently the alkyne, increasing the cationic 0.2/1 ratio. Diphenylphosphine oxide 33 SPO-Ru complexes. Using one equivalent character of the resulting vinyl-ruthenium provided the most narrow size distribution of SPO under mild conditions generated intermediate and so favour nucleophilic at- ofparticlesminimizingtheiragglomeration only the O-coordinated compounds 35a- tack. However, for electron-rich substrate (average diameter 1.6 nm; [Ru (Ph P(O) d, in contrast P-coordinated compounds 37 (R = OMe), the rate-determining step 147 2 H) ]). Despite the ability of SPOs to form 36a-g were obtained using harsher reac- seems to be no longer the nucleophilic 13 very stable molecular complexes,[17,23a] tion conditions (two equiv. of SPOs at 110 attack, in this case the reductive elimina- they were found to be very efficient to sta- °C, see Scheme 13). As observed for other tion process would be accelerated by the bilize Ru-L NPs. The surface of the par- metals, sterically demanding SPOs such as electron-rich ligands. t-Bu P(O)H were unable to coordinate the ticles was carefully analysed showing that 2 all the SPO pre-ligands were coordinated metal centre by either the oxygen or phos- in a network of hydrogen bonds. phorus atom. The authors evaluated these SPO- RuNPs in the hydrogenation of aromatic compounds 34 (Scheme 12). These SPO- RuNPs are highly active using only <0.01 mol% of nanoparticles in neat conditions. The catalyst loading could be reduced to 1:16000 whilst maintaining the high yields (TOF 2700 mol h–1; TON 16200) of hydro- genated compounds even for deactivated Scheme 13. Preparation of Ru-SPO complexes. trifluorotoluene substrates.

Cycloisomerisation of Arenynes According to the literature it is not triv- ial to synthesise Ru/SPOs complexes and only a few arene-containing complexes have been well characterised.[55,57,61] Very recently, Clavier and co-workers described the preparation of ruthenium carbonyl complexes bound either to one SPO or two PA ligands.[12] They showed interesting co- ordination behaviour depending on the re- Scheme 14. Cycloisomerisation of arenynes. 14 CHIMIA 2016, 70, No. 1/2 The French connecTion

Gold–Phosphinous Acid Complexes

Coordination of SPO on gold is rela- tively unexplored, indeed only a few SPO- gold complexes are described in the lit- erature by Schmidbaur and co-workers.[62] Mostly they describe the coordination Scheme 15. Hydrogenation of α,β-unsaturated aldehydes. chemistry of the diphenylphosphine ox- ide derivative giving complexes such as ingly this catalytic combination is able to eroatom-substituted phosphine oxides).[66] [Ph P(OH)AuCl] . promote the reaction even with deactivated These ligands did not improve the activity 2 2 substrates 44. Finally, the authors success- and the n-selectivity was found to be lower Hydrogenation Using Gold fully applied the [t-BuPhP(OH)AuCl] cat- than with SPO-Rh. Furthermore, a side re- Nanoparticles alyst to both the methoxy- and hydroxyl action involving free SPO and Until last year no SPO-gold complexes cyclisation giving high yields in each case. products generated α-hydroxyphosphonic were used either in homogenous and het- acid diesters and so reduced the concen- erogeneous catalysis. Stimulated by ob- tration of pre-ligand and the activity of the taining stable SPO-RuNPs, Van Leuween Rhodium–Phosphinous Acid resulting catalyst. and co-workers used [t-Bu(naphthalen- Complexes 1-yl)P(OH)AuCl] as a precursor to suc- Hydrogenation 2 cessfully stabilise gold nanoparticles.[63] The first application of Rh-SPO com- Association of SPOs with a rhodium Their preparation requires two-step syn- plexes was disclosed in the early 1980s centre applied to hydrogenation reactions theses starting from [(tht)AuCl]; the final when Matsumoto’s group used them as was first achieved by Feringa and de Vries. step produces, in presence of NaBH , the catalysts for hydroformylation of linear [67] This work followed their seminal report 4 desired SPO-Au nanoparticles. This pro- olefins.[64] Since then, other homogenous on the use of iridium SPOs complexes for cedure provides monodisperse particles catalytic reactions have been investigated asymmetric hydrogenation of imines.[68] (1.24 nm) which are highly soluble in apo- giving interesting results. The following Although these catalysts could be dis- lar solvents. Thankfully their non-solubili- section will highlight the latest results of cussed here, the SPO-iridium complexes ty in allows a simple purification Rh-SPO catalysed reactions. will be discussed in their own section and by precipitation. Structural and spectral more closely associated to the asymmetric analyses indicate that SPO 40 coordinates Hydroformylation catalysis part. the gold surface in monodentate fashion as Börner and co-workers reported the use In 2010, Pugin and Pfaltz synthesized an anionic ligand instead of the classic H- offouraryl-substitutedSPOsaspre-ligands new SPO-type ligands derived from chi- bonded ‘bidentate’ coordination. in the rhodium-catalysed hydroformylation ral ferrocenyl and menthyl, JoSPOphos The authors report a highly chemose- of olefins (cyclohexene, 1-octene).[65] Us- and TerSPOphos, respectively (Fig. 7).[69] lective hydrogenation of α,β-unsaturated ing only 26 ppm of [Rh] with a Rh/L ratio These ligands, when associated with [Rh(nbd) ]BF under mild conditions, (2 h, aldehydes 41 in which only the thermo- of 1:4, this catalytic combination produced 2 4 25 °C, 1 bar H ) showed high activities and dynamically more robust C=O group is the desired aldehyde in moderate to good 2 reduced in alcohol to give products 42 yield. In all cases, SPO-Rh catalysts exhib- enantioselectivities in the hydrogenation (Scheme 15). The homogenous catalytic it a better activity than the corresponding of alkenes (90–99% ee).[69] The reactions system displayed a high tolerance toward PPh complex. For 1-octene, good reactivi- were quite fast with turnover frequencies 3 a wide range of functional groups such as ty (up to 88% yield) was observed, but only (TOF) around 2000–20000 h–1. However, alkynes, cyano, nitro, sulphur heteroaro- moderate n-selectivity (53%) was achieved moderate asymmetric inductions were ob- matic rings, esters and carboxylic acids indicating a high isomerisation rate. Curi- served with this catalytic system on hydro- giving again a very high selectivity up ously, electron-poor pre-ligands failed to genated α- and α-ketoesters (89% ee). Of to 99%.[62] Of particular interest for this increase the n-selectivity and for the most note, for such substrates the best enantiose- unprecedented selectivity is the hydroge- electron-deficient pre-ligands the reaction lectivity was obtained combining a ruthe- nation of acrolein to . Other did not occur at all, leaving the 1-octene nium source and JoSPOphos (R1 = t-Bu; AuNPs were evaluated to give no reactiv- totally intact. Finally protic solvents such 92% ee). Interestingly, the authors clearly ity, highlighting the crucial role of the SPO as methanol are crucial to observe a good showed that the absolute configuration at in the reported catalytic activity. level of n-selectivity as they maintain the the phosphorus atom governs the direction integrity of the pseudo ‘bidentate’ coordi- of the asymmetric induction. Cycloisomerisation of Enynes nation.[17] In an extension of this work, the In 2012, Ding and co-workers reported More recently, Fensterbank and co- authors described the use of diaryl-phos- the serendipitous discovery of the hydro- workers described new SPO-gold(i) com- phite pre-ligands, so-called HASPOs (het- lysis of phosphoramidite ligands such as plexes and used them as catalysts in cy- cloisomerisation of enynes and hydroxy- and methoxycyclisation reactions.[11] The [R P(OH)AuCl]complexesweregenerated 2 simply by mixing dimethylsulfide gold(i) with the desired SPO which coordinates the metal centre through its phosphorus atom. The gold complex 43 displayed an interesting catalytic activity in the cyclo- isomerisation of several enynes 44 (X = C(CO Me) ) giving good yields using mild 2 2 reaction conditions (Scheme 16). Interest- Scheme 16. Cycloisomerisation of enyne catalysed by SPO-Au complexes. The French connecTion CHIMIA 2016, 70, No. 1/2 15

DpenPhos into its corresponding second- Fig. 7. Selected chiral ary phosphine oxide form.[70] This latest SPO pre-ligands. SPO-like (S,S)-47 pre-ligand and others were then applied successfully to the en- antioselective rhodium-catalysed hydro- genation of α-ethylphosphonic acids 48. The authors reported both high reactivity (0.01% < catalyst) and enantioselectivity up to 98–99% for this SPO/Rh(i) combina- tion (Scheme 17).[70] This transformation occurs under relatively mild conditions at room temperature using only 10 bars Scheme 17. of hydrogen producing product 49. X-ray Hydrogenation of characterisation of the pre-catalyst showed α-ethylphosphonic acids catalysed by a pseudo ‘bidentate’ coordination of two SPO-Rh(i) complexes. SPO molecules arranged in a square pla- nar fashion around the rhodium centre. The authors highlighted the fact that H-bonding could be involved in the catalysis as poor conversions were observed in protic sol- bonds between two SPO units, was devel- activity; for example cyclohexanone needs vents. Finally this method was applied to oped by Castro et al. to construct unique a 1:10 ratio to achieve 92% conversion (R –1 the synthesis of fosmidomycin analogues systems for rhodium(iii)-catalysed trans- = Ph; TOF = 1825 h ). More fascinating [75] (malaria parasite inhibitor).[70] fer hydrogenation of ketones. By sim- is the result obtained using the BINOL- ply mixing RhCl and the desired SPO in derived chiral SPO 59 ligand (Fig. 9) that In an extension of simple and direct 3 synthesis of valuable bioactive build- , the authors generated a neutral catalysed the asymmetric transfer hydro- ing blocks, the same authors described a complex 58 (Fig. 9) that forms three in- genation of acetophenone with 89% ee, SPO-rhodium(i) catalysed hydrogenation tramolecular hydrogen bonds that do not employing isopropanol as the hydrogen of challenging β,β-diarylacrylic acids[71] interact with adjacent molecules. This re- donor. The positive linear effect observed 50 and α-CF 52 and β-CF 53 substitut- 3 3 ed acrylic acids into corresponding β,β- diarylpropionic acids[72] 51 and trifluo- romethylated propanoic acids 54 and 55 respectively (Scheme 18). In all cases they observed a synergistic effect[73] combining achiral triphenyl phosphine and chiral SPO 56 resulting in a significant enhancement of both conversion (15% to 99%) and en- antioselectivity (38% to 95%) compared with SPO used as the only ligand. This catalytic combination provides a straight- forward methodology to diverse optically active substituted acid derivatives with bio- logical importance. Scheme 18. Hydrogenation of α,β-CF acrylic acids catalysed by SPO-Rh(i) complexes. More recently, Han and co-workers 3 investigated other types of phosphorus li- gands that are able to generate the pseudo by the authors confirmed the preference of ‘bidentate’ coordination mode.[74] A se- a pseudo ‘bidentate’ chelating mode of the ries of chiral phosphinate and phosphite SPOs. were then explored in the asymmetric hy- drogenation of α-acetamidovinnamates. Hydrolysis of - and The easily accessible (R )-menthyl ben- Amine-boranes P zylphosphinate ligand 57 (Fig. 8) combined Due to its high hydrogen capacity (19.6 with cationic rhodium(i) salts produced the wt%) ammonia borane (AB, NH -BH ) 3 3 desired hydrogenated compound in high is considered an attractive candidate for yield (100%) and high enantiopurity (up chemical hydrogen storage applications. to 99%). Interestingly the chiral secondary Fig. 8. Chiral ‘SPO’ derived from . The dehydrocoupling of such a substrate phosphine oxide ligand (R1 = Ph; R2 = t- catalysed by transition metals[76] recently Bu) which produces the best asymmetric sults in the adoption of both pseudo ‘bi- became a field of intensive investigation induction only gave 12% ee. These results dentate’ or ‘tridentate’ coordination mode to efficiently produce hydrogen.[77] One of show the potential and the complementar- for the ensuing chelates. The best catalytic the limitations of this method is that transi- ity of H-phosphinate as an efficient chiral combination with a metal to ligand ratio of tion metal catalysts often require an inert ligand for homogenous catalysis. 1:10 for cyclohexanone (92% conv.; TOF atmosphere. Currently the most effective = 1825 h–1) and 1:3 for acetophenone (18% catalysts for this transformation are rho- Hydrogen Transfer conv.; TOF = 353 h–1) was established us- dium nanoparticles that allow the reaction The self-assembly of SPOs into anionic ing 0.5 mM of rhodium. to proceed at 25 °C in air.[78] Interestingly, bidentate chelates, thanks to their ability The metal to R P(H)O ligand ratio was only few metal catalysts are reported to be 2 to form strong intramolecular hydrogen found to be crucial to obtain good catalytic effective under homogenous conditions. 16 CHIMIA 2016, 70, No. 1/2 The French connecTion

Garralda et al. showed that ligands such Fig. 9. Chiral ‘SPO’ as o-(diphenylphosphino)benzaldehyde derived from binaph- can stabilise the hydridoirida-β-diketones thol. complex 60 by strong O---H---O intramo- lecular hydrogen bonds.[79] In this context, they prepared a rhodium(iii) complex 60 bound by both a phosphine-acyl and a SPO ligand. The phosphinous acid provides the hydrogen atom to form the expected pseudo ‘biden- tate’ coordination mode by intramolecular hydrogen bonding. These rhodium com- plexes were evaluated as catalysts for the hydrolytic release of hydrogen from am- monia- (AB) or tertbutylamine- (TBAB) and dimethylamine-borane (DMAB) sub- strates.[80] The reaction occurs under mild conditions: 40 °C in THF/H2O using only 0.5 mol% of catalyst generating more than

2.5 equiv. of hydrogen (Scheme 19). Neu- Scheme 19. Hydrolysis of amine-boranes catalysed by SPO-Rh(iii) complexes. tral complex 60 displayed better activities than the cationic complex and can be re- used at least six times reproducing the ki- derivatives 63 with terminal allenes that impact on the ee values (up to 91%). The netic profile. The recycling of the catalyst give access to enantioenriched secondary allene partner was then evaluated; once is easily achieved by simply reusing the 64 and tertiary 65 allylic pyrazoles.[82] again the N1 product was obtained in good leftover solution of a completed reaction JoSPOphos was found to efficiently yield and high enantiopurity starting either as the solvent for the next. catalyse this asymmetric transformation from mono- and 1,1-disubstituted allenes. giving both high regio- and enantioselec- The system was also flexible with various N-Coupling of Benzotriazoles tivity. The authors show that the presence functional groups such as phthalimide, In 2014, Breit and co-workers reported of pyridinium p-toluenesulfonate (PPTS) thioether, ether, silyl-ether and alcohol. the rhodium-catalysed N1-selective cou- is crucial to achieve high regioselectivity Substitution at the C(3)- or C(5)-posi- pling of benzotriazoles with allenes.[81] providing the N1 product (Scheme 21). A tion concomitant to the C(4) position ap- Due to the equilibrium between the N1 wide range of substituted 4-pyrazoles 64 pears to slightly decrease the regioselec- and N2 tautomers for 1,2,3-triazoles and were tested giving the corresponding allyl- tivity in some cases but enantioselectivity especially for benzotriazole, a mixture ic pyrazoles in good to excellent yields and remains high in all cases (Scheme 22).[82] of N1 and N 2-subsituted benzotriazole enantioselectivities (Scheme 21).[82] Halo- Globally, electron-rich pyrazoles led to the products is often obtained. An exception- gen, alkyl, fluoroalkyl, phenyl groups at desired N1 products 67 in excellent yields ally high N1 selectivity (product 62) was the 4-position and unsubstituted pyrazoles with high enantiomeric excess, but with a achieved by using a catalytic combination were well tolerated. However, for electron- slightly lower regioselectivity compared of rhodium(i)/JoSPOphos (Scheme 20). withdrawing groups a slight decrease in the to electron-poor ring systems that provide The catalytic system was tolerant toward reactivity was observed without a negative both high regio- and enantioselectivities. modifications on the benzotriazole back- bone and to several allene substrates bear- ing protected hydroxyl and phthalimide groups for example. In addition, this SPO pre-ligand displayed a promising asym- metric induction of 46% ee for the corre- sponding cyclohexylallene adduct.

Asymmetric N-Coupling of Pyrazoles Very recently, the same author de- scribed a SPO-rhodium-catalysed asym- metric N-selective coupling of pyrazole Scheme 20. N-coupling of benzotriazoles catalysed by SPO-Rh(i) complexes.

Scheme 21. Enantioselective addition of symmetric 4-substituted pyrazole derivatives to substituted terminal allenes. The French connecTion CHIMIA 2016, 70, No. 1/2 17

Bimetallic Catalysis

As mentioned in the introduction, SPOs ligands are able to coordinate two different metal centres through both its phosphorus and oxygen atoms. Generally, SPOs react with soft metal ions giving the phosphinous acid complex, which in the presence of a base, react with other metal species (often hard ions) either generating type (h) metallacycle or type (e) bridged complexes (see Scheme 2).[6a,7a,83] Several bifunctional metal complexes have been prepared with different combinations such Scheme 22. Enantioselective addition of symmetric 3,4- or 3,5-substituted pyrazoles to substi- tuted terminal allenes. as Pd-Ti,[16] Pt-Ir and Pt-Rh,[84] Fe-Rh,[85] and others. Curiously the use of such het- erobimetallic complexes that would have 0.5 mol% without affecting the selectiv- control over a stereogenic centre present some cooperative effect in catalysis is ity of the reaction (96.5:3.5). Several for- in the substrate 76 (Scheme 25).[88] Indeed largely uninvestigated in this area.[86] Only mamide substrates 69 and 74 were then the matched case produces a high diaste- the nickel/magnesium bimetallic combi- examined, varying both the substituent on reoselectivity (16:1) and the mismatched nation has been suggested with second- the atom and the number of allylic one a moderate but inverse dr (3:1). The ary phosphine oxides for Kumada-Corriu chains (see Scheme 24).[88] The observed chelation of both metals is suggested to cross-couplings of aryl fluorides.[87] selectivity is not related to the substituent facilitate the C–H activation process by Very recently, the group of Cramer on the nitrogen atom as different groups placing the low-valent nickel metal centre developed an asymmetric hydrocarbonyl- such as alkyl, aryl, benzyl or ester generate in close proximity to the . The ation of homoallylic 69 cata- similar levels of selectivity. Bis-allylated tight geometry of the intermediate might lysed by a hetero-bimetallic SPOs nickel/ formamides 74 lead to the desired lactam be responsible for the good enantiodis- alumina enantiopure complex.[88] First adducts 75 in high yield and enantioselec- crimination in the hydrometalation step. of all, the phosphinous acid form reacts tivities. Interestingly the good diastereose- with Al(Me) through its hydroxyl group, lectivity obtained in this case, point to a 3 releasing a of and pro- good enantiocontrol in the selection of one Asymmetric Catalysis vides an O-coordinated complex type (d). of the two enantiotopic allyl groups. For the formation of catalytically active Finally, the authors show that this cata- The desymmetrisation of the SPOs aluminium-phosphinous acid-Ni complex, lytic system could dictate its own enantio- (different R groups) makes the phospho-

Scheme 23. Enantioselective hydrocarbamoylation catalysed by SPO-Ni/Al complexes. the use of PPh was found essential to dis- Scheme 24. 3 place the cycloctadiene from the nickel Enantioselective to allow the coordination of the SPO pre- hydrocarbamoylation ligand. Enantioenriched SPO 70 derived catalysed by SPO- from phospholane, under mild conditions Ni/Al complexes. (toluene, 40 °C, 12 h), provides γ-lactam 71 in good yield and interesting enantio- meric ratio (er: 14:86, Scheme 23). Inspired by the seminal work of Hama- da[89] in the use of chiral diaminophosphine oxides (DIAPHOXs) in catalytic transfor- mation, the authors developed a new chi- ral HASPO catalyst 72 from the diamine backbone.[88] Using this SPO DI- APHOXs ligand 72, the authors were able to reduce the catalyst loading from 10 to Scheme 25. Matched/mismatched stereocontrol of chiral SPO 72. 18 CHIMIA 2016, 70, No. 1/2 The French connecTion

rus atom stereogenic, providing a door- way to asymmetric catalysis. Fortunately, the stereochemistry at the phosphorus atom is retained throughout the equilib- rium between the secondary phosphine oxide and the phosphinous acid. These chiral SPO ligands[90] are accessible via several methods such as optical resolu- Fig. 10. Chiral SPOs, DIAPHOXs and HASPOs ligands. tion,[91] chiral chromatography,[68] nucleo- philic substitution on chiral (–)-menthyl phenylphosphinate,[22b,c] and dynamic displayed as well this talent and showed [11] F. Schroeder, C. Tugny, E. Salanouve, H. resolution[92] for example.[24,93] These [74] Clavier, L. Giordano, D. Moraleda, Y. Gimbert, characteristics indeed make them very at- promising reactivities. Complementary to this, heteroatom-substituted phosphine V. Mouries-Mansuy, J.-P. Goddard, L. tractive ligands in asymmetric catalysis. Fensterbank, Organometallics 2014, 33, 4051. Since the first report by Minnaard, Feringa oxides (HASPO) and diaminophosphine [12] L. V. Graux, M. Giorgi, G. Buono, H. Clavier, oxides (DIAPHOXs) were also found to be Organometallics 2015, 34, 1864. and de Vries on asymmetric hydrogena- [13] N. W. Alcock, P. Bergamini, T. M. Gomes- tion catalysed by the P-chiral SPOs (40, very interesting ligands displaying high ac- tivity.[98] Due to their original catalytic ac- Carniero, R. D. Jackson, J. Nicholls, A. G. 78)/iridium system,[68] other asymmetric Orpen, P. G. Pringle, S. Sostero, O. Traverso, J. transformations have been successfully ex- tivity, all these pre-ligands should be used Chem. Soc., Chem. Commun. 1990, 980. to reinvestigate classic metal-catalysed plored.[24a,94] Enantiopure SPOs were suc- [14] A. M. Z. Slawin, M. Wainwright, J. Derek reactions. Finally, the use of simple com- Woollins, New J. Chem. 2000, 24, 69. cessfully used in rhodium-catalysed alkene [15] T. Achard, L. Giordano,A. Tenaglia,Y. Gimbert, hydrogenation (SPO-40, 58, 59, 81, Jo- binations of various groups (alkyl, aryl, alkoxy or aryloxy) bound to the phospho- G. Buono, Organometallics 2010, 29, 3936. SPOphos, TerSPOphos; Scheme 17 and [16] T. Mizuta, C. Miyaji, T. Katayama, J.-i. Ushio, rus atom would bring more diversity and Scheme 18, Fig. 10),[67,69,74] SPO-80/pal- K. Kubo, K. Miyoshi, Organometallics 2009, 28, 539. [89a,95] quickly generate a large library of ligands. ladium-catalysed allylic substitution, [17] A. Christiansen, D. Selent, A. Spannenberg, SPO-25/palladium-catalysed allylic ami- Promising but still limited results were ob- tained in asymmetric catalysis using chiral W. Baumann, R. Franke, A. Boerner, nation[89b,96] and SPO-78/palladium-catal- Organometallics 2010, 29, 3139. phosphine oxides. However synthesis of ysed [2+1] cycloaddition.[97] Finally, Rh- [18] D. Martin, D. Moraleda, T. Achard, L. Giordano, JoSPOphos complexes displayed a high enantioenriched SPOs is currently a grow- G. Buono, Chem. Eur. J. 2011, 17, 12729. ing field of investigation. [19] a) J. Bigeault, L. Giordano, I. De Riggi, Y. regio- and enantioselectivity for N1-sub- Gimbert, G. Buono, Org. Lett. 2007, 9, 3567; b) stitution for both benzotriazoles and pyr- Acknowledgements L. Ackermann, H. K. Potukuchi, A. R. Kapdi, azoles substrates.[81,82] C. Schulzke, Chem. Eur. J. 2010, 16, 3300. I would like to thank particularly Prof. [20] a) P. van Leeuwen, C. F. Roobeek, R. L. Wife, Gérard Buono, Dr. Alphonse Tenaglia, Dr. J. H. G. Frijns, J. Chem. Soc. Chem. Com. 1986, Hervé Clavier, and Dr. Laurent Giordano 31; b) P. van Leeuwen, C. F. Roobeek, J. H. G. Conclusion (IsM2-Chirosciences, UMR 7313, Marseille) Frijns, A. G. Orpen, Organometallics 1990, 9, who gave me the opportunity to discover such 1211. Due notably to their air and moisture a fascinating field. [21] L. Ackermann, Synthesis 2006, 1557. stability, secondary phosphine oxides [22] a) T. L. Emmick, R. L. Letsinger, J. Am. Chem. Received: March 31, 2015 Soc. 1968, 90, 3459; b) A. Leyris, J. Bigeault, (SPOs) represent a new and attractive al- D. Nuel, L. Giordano, G. Buono, Tetrahedron ternative to classic phosphine ligands. Lett. 2007, 48, 5247; c) Q. Xu, C.-Q. Zhao, L.- These pre-ligands have been success- [1] ‘A Guide to Organophosphorus Chemistry’, B. Han, J. Am. Chem. Soc. 2008, 130, 12648; fully employed in several metal-catalysed Ed. L. D. Quin, John Wiley & Sons, New York, d) A. J. Bloomfield, J. M. Qian, S. B. Herzon, 2000. Organometallics 2010, 29, 4193. reactions. Clearly the results indicated [2] ‘Phosphorus Ligands in Asymmetric Catalysis [23] a) N. V. Dubrovina, A. Borner, Angew. Chem. that SPOs are more than just ‘phosphine – Synthesis and Application’, Ed. A. Börner, Int. Ed. 2004, 43, 5883; b) L. Ackermann, mimics’ possessing their own special Wiley-VCH: Weinheim, 2008. Synlett 2007, 507. characteristics and promoting original [3] a) W. J. Tang, X. M. Zhang, Chem. Rev. 2003, [24] a) T. M. Shaikh, C.-M. Weng, F.-E. Hong, reactivity. The easy modularity of both 103, 3029; b) K. V. L. Crepy, T. Imamoto, New Coord. Chem. Rev. 2012, 256, 771; b) T. Aspects in Phosphorus Chemistry Iii 2003, 229, Nemoto, Y. Hamada, Tetrahedron 2011, 67, steric and electronic properties of these li- 1. 667. gands makes them attractive for catalysts [4] M. Ohff, J. Holz, M. Quirmbach, A. Borner, [25] a) G. Y. Li, G. Zheng, A. F. Noonan, J. Org. and new reactivity screening. More pre- Synthesis 1998, 1391. Chem. 2001, 66, 8677; b) G. Y. Li, Angew. cisely, the electronic properties of SPOs [5] a) J. Chatt, B. T. Heaton, J. Chem. Soc. A 1968, Chem. 2001, 113, 1561; c) G. Y. Li, P. J. Fagan, 2745; b) A. Christiansen, C. Li, M. Garland, [26] C. Wolf, R. Lerebours, J. Org. Chem. 2003, 68, shows an extraordinary range, going from D. Selent, R. Ludwig, A. Spannenberg, W. 7077. π-accepting phosphite-like compounds, to Baumann, R. Franke, A. Boerner, Eur. J. Org. [27] C. Wolf, R. Lerebours, J. Org. Chem. 2003, 68, extremely electron-donating ligands out- Chem. 2010, 2733. 7551. classing N-heterocyclic carbenes.[18] Once [6] a) D. M. Roundhill, R. P. Sperline, W. B. [28] C. Wolf, R. Lerebours, Org. Biomol. Chem. coordinated to the metal centre through the Beaulieu, Coord. Chem. Rev. 1978, 26, 263; b) 2004, 2, 2161. G. Manca, M. Caporali, A. Ienco, M. Peruzzini, [29] a) C. Wolf, R. Lerebours, E. H. Tanzini, phosphorus atom, the resulting OH group C. Mealli, J. Organomet. Chem. 2014, 760, 177. Synthesis 2003, 2069; b) C. Wolf, R. Lerebours, may act as an acid, as a directing group [7] a) B. Walther, Coord. Chem. Rev. 1984, 60, 67; Org. Lett. 2004, 6, 1147. and as a pendant base responsible for the b) T. Appleby, J. D. Woollins, Coord. Chem. [30] C. Wolf, H. Xu, J. Org. Chem. 2008, 73, 162. interesting activity of SPO-metal com- Rev. 2002, 235, 121. [31] H. Xu, K. Ekoue-Kovi, C. Wolf, J. Org. Chem. [8] D. Magiera, A. Szmigielska, K. M. 2008, 73, 7638. plexes. These ligand–metal cooperative Pietrusiewicz, H. Duddeck, Chirality 2004, 16, [32] a) W. Yang, Y. Wang, J. R. Corte, Org. Lett. effects were found crucial to stabilize ru- 57. 2003, 5, 3131; b) R. R. Poondra, P. M. Fischer, thenium and gold-nanoparticles providing [9] J.-N. Li, L. Liu, Y. Fu, Q.-X. Guo, Tetrahedron N. J. Turner, J. Org. Chem. 2004, 69, 6920; c) S. highly active catalysts. The ability to form 2006, 62, 4453. P. Khanapure, D. S. Garvey, Tetrahedron Lett. pseudo ‘bidentate’ anionic ligands is not [10] The crystallographic data for complexes have 2004, 45, 5283; d) P. Cao, J. Qu, G. Burton, R. been obtained free of charge via the Cambridge A. Rivero, J. Org. Chem. 2008, 73, 7204; e) K. limited solely to SPOs. Phosphinate and Crystallographic Data Centre and cif files were L. Billingsley, S. L. Buchwald, Angew. Chem. phosphite, often synthetic intermediates, reprocessed using Ortep software. Int. Ed. 2008, 47, 4695; f) D. X. Yang, S. L. The French connecTion CHIMIA 2016, 70, No. 1/2 19

Colletti, K. Wu, M. Song, G. Y. Li, H. C. Shen, [57] E. Tomas-Mendivil, F. J. Suarez, J. Diez, V. [80] V. San Nacianceno, L. Ibarlucea, C. Org. Lett. 2009, 11, 381; g) Y.-Y. Chang, F.-E. Cadierno, Chem. Commun. 2014, 50, 9661. Mendicute-Fierro, A. Rodriguez-Dieguez, J. Hong, Tetrahedron 2013, 69, 2327. [58] E. Tomas-Mendivil, L. Menendez-Rodriguez, M. Seco, I. Zumeta, C. Ubide, M. A. Garralda, [33] K. Ekoue-Kovi, H. Xu, C. Wolf, Tetrahedron J. Francos, P. Crochet, V. Cadierno, RSC Adv. Organometallics 2014, 33, 6044. Lett. 2008, 49, 5773. 2014, 4, 63466. [81] K. Xu, N. Thieme, B. Breit, Angew. Chem. Int. [34] R. Lerebours, A. Camacho-Soto, C. Wolf, J. [59] F. Wang, W. E. Buhro, J. Am. Chem. Soc. 2012, Ed. 2014, 53, 7268. Org. Chem. 2005, 70, 8601. 134, 5369. [82] A. M. Haydl, K. Xu, B. Breit, Angew. Chem. [35] Z. Zhang, Z. Hu, Z. Yu, P. Lei, H. Chi, Y. Wang, [60] E. Rafter, T. Gutmann, F. Loew, G. Buntkowsky, Int. Ed. 2015, 54, 7149. R. He, Tetrahedron Lett. 2007, 48, 2415. K. Philippot, B. Chaudret, P. W. N. M. van [83] T. Appleby, J. D. Woollins, Coord. Chem. Rev. [36] R. Lerebours, C. Wolf, Org. Lett. 2007, 9, 2737. Leeuwen, Catal. Sci. Technol. 2013, 3, 595. 2002, 235, 121. [37] L. Ackermann, R. Vicente, N. Hofmann, Org. [61] a) R. Torres-Lubián, M. J. Rosales-Hoz, A. [84] L. R. Falvello, J. Forniés, A. Martín, J. Gómez, Lett. 2009, 11, 4274. M. Arif, R. D. Ernst, M. A. Paz-Sandoval, E. Lalinde, M. T. Moreno, J. Sacristán, Inorg. [38] R. Lerebours, C. Wolf, J. Am. Chem. Soc. 2006, J. Organomet. Chem. 1999, 585, 68; b) Chem. 1999, 38, 3116. 128, 13052. C. J. den Reijer, M. Wörle, P. S. Pregosin, [85] H. Brunner, R. Eder, B. Hammer, U. Klement, [39] a) S. Wang, Z. Zhang, Z. Hu, Y. Wang, P. Lei, Organometallics 2000, 19, 309. J. Organomet. Chem. 1990, 394, 555. H. Chi, J. Environ. Sci.-China 2009, 21 Suppl [62] a) H. Schmidbaur, A. A. M. Aly, Angew. Chem. [86] a) R. M. Bullock, C. P. Casey, Acc. Chem. Res. 1, S124; b) Z. Zhang, Z. Hu, S. Wang, P. Lei, Int. Ed. 1980, 19, 71; b) C. Hollatz, A. Schier, 1987, 20, 167; b) N. Wheatley, P. Kalck, Chem. H. Chi, R. He, J. Environ. Sci.-China 2009, 21 H. Schmidbaur, J. Am. Chem. Soc. 1997, 119, Rev. 1999, 99, 3379; c) B. G. Cooper, J. W. Suppl 1, S65; c) B. Li, C. Wang, G. Chen, Z. 8115; c) C. Hollatz, A. Schier, J. Riede, H. Napoline, C. M. Thomas, Catal. Rev. 2012, 54, Zhang, J. Environ. Sci-China 2013, 25, 1083. Schmidbaur, J. Chem. Soc., Dalton Trans. 1999, 1. [40] D.-F. Hu, C.-M. Weng, F.-E. Hong, 111; d) C. Hollatz, A. Schier, H. Schmidbaur, [87] a) L. Ackermann, A. Althammer, Chemie in Organometallics 2011, 30, 1139. Inorg. Chim. Acta 2000, 300, 191. unserer Zeit 2009, 43, 74; b) Z. Jin, Y.-J. Li, [41] T. M. Shaikh, F. E. Hong, Beilstein J. Org. [63] I. Cano, A. M. Chapman, A. Urakawa, P. W. N. Y.-Q. Ma, L.-L. Qiu, J.-X. Fang, Chem. Eur. J. Chem. 2013, 9, 1578. M. van Leeuwen, J. Am. Chem. Soc. 2014, 136, 2012, 18, 446. [42] Y.-C. Chang, W.-C. Chang, C.-Y. Hu, F.-E. 2520. [88] P. A. Donets, N. Cramer, J. Am. Chem. Soc. Hong, Organometallics 2014, 33, 3523. [64] M. Matsumoto, M. Tamura, J. Mol. Catal. 1983, 2013, 135, 11772. [43] J. Bigeault, L. Giordano, G. Buono, Angew. 19, 365. [89] a) T. Nemoto, T. Matsumoto, T. Masuda, T. Chem. Int. Ed. 2005, 44, 4753. [65] A. Christiansen, C. Li, M. Garland, D. Selent, R. Hitomi, K. Hatano, Y. Hamada, J. Am. Chem. [44] J. Bigeault, I. de Riggi,Y. Gimbert, L. Giordano, Ludwig, R. Franke, A. Boerner, Chemcatchem Soc. 2004, 126, 3690; b) T. Nemoto, T. Masuda, G. Buono, Synlett 2008, 1071. 2010, 2, 1278. Y. Akimoto, T. Fukuyama, Y. Hamada, Org. [45] a) T. Ghaffar, A. W. Parkins, Tetrahedron Lett. [66] A. Christiansen, D. Selent, A. Spannenberg, M. Lett. 2005, 7, 4447. 1995, 36, 8657; b) T. Ghaffar, A. W. Parkins, J. Koeckerling, H. Reinke, W. Baumann, H. Jiao, [90] L. Ackermann, ‘Chiral Secondary Phosphine Mol. Catal. A: Chem. 2000, 160, 249. R. Franke, A. Boerner, Chem. Eur. J 2011, 17, Oxides and Heteroatom-Substituted Secondary [46] a) T. J. Ahmed, S. M. M. Knapp, D. R. Tyler, 2120. Phosphine Oxides as Preligands’ in ‘Phosphorus Coord. Chem. Rev. 2011, 255, 949; b) R. [67] X. B. Jiang, M. van den Berg, A. J. Minnaard, Ligands in Asymmetric Catalysis’, Vol. I–III, Garcia-Alvarez, P. Crochet, V. Cadierno, Green B. L. Feringa, J. G. de Vries, Tetrahedron- Ed.: A. Boerner, Wiley-VCH, Weinheim, 2008, Chem. 2013, 15, 46. Asymmetry 2004, 15, 2223. pp. 831–847. [47] W. F. Yates, R. L. Heider, J. Am. Chem. Soc. [68] X. B. Jiang, A. J. Minnaard, B. Hessen, B. L. [91] a) J. Drabowicz, P. Lyzwa, J. Omelanczuk, K. M. 1952, 74, 4153. Feringa, A. L. L. Duchateau, J. G. O. Andrien, Pietrusiewicz, M. Mikolajczyk, Tetrahedron- [48] S. M. M. Knapp, T. J. Sherbow, J. J. Juliette, D. J. A. F. Boogers, J. G. de Vries, Org. Lett. 2003, Asymmetry 1999, 10, 2757; b) R. K. Haynes, T. R. Tyler, Organometallics 2012, 31, 2941. 5, 1503. L. Au-Yeung, W. K. Chan, W. L. Lam, Z. Y. Li, [49] a) T. J. Ahmed, B. R. Fox, S. M. M. Knapp, R. [69] H. Landert, F. Spindler, A. Wyss, H.-U. Blaser, L. L. Yeung, A. S. C. Chan, P. Li, M. Koen, C. B. Yelle, J. J. Juliette, D. R. Tyler, Inorg. Chem. B. Pugin, Y. Ribourduoille, B. Gschwend, B. R. Mitchell, S. C. Vonwiller, Eur. J. Org. Chem. 2009, 48, 7828; b) S. Knapp, T. Sherbow, T. Ramalingam, A. Pfaltz, Angew. Chem. Int. Ed. 2000, 3205; c) A. Leyris, D. Nuel, L. Giordano, Ahmed, I. Thiel, L. Zakharov, J. J. Juliette, D. 2010, 49, 6873. M. Achard, G. Buono, Tetrahedron Lett. 2005, Tyler, J. Inorg. Organomet. Polym. Mater. 2014, [70] K. Dong, Z. Wang, K. Ding, J. Am. Chem. Soc. 46, 8677; d) J. Holt, A. M. Maj, E. P. Schudde, 24, 145. 2012, 134, 12474. K. M. Pietrusiewicz, L. Sieron, W. Wieczorek, [50] T. Achard, A. Lepronier, Y. Gimbert, H. Clavier, [71] Y. Li, K. Dong, Z. Wang, K. Ding, Angew. T. Jerphagnon, I. W. C. E. Arends, U. Hanefeld, L. Giordano, A. Tenaglia, G. Buono, Angew. Chem. Int. Ed. 2013, 52, 6748. A. J. Minnaard, Synthesis 2009, 2061; e) M. Chem. Int. Ed. 2011, 50, 3552. [72] K. Dong, Y. Li, Z. Wang, K. Ding, Angew. Stankevic, K. M. Pietrusiewicz, J. Org. Chem. [51] L. Ackermann, Org. Lett. 2005, 7, 3123. Chem. Int. Ed. 2013, 52, 14191. 2007, 72, 816. [52] a) L. Ackermann, A. Althammer, R. Born, [73] Y. T. Fan, M. Cong, L. Peng, Chem. Eur. J 2014, [92] F. A. Kortmann, M.-C. Chang, E. Otten, E. P. A. Angew. Chem. Int. Ed. 2006, 45, 2619; b) L. 20, 2698. Couzijn, M. Lutz, A. J. Minnaard, Chem. Sci. Ackermann, Isr. J. Chem. 2010, 50, 652; c) L. [74] X.-B. Wang, M. Goto, L.-B. Han, Chem. Eur. J 2014, 5, 1322. Ackermann, A. V. Lygin, Org. Lett. 2011, 13, 2014, 20, 3631. [93] O. I. Kolodiazhnyi, Tetrahedron: Asymmetry 3332; d) L. Ackermann, M. Mulzer, Org. Lett. [75] P. M. Castro, H. Gulyas, J. Benet-Buchholz, C. 2012, 23, 1. 2008, 10, 5043. Bo, Z. Freixa, P. W. N. M. van Leeuwen, Catal. [94] For a recent review see: ref. [24a]. [53] For reviews see: a) L. Ackermann, Acc. Chem. Sci. Technol. 2011, 1, 401. [95] a) W.-M. Dai, K. K. Y. Yeung, W. H. Leung, R. Res. 2014, 47, 281; b) L. Ackermann, Org. [76] A. Staubitz, A. P. M. Robertson, I. Manners, K. Haynes, Tetrahedron: Asymmetry 2003, 14, Process Res. Dev. 2015, 19, 260. Chem. Rev. 2010, 110, 4079. 2821; b) T. Nemoto, T. Sakamoto, T. Fukuyama, [54] L. Ackermann, E. Diers, A. Manvar, Org. Lett. [77] a) H. W. Langmi, G. S. McGrady, Coord. Chem. Y. Hamada, Tetrahedron Lett. 2007, 48, 4977. 2012, 14, 1154. Rev. 2007, 251, 925; b) T. Umegaki, J.-M. Yan, [96] T. Nemoto, T. Fukuyama, E. Yamamoto, S. [55] R. Garcia-Alvarez, J. Francos, E. Tomas- X.-B. Zhang, H. Shioyama, N. Kuriyama, Q. Tamura, T. Fukuda, T. Matsumoto, Y. Akimoto, Mendivil, P. Crochet, V. Cadierno, J. Xu, Int. J. Hydrogen Energ. 2009, 34, 2303. Y. Hamada, Org. Lett. 2007, 9, 927. Organomet. Chem. 2014, 771, 93. [78] M. Fetz, R. Gerber, O. Blacque, C. M. Frech, [97] D. Gatineau, D. Moraleda, J.-V. Naubron, T. [56] S. M. M. Knapp, T. J. Sherbow, R. B. Yelle, J. J. Chem. Eur. J. 2011, 17, 4732. Buergi, L. Giordano, G. Buono, Tetrahedron- Juliette, D. R. Tyler, Organometallics 2013, 32, [79] M. A. Garralda, R. Hernández, L. Ibarlucea, E. Asymmetry 2009, 20, 1912. 3744. Pinilla, M. R. Torres, Organometallics 2003, [98] L. Ackermann, Isr. J. Chem. 2010, 50, 652. 22, 3600.