ACS Publications. Most Trusted. Most Cited. Most Read
H-Bonded Counterion-Directed Enantioselective Au(I) Catalysis
My Activity

Figure 1Loading Img
  • Open Access
Article

H-Bonded Counterion-Directed Enantioselective Au(I) Catalysis
Click to copy article linkArticle link copied!

  • Allegra Franchino
    Allegra Franchino
    Institute of Chemical Research of Catalonia (ICIQ), Barcelona Institute of Science and Technology, Av. Països Catalans 16, 43007 Tarragona, Spain
  • Àlex Martí
    Àlex Martí
    Institute of Chemical Research of Catalonia (ICIQ), Barcelona Institute of Science and Technology, Av. Països Catalans 16, 43007 Tarragona, Spain
    Departament de Química Orgànica i Analítica, Universitat Rovira i Virgili, C/Marcel·lí Domingo s/n, 43007 Tarragona, Spain
    More by Àlex Martí
  • Antonio M. Echavarren*
    Antonio M. Echavarren
    Institute of Chemical Research of Catalonia (ICIQ), Barcelona Institute of Science and Technology, Av. Països Catalans 16, 43007 Tarragona, Spain
    Departament de Química Orgànica i Analítica, Universitat Rovira i Virgili, C/Marcel·lí Domingo s/n, 43007 Tarragona, Spain
    *[email protected]
Open PDFSupporting Information (1)

Journal of the American Chemical Society

Cite this: J. Am. Chem. Soc. 2022, 144, 8, 3497–3509
Click to copy citationCitation copied!
https://doi.org/10.1021/jacs.1c11978
Published February 9, 2022

Copyright © 2022 The Authors. Published by American Chemical Society. This publication is licensed under

CC-BY-NC-ND 4.0 .

Abstract

Click to copy section linkSection link copied!

A new strategy for enantioselective transition-metal catalysis is presented, wherein a H-bond donor placed on the ligand of a cationic complex allows precise positioning of the chiral counteranion responsible for asymmetric induction. The successful implementation of this paradigm is demonstrated in 5-exo-dig and 6-endo-dig cyclizations of 1,6-enynes, combining an achiral phosphinourea Au(I) chloride complex with a BINOL-derived phosphoramidate Ag(I) salt and thus allowing the first general use of chiral anions in Au(I)-catalyzed reactions of challenging alkyne substrates. Experiments with modified complexes and anions, 1H NMR titrations, kinetic data, and studies of solvent and nonlinear effects substantiate the key H-bonding interaction at the heart of the catalytic system. This conceptually novel approach, which lies at the intersection of metal catalysis, H-bond organocatalysis, and asymmetric counterion-directed catalysis, provides a blueprint for the development of supramolecularly assembled chiral ligands for metal complexes.

This publication is licensed under

CC-BY-NC-ND 4.0 .
  • cc licence
  • by licence
  • nc licence
  • nd licence
Copyright © 2022 The Authors. Published by American Chemical Society

Note Added after ASAP Publication

This paper was published February 9, 2022. References 2a,b have been added and the paper was re-posted February 17, 2022.

Introduction

Click to copy section linkSection link copied!

Groundbreaking work by Toste and co-workers proved that it was possible to perform gold catalysis enantioselectively by using achiral ligands together with chiral anions, combining a dinuclear gold complex with a chiral silver phosphate to realize the asymmetric cyclization of allenes (Scheme 1A). (1) This contributed to the development of the field of “asymmetric-counteranion directed catalysis” (ACDC), (2) and chiral anions have since been employed as stereodirecting elements in other Au(I)-catalyzed transformations, in combination with both achiral (3) and chiral (4) gold complexes. However, nearly all examples of Au(I)-ACDC reported to date refer to reactions of prochiral allene substrates. (3,4) The only exceptions are the desymmetrization of diynes described by the group of Czekelius (Scheme 1B) (5) and cases of tandem Au/chiral acid catalysis wherein gold is generally not involved in the enantiodetermining Brønsted acid-catalyzed step. (6) To the best of our knowledge, all other attempts to leverage chiral counterions in more challenging asymmetric Au(I)-catalyzed reactions of alkynes, (7) which, unlike allenes, cannot be prochiral in themselves, were met with failure until now. In this regard, our group reported that mixing triphenylphosphinegold(I) chloride with BINOL-derived silver phosphates affords neutral gold complexes with an anionic phosphate ligand, which are not catalytically competent in enyne cycloisomerizations. (8) In fact, the counteranion has a profound impact on the reactivity and selectivity of all gold-catalyzed transformations, since properties such as its basicity, coordinating ability, and steric bulk influence the energy barriers for various elementary steps of the catalytic cycle. (9,10)

Scheme 1

Scheme 1. Asymmetric Counterion-Directed Gold Catalysis
A further layer of complexity is added when the anion is chiral, as its position with respect to the reaction center, key for a successful ACDC, depends on difficult-to-predict electrostatic interactions with the cationic complex/intermediate or on the presence of a protic group on the substrate. Therefore, in the vast majority of cases, the noncovalent interactions required for enantioinduction (11) are not built into the catalytic system a priori but rather selected during optimization by a lengthy trial-and-error approach and rationalized only aposteriori. (12) Alternatively, the chiral anion can be precisely positioned by a rigid covalent linkage to the ligand, as in the “tethered counterion-directed catalysis” strategy recently disclosed by Marinetti, Guinchard, and co-workers for the enantioselective Au(I)-catalyzed tandem cycloisomerization–nucleophile addition to 2-alkynyl enones. (13) However, despite its elegance, this latter approach becomes akin to using a chiral ligand, so it is devoid of the flexibility and combinatorial potential offered by the original two-component ACDC.
Herein we detail a conceptually novel “H-bonded counterion-directed catalysis” strategy that enables the first general use of chiral anions in gold(I)-catalyzed reactions of alkyne substrates, as opposed to well-established allenes (Scheme 1c). Various bi- and tricyclic structures were accessed in good to excellent yield and enantioselectivity from 1,6-enynes, via 5-exo-dig and 6-endo-dig cyclizations with or without external nucleophiles. (14) To overcome the reactivity (8) and enantiocontrol (15) challenges described above, we envisioned to prepare gold(I) complexes with bifunctional phosphine ligands (16) incorporating dual H-bond donor groups (17) such as (thio)ureas (18) and combine them with BINOL-derived chiral anions (19) (Scheme 2). The pendant H-bond donor would serve two purposes: (1) remove the chiral anionic ligand from the Au(I) coordination sphere, thus enabling substrate coordination and subsequent catalysis at the metal center; (2) place the chiral anion close to the substrate, ensuring good transmission of the stereochemical information.

Scheme 2

Scheme 2. Design for Asymmetric H-Bonded Counterion-Directed Au(I) Catalysis
The feasibility of ligand abstraction via H-bonding interactions was supported by our previous studies on the self-activation of Au(I) chloride complexes featuring PPh3-based phosphinosquaramides and phosphinoureas. (18b) Regarding stereochemical considerations, we chose Buchwald-type phosphines functionalized at the meta or para position of the distal aryl ring due to both their easy electronic tuning and attractive geometric features in the context of linear dicoordinated Au(I) complexes (Scheme 2). (16,20) In fact, the blocked rotation about the P–Caryl bond, enforced by the interlocking of the ortho-proton Ho between the two bulky alkyl groups, causes the P–Au–Cl and biphenyl axes to be parallel. This in turn projects the urea “anchor” for the chiral anion close to the substrate coordination site. This H-bonded ACDC approach features a modular and easy synthesis for both pieces of the catalytic system, which would click together based on predictable, well-precedented H-bond interactions between (thio)ureas and phosphate anions. (21) This strategy can thus benefit from all the typical advantages offered by supramolecularly assembled ligands for enantioselective metal catalysis, (22) most notably (i) simple preparation and tuning of the two components, which avoids long syntheses of new chiral ligands from scratch, and (ii) generation of a large library of catalysts through combinatorial and potentially automated methods, with the aim to speed up screening and optimization.

Results and Discussion

Click to copy section linkSection link copied!

Synthesis of the Components of the Catalytic System

The bifunctional phosphino(thio)urea Au(I) chloride complexes were prepared with a high-yielding and modular three-step sequence from bromoanilines A13, in turn accessible in one or two steps from commercial building blocks (Scheme 3, see Supporting Information for details). A well-scalable Pd-catalyzed P–C coupling (23) allowed the introduction of different phosphine substituents on the biphenyl scaffold without the need to protect the free amine group. In this sense, the use of potassium phosphate in place of stronger bases such as sodium tert-butoxide was instrumental in suppressing undesired Buchwald–Hartwig coupling between the aryl bromide and the NH2 group. Phosphinoanilines B15 were then reacted with the iso(thio)cyanate of choice for the late-stage introduction of the key H-bond donor on ligands L113. Finally, ligand exchange with [(Me2S)AuCl] afforded Au(I) chloride complexes Au113. Thus, all complexes were prepared in four to five total steps and 42–60% overall yields except for Au12 featuring an N-aryl-N′-alkylurea, obtained in 31% overall yield. Previously described complexes Au1416 with a triarylphosphine backbone were included for comparison. (18b)

Scheme 3

Scheme 3. Synthesis of Phosphino(thio)urea Au(I) Chloride Complexes Au113a

aSelected X-ray structures displayed with ORTEP ellipsoid at 50% probability level; solvent molecules and selected H atoms omitted for clarity.

Representative complexes were characterized also by single crystal X-ray diffraction (see Supporting Information for a complete overview). In the solid state, all novel phosphinourea Au(I) complexes display an intact Au–Cl bond with the P–Au–Cl axis nearly parallel to the biphenyl axis, as expected. The urea moiety generally adopts an anti,anti-conformation (Au1, Au6, Au10, and Au13) but displays the much rarer anti,syn one (24) in Au4 and Au8 and shows various degrees of out-of-plane twisting of the aryl substituents (compare Au1 and Au10 in Scheme 3). In the solid state, the urea group establishes H-bonds either with chloride ligands or with other urea units, depending on the ligand scaffold. As for phosphinothiourea Au(I) complexes, Au7 possessing a para-substituted ligand shows a classical [LAuCl] structure with intramolecular NH···Cl contacts. In the crystal structure of Au9, on the contrary, the S atom of the meta thiourea coordinates to gold with an almost linear P–Au–S axis, while the displaced chloride ion is stabilized by two H-bonds with the NH groups.
Regarding the other component of the catalytic system, we planned to introduce the chiral anion as a metal salt in order to simultaneously scavenge the chloride ligand from the Au(I) center. To this end, eight chiral salts (Ag16, Na6, and Cu6) were synthesized in 2–3 steps and 30–77% overall yield from commercially available (R)-configured binaphthols C14 (Scheme 4). After treatment of the latter with phosphoryl chloride, the desired sulfonamide was added to the reaction mixture affording phosphoramidates (25)D16. Deprotonation of these Brønsted acids with silver carbonate delivered chiral silver salts Ag16. Crystals of Ag6 grown in toluene/pentane reveal a dimeric structure where the anion behaves as a bidentate O,O-ligand through the phosphoryl and sulfonyl O atoms (Scheme 4). The Ag(I) centers are further stabilized by η2-interactions with the anthracenyl substituents, and one of them is also bound in a η12 fashion to a molecule of toluene.

Scheme 4

Scheme 4. Synthesis of Ag(I), Na(I), and Cu(II) Chiral Saltsa

aX-ray structure displayed with ORTEP ellipsoid at 50% probability level; binaphthol scaffold and toluene in wireframe; selected solvent molecules and all H atoms omitted for clarity.

Optimization

Having built the library of [LAuCl] complexes and chiral anions, to validate our H-bonded ACDC design, we investigated the cycloisomerization of 1,6-enynes of type 1 (Table 1), (26,27) a reaction that did not proceed at all employing [(Ph3P)Au(TRIP)]. (8) We commenced by screening different silver salts in combination with phosphinourea gold(I) chloride Au1 at room temperature. This complex combined with AgTRIP was catalytically inactive, whereas together with less basic N-triflyl phosphoramidate (25) salt Ag1 afforded the desired product 2a in good yield and encouraging 79:21 er (Table 1, entries 1 and 2). The more basic N-mesyl and N-phenylsulfonyl analogues Ag2 and Ag3 induced comparable enantioselectivity but much lower reactivity (Table 1, entries 3 and 4). A clear trend between reactivity and basicity emerged: less basic counterions from stronger parent Brønsted acids (19,28) are essential for reactivity presumably because they coordinate less strongly to the cationic Au(I) center (which is isolobal to a proton) (29) and therefore can be abstracted more easily via H-bonds. Among N-triflyl phosphoramidate salts, Ag1 provided better enantiocontrol than Ag4 and Ag5 (Table 1, entries 5 and 6), which bear different groups at the 3,3′-positions of the binaphthol backbone. Overall, the results summarized in Table 1 indicate that for a given anion, substitution at phosphorus influences basicity and coordinating ability, hence reactivity, while 3,3′ residues are responsible for enantioselectivity, as expected. (30)
Table 1. Screening of Silver Salts with Au1
entry(R)-AgXtime (h)yield (%)aerb
1AgTRIP96<5 
2Ag14.57079:21
3Ag24.5781.5:18.5
4Ag34.5983.5:16.5
5Ag44.54257:43
6Ag54.52251:49
a

Determined by 1H NMR against internal standard.

b

Determined by HPLC on chiral stationary phase.

Next, the library of gold complexes was evaluated with the optimal silver salt (R)-Ag1. Only key data providing insight into the working mode of the catalytic system are shown in Table 2 (see Supporting Information for comprehensive optimization studies). The presence of the urea group on the ligand is important for reactivity and essential for enantioselectivity, as [(JohnPhos)AuCl] with Ag1 delivered product 2a in lower yield and racemic form (Table 2, entries 1 and 2). Complex Au5, the thiourea analogue of Au1, was completely inactive, most likely because the S atom of the thiourea coordinates too strongly to the metal center, preventing substrate activation. This is consistent with the known thiophilicity of gold, (31) the Au–S bond observed in the crystal structure of related Au9, and previous studies on PPh3-based phosphinothourea Au(I) complexes. (18b) Moving the urea group from the para to the meta position of the biphenyl scaffold reversed the sense of enantioinduction (Table 2, entry 4 vs 1). Remarkably, either enantiomer of the product can thus be obtained preferentially using the same enantiomer of the chiral anion, in combination with a different achiral cocatalyst. Complexes Au1416, equipped with urea or squaramide groups on triarylphosphine scaffolds, (18b) were poorly active and afforded (almost) racemic product (Table 2, entries 5–7).
Table 2. Screening of Selected Au(I) Complexes with Ag1
entryLAuCltime (h)yield (%)aerb
1Au14.57079:21
2[(JohnPhos)AuCl]64050:50
3Au5440 
4Au84.57027.5:72.5
5Au14443250:50
6Au15441045:55
7Au16244656:44
8cAu10.79087:13
a

Determined by 1H NMR against internal standard.

b

Determined by HPLC on chiral stationary phase.

c

In benzene.

This data set highlights that a H-bond donor on the ligand is crucial for both reactivity and enantiocontrol. The mere presence of a chiral anion in the reaction mixture is not sufficient to transfer effectively the stereochemical information, as its proximity to the reaction center cannot be guaranteed in the absence of a suitably placed H-bond donor. These observations lend credibility to the original design, wherein the pendant urea was envisioned to enable both anion abstraction and precise positioning of the source of chirality (Scheme 2). The presence of H-bonding at the heart of the catalytic system can be inferred also from solvent effects. For instance, the performance of the optimal Au1/Ag1 combination improved by replacing dichloromethane with benzene (Table 2, entry 1 vs 8, 79:21 vs 87:13 er). The enantiocontrol generally increased at −20 °C (Table 3). A qualitative correlation between solvent polarity (defined by ENT values) (32) and enantioselectivity levels was observed, in agreement with H-bonding interactions between the urea and the anion being favored in apolar solvents. The least polar solvent for either the chlorinated or aromatic series (chloroform for entries 1–3 and toluene for entries 4–8 in Table 3) delivered the product with the highest enantiomeric ratio, up to 94.5:5.5 er.
Table 3. Solvent Effect with Au1/Ag1 Catalytic Systema
entrysolventENT (ref  (32))yield (%)aerb
1CHCl30.2593993:7
2CH2Cl20.3098592:8
3ClCH2CH2Cl0.3276088:12
4C6H5CH30.099>9594.5:5.5
51,2-(CH3)2C6H4na8693:7
6C6H5Cl0.188>9591:9
7C6H5F0.1947590:10
8C6H5CF3na8990.5:9.5
a

Determined by 1H NMR against internal standard.

b

Determined by HPLC on chiral stationary phase.

Once identified toluene as the best solvent, a final fine-tuning of the catalytic system was performed (Table 4). Replacing Ag1 with Ag6, i.e., swapping the triisopropylphenyl groups for 9-anthracenyl substituents at the 3,3′-binaphthol positions, led to a considerable improvement in enantioselectivity, although at the expense of conversion (Table 4, entries 1 and 2). In order to increase reactivity, more electrophilic Au(I) complex Au10 with a trifluoromethyl group meta to the phosphine was employed, and an excellent yield of 2a could be obtained at −10 °C while maintaining the high enantioselectivity (Table 4, entries 3 and 4). Complexes Au24, Au11, and Au12 with different substituents on the urea were also evaluated, since electronic properties influence H-bonding ability. (17,21) However, none of them surpassed the performance of complex Au10 possessing a simple phenyl urea, which offered the right balance between electronic activation and steric hindrance (see Supporting Information for details).
Table 4. Fine Tuning of the Catalytic Systema
entryLAuCl(R)-AgXT (°C)yield (%)aerb
1Au1Ag1–20>9594.5:5.5
2Au1Ag6–205098:2
3Au10Ag6–207898:2
4Au10Ag6–10>9598:2
a

Determined by 1H NMR against internal standard.

b

Determined by HPLC on chiral stationary phase.

Generality of the Enantioselective 5-exo-dig and 6-exo-dig Cyclizations

The substrate scope of the formal [4 + 2] cycloaddition of enynes 1 catalyzed by the Au10/Ag6 system was then assessed (Scheme 5). Enynes 1ah bearing various linkers underwent cyclization to deliver the corresponding products 2ah in high yield (84–95%) and with excellent enantioselectivity (96:4 to 98:2 er). The standard reaction to 2a could be performed on 2-mmol scale in 24 h under air and in technical-grade toluene, with Au(I) and Ag(I) loading reduced to 1 mol %, with a 99% yield and 98:2 er for the product. This catalyst loading is the lowest reported so far in asymmetric cycloisomerizations of enynes 1, which is remarkable considering that all previous methods relied on stereochemical information covalently embedded in a chiral ligand. (27) Apart from ethers (2a, 2b), also allyl (2c), ester (2d, 2h), carbamate (2e), tosylate (2f), and acetal groups (2g) were tolerated. As for variations on the alkyne moiety, products with electron-poor as well as electron-rich groups at the para (2il), ortho (2m,n), and meta position (2o,p) of the aromatic ring were obtained with good to excellent yield and enantiocontrol (85–98%, 89.5:10.5 to 98.5:1.5 er). Also substrates 1q and 1r, possessing respectively 1-naphthyl and benzothiophenyl substituents at the alkyne terminus, underwent reaction smoothly and enantioselectively. The attainment of such high levels of enantiocontrol across a broad scope is unprecedented in this transformation. (27)

Scheme 5

Scheme 5. Enantioselective Formal [4 + 2] Cycloadditions of 1,6-Enynes Based on 5-exo-dig Cyclizationsa

aReactions performed under Ar or N2 in anhydrous toluene (0.1 or 0.2 M), unless otherwise stated. Yields of material isolated after purification, er determined by HPLC or SFC on chiral stationary phase.

bCarried out at 2 mmol scale, with 1 mol % Au10 and 1 mol % Ag6, in technical-grade toluene (0.6 M) under air for 24 h.

cAt 23 °C.

dAt 0 °C.

eWith 10 mol % Au10 and 10 mol % Ag6.

fIncluding 5% of inseparable 6-endo-isomer.

gReaction time: 96 h.

Compared to linker and alkyne variations, the catalytic system showed higher sensitivity to changes to the alkene part (compare 2a with 2sw in Scheme 5), in line with the fact that the chiral catalytic ensemble needs to discriminate the two enantiotopic faces of the double bond. Thus, enynes 1s and 1t possessing geranyl or trans-cinnamyl groups cyclized to products 2s and 2t in less than 48 h (90–99% yield, 89:11 to 95.5:4.5 er), even though higher temperatures and/or catalyst loadings were required. Nevertheless, this marks a significant improvement with respect to previous chiral gold catalysts, which required 7–14 days at room temperature. (27e) Enyne 1u possessing a cis-cinnamyl substituent afforded product 2u in lower yield (48%, due to a competing cycloisomerization to an achiral diene product) and enantioselectivity (82:18 er). Formation of epimeric products 2t and 2u highlights the stereospecificity of the reaction with respect to the alkene configuration. More hindered substrates such as 1v and 1w, bearing respectively a cyclohexene or a tetrasubstituted alkene, delivered products 2v and 2w in good yield (61–95%) and moderate but encouraging enantioselectivity (80:20 to 85:15 er), considering that such bulky enynes were never engaged in this cycloisomerization, let alone asymmetrically.
Judging by the enantioselectivity of the 5-exo-dig cyclizations presented in Scheme 5, the key interaction between the urea and the anion is not disrupted by H-bond acceptors on the substrates (such as ester, carbamate, tosylate, and nitro groups in 1df, 1h, 1l), even though they are present in up to 40-fold excess with respect to the chiral anion. Protic additives with H-bond donor ability were also tolerated, as spiking the reaction of model enyne 1a with 5 equiv of methanol still delivered product 2a in 80% NMR yield and 97:3 er. These observations suggest that H-bonding between the urea and the anion is strong enough to make for an effective and robust catalytic system, even if based on noncovalent interactions. To further prove this point, the Au10/Ag6 catalytic system was applied also to the 6-endo-dig cyclization of enynes, with and without the addition of protic exogenous nucleophiles (Scheme 6).

Scheme 6

Scheme 6. Enantioselective 6-endo-dig Cyclizations of 1,6-Enynes without (A) or with (B) Nucleophile Additiona

aYields of material isolated after purification, er determined by HPLC or SFC on chiral stationary phase.

bFor 48 h.

cAt 23 °C.

Thus, O-tethered enyne 3 was converted to oxabicyclo[4.1.0]hept-4-ene 4 in 70% yield and 92.5:7.5 er. (33) Cyclization of benzene-tethered enyne 5 followed by the addition of O-based nucleophiles (27e,34) delivered compounds 6ae in moderate to good yield and er (55–77%, 88.5:11.5 to 92.5:7.5 er). (35) The addition of a N-centered nucleophile to enynes of type 5 was demonstrated for the first time, obtaining azide 6f in 62% yield and 90:10 er. Instead, fluoride addition afforded product 6g in 69% yield and only 61:39 er. The low enantiocontrol in the formation of 6g can be explained by the strong tendency of the fluoride ion to H-bond with the urea, (36) thus preventing the key interaction that keeps the chiral anion in place.
Importantly, derivatization of the water- and azide-addition products gave access to O- and N-derivatives equivalent to the formal addition of noncompetent nucleophiles, thus expanding the scope and underlying the usefulness of this enantioselective protocol. For example, alcohol 6a was transformed into benzoate 6h, as well as into diene 7, which displays the core of the carexane natural products (Scheme 7A). (27e,37) Triazole 6i, primary amine 6j, and amide 6k were easily obtained from azide 6f without erosion of the enantiopurity (Scheme 7B). Cyclizations of 1,6-enynes bearing a terminal alkyne generally proceeded with lower enantioselectivity than those of internal aryl alkyne substrates (Scheme 7C). The methoxycyclization of benzene-tethered enyne 8 delivered product 9 in 94% yield and 78.5:21.5 er, which could be increased to 87:13 er at the expense of yield (see Supporting Information for details). (38) In order to preserve the stereocenter, the cyclopropyl gold carbene generated upon 5-exo-dig cycloisomerization of O-tethered enyne 10 was trapped in situ with diphenyl sulfoxide. (39) Under unoptimized conditions, cyclopropyl aldehyde 11 was obtained in 42% yield and 83:17 er, which compares favorably with the only other enantioselective preparation reported so far (3% ee). (40)

Scheme 7

Scheme 7. Derivatization and Scope Extension

aPrepared from a batch of 6f with 93:7 er (see Supporting Information).

Finally, we sought to extend this H-bonded counterion-directed catalysis strategy to other types of reactions. At 1 mol % loading, phosphinosquaramide complex Au16 in combination with (R)-Ag6 catalyzes the tandem cyclization–indole addition to 2-alkynyl enones 12, affording furans 14ad in good yield and enantioselectivity (Scheme 8). (41,13) In this case, the stereocenter is not created during the cycloisomerization but forms in the subsequent intermolecular nucleophilic attack to a carbocation intermediate. Chiral salt (R)-Ag6 alone affords predominantly the opposite enantiomer of product 14a, and the Au10/(R)-Ag6 combination employed for 1,6-enynes gives slightly lower enantioselectivity. These results indicate that tuning of the achiral catalytic component to get high enantioselectivity is required for each reaction class (see Supporting Information for more examples, including an allenol (1) cyclization).

Scheme 8

Scheme 8. Enantioselective Cycloisomerization–Indole Addition to 2-Alkynyl Enonesa

aYields of material isolated after purification, er determined by HPLC or SFC on chiral stationary phase.

Mechanistic Studies

Additional control experiments were conducted to shed light on the working mode of the final, optimized Au10/Ag6 catalytic system (Scheme 9 and Table 5, see Supporting Information for further tests). To this end, cationic complex Au17 was prepared by treatment of Au10 with equimolar AgSbF6 in the presence of acetonitrile. Its structure was confirmed by X-ray diffraction, showing that in the solid state the urea H-bonds to the fluoride of the counteranion. When AgSbF6 was replaced by (R)-Ag6, neutral complex Au18 was isolated instead, with no incorporation of acetonitrile as indicated by NMR.

Scheme 9

Scheme 9. Synthesis of Complexes with Ureaphosphine L10
Table 5. Control Experiments
entry[Au][Na or Ag]yield (%)aerb
1Au10(R)-Ag6>9598:2
2Au18 >9593:7
3[(JohnPhos)AuCl](R)-Ag60 
4Au10(R)-Na60 
5[(JohnPhos)Au(NCMe)]SbF6(R)-Ag62450:50
6[(JohnPhos)Au(NCMe)]SbF6(R)-Na6350:50
7Au17(R)-Ag6>9550:50
8Au17(R)-Na6>9579:21
9cAu17(R)-Na6>9590:10
a

Determined by 1H NMR against internal standard.

b

Determined by HPLC on chiral stationary phase.

c

With 15-crown-5 (100 mol %).

These complexes were then used in a series of informative control experiments, which highlighted how the chloride ligand and the countercation of the chiral salt play a role too (Table 5). Preformed complex Au18 delivered product 2a with yield and enantioselectivity comparable to the in situ combination of Au10 and (R)-Ag6 (93:7 er vs 98:2 er, Table 5, entries 1 and 2). No reactivity was detected employing either [(JohnPhos)AuCl] with (R)-Ag6 or Au10 with (R)-Na6 (Table 5, entries 3 and 4), indicating respectively that the urea is required to remove the chiral anion from Au and that sodium, unlike silver, is not able to scavenge the chloride ligand. Experiments with cationic complexes Au17 and its urea-free counterpart [(JohnPhos)Au(NCMe)]SbF6 were then performed. Combining [(JohnPhos)Au(NCMe)]SbF6 with (R)-Ag6 and (R)-Na6 delivered racemic material in low yield (Table 5, entries 5 and 6), thus emphasizing the importance of the tethered urea not only for reactivity but also for enantioselectivity, achieved through precise positioning of the chiral anion via H-bonding. In an apparently surprising outcome, also cationic complex Au17 combined with (R)-Ag6 yielded racemic product (Table 5, entry 7). This can actually be explained by catalysis carried out by achiral cationic species Au17 on its own, because the chiral anion associates preferentially to Ag+ over Au+ (or the urea). In agreement with this picture, product 2a was obtained with 79:21 er when (R)-Na6 was used (Table 5, entry 8). Moreover, when 1 equiv of 15-crown-5 ether was added in order to chelate Na+ and thus direct the anion to Au+, the enantiomeric ratio of the product further improved to 90:10 (Table 5, entry 9). Therefore, in order to attain high enantioselectivity using this H-bonded system, it is crucial to tie the generation of the catalytically competent cationic Au(I) center to the removal of other cations (Ag+ and to a lesser extent Na+), which would otherwise “sequester” the chiral anion, leading to racemic background reactivity. In this sense, when combining Au10 and (R)-Ag6in situ, precipitation of AgCl not only frees up a coordination site on gold but also ensures that Ag+ is removed from deleterious solution equilibria with the anion.
Further insights into the catalytic system are presented in Scheme 10. First of all, single-crystal X-ray diffraction of products 2h and 6a and comparison of optical rotations with available literature values indicate that the newly created stereocenter is (R)-configured in both 5-exo-dig and 6-endo-dig reactions. This implies that in the enantiodetermining step of both cyclization modes, the Si face of the alkene attacks the Au(I)-activated alkyne, delivering the corresponding cyclopropyl Au(I) carbene intermediates (Scheme 10A). The absence of nonlinear effects (Scheme 10B) suggests that only one chiral anion is involved in the enantiodetermining step, in line with formation of the expected 1:1 urea:anion complex. (21b,c)

Scheme 10

Scheme 10. Mechanistic Investigations: (A) Enantiofacial Selectivity, (B) Study of Nonlinear Effects, (C) Use of Complex Au4 as Au10 Surrogate, (D) 1H NMR Titration of Au4 with (R)-Na6 (298 K, CD2Cl2), (E–J) Kinetic Studies
Spectroscopic observation of the H-bonding interaction between the urea of complex Au10 and the chiral anion proved to be nontrivial. Mixing of Au10 and (R)-Ag6, in various ratios and in the absence of substrate, invariably resulted in at least partial chloride abstraction. In turn this led to poorly resolved NMR spectra, where species tentatively identified as chiral [LAuX] Au18 with the anion behaving as anionic ligand, and achiral chloride-bridged dinuclear complexes predominated. On the other hand, when Na6 was used instead of Ag6 to circumvent chloride scavenging, no changes were detected by NMR. In this last case, the chiral anion presumably remained associated with Na+ without interacting at all with the neutral urea. We resorted to study the combination of (R)-Na6 and complex Au4 in the presence of a constant excess of 15-crown-5, to force dissociation of the sodium cation, while at the same time avoiding undesired chloride scavenging. Au(I) chloride complex Au4 was chosen as a convenient surrogate for Au10, since the NH signals of its bis(trifluoromethyl)phenylurea resonate in a free region of the 1H NMR spectrum. The performance of Au4 in the standard asymmetric reaction is comparable to that of Au10, even if with marginally lower reactivity (consistent with the absence of the meta CF3 group) and enantioselectivity (Scheme 10C). Gratifyingly, 1H NMR titration of complex Au4 (10 mM in CD2Cl2 at 25 °C) with increasing amounts of (R)-Na6 in the presence of 20 equiv of 15-crown-5 resulted in clear deshielding of both NH signals of the urea, indicating their engagement in H-bonds with the anion (Scheme 10D). (42) The titration was recorded at the same catalyst concentration present in the reaction mixtures (5–10 mM), and the establishment of such H-bonds during the reaction is expected to be entropically even more facile because after AgCl precipitation the Au(I)-bound chiral anion should already be in close proximity to the H-bond donor.
Finally, kinetic studies on the cycloisomerization of 1a to 2a catalyzed by the Au10/(R)-Ag6 system in toluene-d8 at −10 °C were undertaken. By use of the variable time normalization analysis (VTNA) introduced by Burés, (43) the reaction was found to be approximately first order in catalyst (0.9) and 0.5 order in substrate (Scheme 10E,F). A first order in catalyst is common to most catalyzed transformations, provided that the catalyst does not decompose or aggregate into off-cycle species. A partial, noninteger order in substrate is expected for unimolecular catalyzed reactions that follow Briggs–Haldane kinetics (44) and possess a Michaelis–Menten constant (45) (KM) similar to substrate concentration. (18b) To verify that this was indeed the case for the cycloisomerization under study, the reaction progress kinetic analysis (RPKA) popularized by Blackmond (46) was carried out. 1H NMR monitoring of the reaction indicated clean conversion of enyne 1a to product 2a (Scheme 10G), but rate analysis revealed an initial induction period (Scheme 10H), most likely related to a noninstantaneous chloride abstraction. (47) The double reciprocal Lineweaver–Burk plot was thus constructed using data in the 1.5–10 h time range (Scheme 10I), (48) obtaining a Michaelis–Menten constant (KM) of 74 ± 4 mM. (49) Given the 0–100 mM substrate concentration present during the reaction, this intermediate KM value justifies the partial order in substrate determined by VTNA. The experimental 0.5 order found for the entire reaction course falls within the range of the calculated elasticity coefficient ε, (50) which predicts the changing order in substrate for Briggs–Haldane kinetic regimes from KM and substrate concentration (Scheme 10J).
Scheme 11 presents a tentative mechanism for the enantioselective formal [4 + 2] cycloaddition of enyne 1a catalyzed by Au10/(R)-Ag6, which takes into account all the experimental, spectroscopic, and kinetic evidence discussed above, as well as previous studies on this cycloisomerization. (26b,27e) Upon mixing Au(I) chloride complex Au10 with (R)-Ag6, neutral complex I possessing a phosphoramidate anionic ligand forms. The expected chloride scavenging accompanied by precipitation of AgCl matches the observed formation of a solution (with very few solid grains) upon addition of a solution of (R)-Ag6 to a thick whitish suspension of Au10 and enyne in toluene. Species I represents the entry to the catalytic cycle and coincides with Au18, prepared ex situ (see Scheme 9) and catalytically competent (see Table 5). In the presence of enyne 1, neutral complex I is proposed to be in equilibrium with cationic complex II, wherein the anion is H-bonded to the pendant urea and the alkyne coordinates to Au. The presence of this equilibrium is consistent with the partial order in substrate observed in the Briggs–Haldane kinetics. Additionally, H-bonding interactions between the anion and both NH groups of the urea were observed spectroscopically in a model system (see Scheme 10D). Regarding their precise geometrical arrangement, we propose that the NH residues establish two H-bonds with the phosphoryl O atom. This speculation is based on solid state considerations and DFT calculations with implicit solvent models. (51) In the crystal structures of (R)-Ag6 and related triethylammonium salt (R)-Et3NH6, the P–O and S–O bonds are almost parallel with the two O atoms forming a ∼3 Å wide “pincer” (see inset in Scheme 11), while Au10 has a 2.0 Å H–H distance between the urea NH groups. The most stable conformer computed by DFT for the model diphenylurea–chiral phosphoramidate couple shows two H-bonds to the phosphoryl O atom; however, a conformer where the NH groups H-bond to both the phosphoryl and sulfonyl O atoms is only 0.8 kcal/mol higher in energy. (51)

Scheme 11

Scheme 11. Proposed Mechanism for the Cyclization of Enyne 1 under H-Bonded Counterion-Directed Au(I) Catalysis
The enyne is expected to be oriented as depicted for complex II (Scheme 11), with the arene pointing toward the more crowded BINOL region and the linker in an unencumbered zone. This would be consistent with the lower er (92.5:7.5 to 95:5) observed in the cyclization of ortho-substituted substrates 1m, 1n, and 1q, which are not so well accommodated, and with the catalyst ability to tolerate instead even very bulky linkers on substrates 1ag. At this point, the enantiodetermining C–C bond formation takes place, i.e., the attack of the Si face of the alkene to the activated alkyne, leading to cyclopropyl Au(I) carbene III. (52) Friedel–Crafts-type ring expansion, deprotonation of the Wheland intermediate, and protodeauration then afford species IV. Product/substrate ligand exchange is likely mediated by the anion (18b) via the intermediacy of species I, thus closing the cycle and releasing product 2.

Conclusions

Click to copy section linkSection link copied!

We describe the concept of asymmetric H-bonded counterion-directed catalysis, based on H-bonding interactions between a chiral anion and a suitably positioned H-bond donor group on JohnPhos-type ligands for Au(I). The presence of such interactions was substantiated by 1H NMR titrations and structure–activity studies with modified ligands and chiral salts, as well as by the observed solvent and fluoride effects. For the first time, a broad range of alkyne substrates was engaged in challenging enantioselective gold-catalyzed reactions using chiral anions as the source of the stereochemical information, at catalyst loading down to 1 mol % and with excellent functional group tolerance.
This new paradigm, with a modular, short, and tunable synthesis of the two catalytic components, has the potential to speed up the development of enantioselective versions of various transition-metal catalyzed reactions, provided that the ligand for the metal of choice is equipped with a suitably placed H-bond donor group for a chiral anion.

Supporting Information

Click to copy section linkSection link copied!

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/jacs.1c11978.

  • Optimization tables, procedures, characterization, kinetic data, NMR spectra, SFC and HPLC traces, DFT computations, crystallographic data (PDF)

Accession Codes

CCDC 21077462107758 contain the supplementary crystallographic data for this paper. These data can be obtained free of charge via www.ccdc.cam.ac.uk/data_request/cif, or by emailing [email protected], or by contacting The Cambridge Crystallographic Data Centre, 12 Union Road, Cambridge CB2 1EZ, UK; fax: +44 1223 336033.

Terms & Conditions

Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.

Author Information

Click to copy section linkSection link copied!

  • Corresponding Author
    • Antonio M. Echavarren - Institute of Chemical Research of Catalonia (ICIQ), Barcelona Institute of Science and Technology, Av. Països Catalans 16, 43007 Tarragona, SpainDepartament de Química Orgànica i Analítica, Universitat Rovira i Virgili, C/Marcel·lí Domingo s/n, 43007 Tarragona, SpainOrcidhttps://orcid.org/0000-0001-6808-3007 Email: [email protected]
  • Authors
    • Allegra Franchino - Institute of Chemical Research of Catalonia (ICIQ), Barcelona Institute of Science and Technology, Av. Països Catalans 16, 43007 Tarragona, Spain
    • Àlex Martí - Institute of Chemical Research of Catalonia (ICIQ), Barcelona Institute of Science and Technology, Av. Països Catalans 16, 43007 Tarragona, SpainDepartament de Química Orgànica i Analítica, Universitat Rovira i Virgili, C/Marcel·lí Domingo s/n, 43007 Tarragona, Spain
  • Notes
    The authors declare no competing financial interest.

Acknowledgments

Click to copy section linkSection link copied!

We thank the MCIN/AEI/10.13039/501100011033 (Grants PID2019-104815GB-I00 and CEX2019-000925-S), the European Union (Horizon 2020 Marie Skłodowska-Curie COFUND Postdoctoral Fellowship 754510 to A.F.), the European Research Council (Advanced Grant 835080), the AGAUR (Grant 2017 SGR 1257 and FI Fellowship to À.M.), and CERCA Program/Generalitat de Catalunya for financial support. We also sincerely thank the ICIQ X-ray diffraction (especially Dr. Eduardo Escudero), NMR, chromatography, and mass spectrometry units.

References

Click to copy section linkSection link copied!

This article references 52 other publications.

  1. 1
    Hamilton, G. L.; Kang, E. J.; Mba, M.; Toste, F. D. A Powerful Chiral Counterion Strategy for Asymmetric Transition Metal Catalysis. Science 2007, 317, 496499,  DOI: 10.1126/science.1145229
  2. 2
    (a) Mayer, S.; List, B. Asymmetric Counteranion-Directed Catalysis. Angew. Chem. Int. Ed. 2006, 45, 41934195,  DOI: 10.1002/anie.200600512
    (b) Mukherjee, S.; List, B. Chiral Counteranions in Asymmetric Transition-Metal Catalysis: Highly Enantioselective Pd/Brønsted Acid-Catalyzed Direct α-Allylation of Aldehydes. J. Am. Chem. Soc. 2007, 129, 1133611337,  DOI: 10.1021/ja074678r
    (c) Phipps, R. J.; Hamilton, G. L.; Toste, F. D. The Progression of Chiral Anions from Concepts to Applications in Asymmetric Catalysis. Nat. Chem. 2012, 4, 603614,  DOI: 10.1038/nchem.1405
    (d) Mahlau, M.; List, B. Asymmetric Counteranion-Directed Catalysis: Concept, Definition, and Applications. Angew. Chem., Int. Ed. 2013, 52, 518533,  DOI: 10.1002/anie.201205343
    (e) Brak, K.; Jacobsen, E. N. Asymmetric Ion-Pairing Catalysis. Angew. Chem., Int. Ed. 2013, 52, 534561,  DOI: 10.1002/anie.201205449
    (f) Shirakawa, S.; Maruoka, K. Recent Developments in Asymmetric Phase-Transfer Reactions. Angew. Chem., Int. Ed. 2013, 52, 43124348,  DOI: 10.1002/anie.201206835
  3. 3

    For work on the combination of achiral Au(I) complexes and chiral anions, see the following:

    (a) Reference (1).
    (b) LaLonde, R. L.; Wang, Z. J.; Mba, M.; Lackner, A. D.; Toste, F. D. Gold(I)-Catalyzed Enantioselective Synthesis of Pyrazolidines, Isoxazolidines, and Tetrahydrooxazines. Angew. Chem., Int. Ed. 2010, 49, 598601,  DOI: 10.1002/anie.200905000
    (c) Zi, W.; Toste, F. D. Gold(I)-Catalyzed Enantioselective Desymmetrization of 1,3-Diols through Intramolecular Hydroalkoxylation of Allenes. Angew. Chem., Int. Ed. 2015, 54, 1444714451,  DOI: 10.1002/anie.201508331
    (d) Pedrazzani, R.; An, J.; Monari, M.; Bandini, M. New Chiral BINOL-Based Phosphates for Enantioselective [Au(I)]-Catalyzed Dearomatization of β-Naphthols with Allenamides. Eur. J. Org. Chem. 2021, 2021, 17321736,  DOI: 10.1002/ejoc.202100166
  4. 4

    For work on the combination of chiral Au(I) complexes and chiral anions, see the following:

    (a) Aikawa, K.; Kojima, M.; Mikami, K. Axial Chirality Control of Gold(biphep) Complexes by Chiral Anions: Application to Asymmetric Catalysis. Angew. Chem., Int. Ed. 2009, 48, 60736077,  DOI: 10.1002/anie.200902084
    (b) Aikawa, K.; Kojima, M.; Mikami, K. Synergistic Effect: Hydroalkoxylation of Allenes through Combination of Enantiopure BIPHEP-Gold Complexes and Chiral Anions. Adv. Synth. Catal. 2010, 352, 31313135,  DOI: 10.1002/adsc.201000672
    Corrigendum:Aikawa, K.; Kojima, M.; Mikami, K. Adv. Synth. Catal. 2011, 353, 28822883,  DOI: 10.1002/adsc.201100838
    (c) Barreiro, E. M.; Broggini, D. F. D.; Adrio, L. A.; White, A. J. P.; Schwenk, R.; Togni, A.; Hii, K. K. Gold(I) Complexes of Conformationally Constricted Chiral Ferrocenyl Phosphines. Organometallics 2012, 31, 37453754,  DOI: 10.1021/om300222k
    (d) Miles, D. H.; Veguillas, M.; Toste, F. D. Gold(I)-Catalyzed Enantioselective Bromocyclization Reactions of Allenes. Chem. Sci. 2013, 4, 34273431,  DOI: 10.1039/c3sc50811k
    (e) Handa, S.; Lippincott, D. J.; Aue, D. H.; Lipshutz, B. H. Asymmetric Gold-Catalyzed Lactonizations in Water at Room Temperature. Angew. Chem., Int. Ed. 2014, 53, 1065810662,  DOI: 10.1002/anie.201404729
  5. 5
    (a) Mourad, A. K.; Leutzow, J.; Czekelius, C. Anion-Induced Enantioselective Cyclization of Diynamides to Pyrrolidines Catalyzed by Cationic Gold Complexes. Angew. Chem., Int. Ed. 2012, 51, 1114911152,  DOI: 10.1002/anie.201205416
    (b) Spittler, M.; Lutsenko, K.; Czekelius, C. Total Synthesis of (+)-Mesembrine Applying Asymmetric Gold Catalysis. J. Org. Chem. 2016, 81, 61006105,  DOI: 10.1021/acs.joc.6b00985
  6. 6

    For a review on the combination of Au(I) methyl complexes and excess Brønsted acid for tandem reactions, see the following:

    (a) Inamdar, S. M.; Konala, A.; Patil, N. T. When gold meets chiral Brønsted acid catalysis: extending the boundaries of enantioselective gold catalysis. Chem. Commun. 2014, 50, 1512415135,  DOI: 10.1039/C4CC04633A

    For selected examples, see the following:

    (b) Han, Z.-Y.; Xiao, H.; Chen, X.-H.; Gong, L.-Z. Consecutive Intramolecular Hydroamination/Asymmetric Transfer Hydrogenation under Relay Catalysis of an Achiral Gold Complex/Chiral Brønsted Acid Binary System. J. Am. Chem. Soc. 2009, 131, 91829183,  DOI: 10.1021/ja903547q
    (c) Muratore, M. E.; Holloway, C. A.; Pilling, A. W.; Storer, R. I.; Trevitt, G.; Dixon, D. J. Enantioselective Brønsted Acid-Catalyzed N-Acyliminium Cyclization Cascades. J. Am. Chem. Soc. 2009, 131, 1079610797,  DOI: 10.1021/ja9024885
    (d) Liu, X.-Y.; Che, C.-M. Highly Enantioselective Synthesis of Chiral Secondary Amines by Gold(I)/Chiral Brønsted Acid Catalyzed Tandem Intermolecular Hydroamination and Transfer Hydrogenation Reactions. Org. Lett. 2009, 11, 42044207,  DOI: 10.1021/ol901443b
    (e) Tu, X.-F.; Gong, L.-Z. Highly Enantioselective Transfer Hydrogenation of Quinolines Catalyzed by Gold Phosphates: Achiral Ligand Tuning and Chiral-Anion Control of Stereoselectivity. Angew. Chem., Int. Ed. 2012, 51, 1134611349,  DOI: 10.1002/anie.201204179
    (f) Cala, L.; Mendoza, A.; Fañanás, F. J.; Rodríguez, F. A Catalytic Multicomponent Coupling Reaction for the Enantioselective Synthesis of Spiroacetals. Chem. Commun. 2013, 49, 27152717,  DOI: 10.1039/c3cc00118k
    (g) Zhou, S.; Li, Y.; Liu, X.; Hu, W.; Ke, Z.; Xu, X. Enantioselective Oxidative Multi-Functionalization of Terminal Alkynes with Nitrones and Alcohols for Expeditious Assembly of Chiral α-Alkoxy-β-amino-ketones. J. Am. Chem. Soc. 2021, 143, 1470314711,  DOI: 10.1021/jacs.1c06178
  7. 7
    (a) Dorel, R.; Echavarren, A. M. Gold(I)-Catalyzed Activation of Alkynes for the Construction of Molecular Complexity. Chem. Rev. 2015, 115, 90289072,  DOI: 10.1021/cr500691k
    (b) Zuccarello, G.; Escofet, I.; Caniparoli, U.; Echavarren, A. M. New-Generation Ligand Design for the Gold Catalyzed Asymmetric Activation of Alkynes. ChemPlusChem 2021, 86, 12831296,  DOI: 10.1002/cplu.202100232
    (c) Barbazanges, M.; Augé, M.; Moussa, J.; Amouri, H.; Aubert, C.; Desmarets, C.; Fensterbank, L.; Gandon, V.; Malacria, M.; Ollivier, C. Enantioselective IrI-Catalyzed Carbocyclization of 1,6-Enynes by the Chiral Counterion Strategy. Chem.─Eur. J. 2011, 17, 1378913794,  DOI: 10.1002/chem.201102723
  8. 8
    Raducan, M.; Moreno, M.; Bour, C.; Echavarren, A. M. Phosphate Ligands in the Gold(I)-Catalysed Activation of Enynes. Chem. Commun. 2012, 48, 5254,  DOI: 10.1039/C1CC15739F
  9. 9

    For reviews, see the following:

    (a) Jia, M.; Bandini, M. Counterion Effects in Homogeneous Gold Catalysis. ACS Catal. 2015, 5, 16381652,  DOI: 10.1021/cs501902v
    (b) Zuccaccia, D.; Del Zotto, A.; Baratta, W. The Pivotal Role of the Counterion in Gold Catalyzed Hydration and Alkoxylation of Alkynes. Coord. Chem. Rev. 2019, 396, 103116,  DOI: 10.1016/j.ccr.2019.06.007
    (c) Lu, Z.; Li, T.; Mudshinge, S. R.; Xu, B.; Hammond, G. B. Optimization of Catalysts and Conditions in Gold(I) Catalysis─Counterion and Additive Effects. Chem. Rev. 2021, 121, 84528477,  DOI: 10.1021/acs.chemrev.0c00713
  10. 10

    For selected articles, see the following:

    (a) Zuccaccia, D.; Belpassi, L.; Tarantelli, F.; Macchioni, A. Ion Pairing in Cationic Olefin-Gold(I) Complexes. J. Am. Chem. Soc. 2009, 131, 31703171,  DOI: 10.1021/ja809998y
    (b) Zuccaccia, D.; Belpassi, L.; Rocchigiani, L.; Tarantelli, F.; Macchioni, A. A Phosphine Gold(I) π-Alkyne Complex: Tuning the Metal-Alkyne Bond Character and Counterion Position by the Choice of the Ancillary Ligand. Inorg. Chem. 2010, 49, 30803082,  DOI: 10.1021/ic100093n
    (c) Bandini, M.; Bottoni, A.; Chiarucci, M.; Cera, G.; Miscione, G. P. Mechanistic Insights into Enantioselective Gold-Catalyzed Allylation of Indoles with Alcohols: The Counterion Effect. J. Am. Chem. Soc. 2012, 134, 2069020700,  DOI: 10.1021/ja3086774
    (d) Zhdanko, A.; Maier, M. E. Explanation of Counterion Effects in Gold(I)-Catalyzed Hydroalkoxylation of Alkynes. ACS Catal. 2014, 4, 27702775,  DOI: 10.1021/cs500446d
    (e) Biasiolo, L.; Del Zotto, A.; Zuccaccia, D. Toward Optimizing the Performance of Homogeneous L-Au-X Catalysts through Appropriate Matching of the Ligand (L) and Counterion (X). Organometallics 2015, 34, 17591765,  DOI: 10.1021/acs.organomet.5b00308
    (f) Lu, Z.; Han, J.; Okoromoba, O. E.; Shimizu, N.; Amii, H.; Tormena, C. F.; Hammond, G. B.; Xu, B. Predicting Counterion Effects Using a Gold Affinity Index and a Hydrogen Bonding Basicity Index. Org. Lett. 2017, 19, 58485851,  DOI: 10.1021/acs.orglett.7b02829
    (g) Schießl, J.; Schulmeister, J.; Doppiu, A.; Wörner, E.; Rudolph, M.; Karch, R.; Hashmi, A. S. K. An Industrial Perspective on Counter Anions in Gold Catalysis: Underestimated with Respect to “Ligand Effects”. Adv. Synth. Catal. 2018, 360, 24932502,  DOI: 10.1002/adsc.201800233
  11. 11
    (a) Neel, A. J.; Hilton, M. J.; Sigman, M. S.; Toste, F. D. Exploiting Non-covalent π Interactions for Catalyst Design. Nature 2017, 543, 637646,  DOI: 10.1038/nature21701
    (b) Fanourakis, A.; Docherty, P. J.; Chuentragool, P.; Phipps, R. J. Recent Developments in Enantioselective Transition Metal Catalysis Featuring Attractive Noncovalent Interactions between Ligand and Substrate. ACS Catal. 2020, 10, 1067210714,  DOI: 10.1021/acscatal.0c02957
  12. 12
    (a) Duarte, F.; Paton, R. S. Molecular Recognition in Asymmetric Counteranion Catalysis: Understanding Chiral Phosphate-Mediated Desymmetrization. J. Am. Chem. Soc. 2017, 139, 88868896,  DOI: 10.1021/jacs.7b02468
    (b) Orlandi, M.; Coelho, J. A. S.; Hilton, M. J.; Toste, F. D.; Sigman, M. S. Parametrization of Non-covalent Interactions for Transition State Interrogation Applied to Asymmetric Catalysis. J. Am. Chem. Soc. 2017, 139, 68036806,  DOI: 10.1021/jacs.7b02311
    (c) Shoja, A.; Reid, J. P. Computational Insights into Privileged Stereocontrolling Interactions Involving Chiral Phosphates and Iminium Intermediates. J. Am. Chem. Soc. 2021, 143, 72097215,  DOI: 10.1021/jacs.1c03829
  13. 13
    Zhang, Z.; Smal, V.; Retailleau, P.; Voituriez, A.; Frison, G.; Marinetti, A.; Guinchard, X. Tethered Counterion-Directed Catalysis: Merging the Chiral Ion-Pairing and Bifunctional Ligand Strategies in Enantioselective Gold(I) Catalysis. J. Am. Chem. Soc. 2020, 142, 37973805,  DOI: 10.1021/jacs.9b11154
  14. 14
    Michelet, V. Noble Metal-Catalyzed Enyne Cycloisomerizations and Related Reactions. Comprehensive Organic Synthesis, 2nd ed.; Elsevier, 2014; Vol. 5, pp 14831536.
  15. 15

    For recent reviews on asymmetric gold catalysis, see the following:

    (a) Zi, W.; Toste, F. D. Recent Advances in Enantioselective Gold Catalysis. Chem. Soc. Rev. 2016, 45, 45674589,  DOI: 10.1039/C5CS00929D
    (b) Li, Y.; Li, W.; Zhang, J. Gold-Catalyzed Enantioselective Annulations. Chem.─Eur. J. 2017, 23, 467512,  DOI: 10.1002/chem.201602822
    (c) Jiang, J.-J.; Wong, M.-K. Recent Advances in the Development of Chiral Gold Complexes for Catalytic Asymmetric Catalysis. Chem.─Asian J. 2021, 16, 364377,  DOI: 10.1002/asia.202001375
    (d) Porcel García, S. Gold Catalyzed Asymmetric Transformations. IntechOpen, 2021;  DOI: 10.5772/intechopen.97519 .
    (e) Reference (7b).
  16. 16
    (a) Cheng, X.; Zhang, L. Designed Bifunctional Ligands in Cooperative Homogeneous Gold Catalysis. CCS Chem. 2021, 3, 19892002,  DOI: 10.31635/ccschem.020.202000454
    (b) Zuccarello, G.; Zanini, M.; Echavarren, A. M. Buchwald-Type Ligands on Gold(I) Catalysis. Isr. J. Chem. 2020, 60, 360372,  DOI: 10.1002/ijch.201900179
  17. 17
    (a) Schreiner, P. R. Metal-Free Organocatalysis through Explicit Hydrogen Bonding Interactions. Chem. Soc. Rev. 2003, 32, 289296,  DOI: 10.1039/b107298f
    (b) Doyle, A. G.; Jacobsen, E. N. Small-Molecule H-Bond Donors in Asymmetric Catalysis. Chem. Rev. 2007, 107, 57135743,  DOI: 10.1021/cr068373r
  18. 18

    For the only previous examples of phosphino(thio)urea Au(I) complexes, see the following:

    (a) Campbell, M. J.; Toste, F. D. Enantioselective Synthesis of Cyclic Carbamimidates via a Three-Component Reaction of Imines, Terminal Alkynes, and p-Toluenesulfonylisocyanate Using a Monophosphine Gold(I) Catalyst. Chem. Sci. 2011, 2, 13691378,  DOI: 10.1039/c1sc00160d
    (b) Franchino, A.; Martí, À.; Nejrotti, S.; Echavarren, A. M. Silver-Free Au(I) Catalysis Enabled by Bifunctional Urea- and Squaramide-Phosphine Ligands via H-Bonding. Chem.─Eur. J. 2021, 27, 1198911996,  DOI: 10.1002/chem.202101751
  19. 19
    (a) Akiyama, T. Stronger Brønsted Acids. Chem. Rev. 2007, 107, 57445758,  DOI: 10.1021/cr068374j
    (b) Terada, M. Binaphthol-derived Phosphoric Acid as a Versatile Catalyst for Enantioselective Carbon-Carbon Bond Forming Reactions. Chem. Commun. 2008, 40974112,  DOI: 10.1039/b807577h
    (c) Akiyama, T.; Mori, K. Stronger Brønsted Acids: Recent Progress. Chem. Rev. 2015, 115, 92779306,  DOI: 10.1021/acs.chemrev.5b00041
    (d) Parmar, D.; Sugiono, E.; Raja, S.; Rueping, M. Complete Field Guide to Asymmetric BINOL-Phosphate Derived Brønsted Acid and Metal Catalysis: History and Classification by Mode of Activation; Brønsted Acidity, Hydrogen Bonding, Ion Pairing, and Metal Phosphates. Chem. Rev. 2014, 114, 90479153,  DOI: 10.1021/cr5001496
    Correction:Parmar, D.; Sugiono, E.; Raja, S.; Rueping, M. Chem. Rev. 2017, 117, 1060810620,  DOI: 10.1021/acs.chemrev.7b00197
  20. 20
    Wang, Y.; Wang, Z.; Li, Y.; Wu, G.; Cao, Z.; Zhang, L. A General Ligand Design for Gold Catalysis Allowing Ligand-Directed anti-Nucleophilic Attack of Alkynes. Nat. Commun. 2014, 5, 3470,  DOI: 10.1038/ncomms4470
  21. 21
    (a) Zhang, Z.; Schreiner, P. R. (Thio)urea Organocatalysis─What Can Be Learnt from Anion Recognition?. Chem. Soc. Rev. 2009, 38, 11871198,  DOI: 10.1039/b801793j
    (b) Amendola, V.; Fabbrizzi, L.; Mosca, L. Anion Recognition by Hydrogen Bonding: Urea-Based Receptors. Chem. Soc. Rev. 2010, 39, 38893915,  DOI: 10.1039/b822552b
    (c) Bregović, V. B.; Basarić, N.; Mlinarić-Majerski, K. Anion Binding with Urea and Thiourea Derivatives. Coord. Chem. Rev. 2015, 295, 80124,  DOI: 10.1016/j.ccr.2015.03.011
  22. 22

    For reviews on supramolecular ligands for asymmetric metal catalysis, see the following:

    (a) Meeuwissen, J.; Reek, J. N. H. Supramolecular Catalysis Beyond Enzyme Mimics. Nat. Chem. 2010, 2, 615621,  DOI: 10.1038/nchem.744
    (b) Carboni, S.; Gennari, C.; Pignataro, L.; Piarulli, U. Supramolecular Ligand-Ligand and Ligand-Substrate Interactions for Highly Selective Transition Metal Catalysis. Dalton Trans. 2011, 40, 43554373,  DOI: 10.1039/c0dt01517b
    (c) Bellini, R.; van der Vlugt, J. I.; Reek, J. N. H. Supramolecular Self-Assembled Ligands in Asymmetric Transition Metal Catalysis. Isr. J. Chem. 2012, 52, 613629,  DOI: 10.1002/ijch.201200002
    (d) Raynal, M.; Ballester, P.; Vidal-Ferran, A.; van Leeuwen, P. W. N. M. Supramolecular Catalysis. Part 1: Non-Covalent Interactions as a Tool for Building and Modifying Homogeneous Catalysts. Chem. Soc. Rev. 2014, 43, 16601733,  DOI: 10.1039/C3CS60027K
    (e) Ohmatsu, K.; Ooi, T. Design of Supramolecular Chiral Ligands for Asymmetric Metal Catalysis. Tetrahedron Lett. 2015, 56, 20432048,  DOI: 10.1016/j.tetlet.2015.02.096
    (f) Trouvé, J.; Gramage-Doria, R. Beyond Hydrogen Bonding: Recent Trends of Outer Sphere Interactions in Transition Metal Catalysis. Chem. Soc. Rev. 2021, 50, 35653584,  DOI: 10.1039/D0CS01339K

    For a seminal example involving ion-pairing interactions between an achiral ligand and a chiral anion, see the following:

    (g) Ohmatsu, K.; Ito, M.; Kunieda, T.; Ooi, T. Ion-Paired Chiral Ligands for Asymmetric Palladium Catalysis. Nat. Chem. 2012, 4, 473477,  DOI: 10.1038/nchem.1311
  23. 23
    Murata, M.; Buchwald, S. L. A General and Efficient Method for the Palladium-Catalyzed Cross-Coupling of Thiols and Secondary Phosphines. Tetrahedron 2004, 60, 73977403,  DOI: 10.1016/j.tet.2004.05.044
  24. 24
    Luchini, G.; Ascough, D. M. H.; Alegre-Requena, J. V.; Gouverneur, V.; Paton, R. S. Data-Mining the Diaryl(thio)urea Conformational Landscape: Understanding the Contrasting Behavior of Ureas and Thioureas with Quantum Chemistry. Tetrahedron 2019, 75, 697702,  DOI: 10.1016/j.tet.2018.12.033
  25. 25
    Nakashima, D.; Yamamoto, H. Design of Chiral N-Triflyl Phosphoramide as a Strong Chiral Brønsted Acid and Its Application to Asymmetric Diels-Alder Reaction. J. Am. Chem. Soc. 2006, 128, 96269627,  DOI: 10.1021/ja062508t
  26. 26
    (a) Nieto-Oberhuber, C.; López, S.; Echavarren, A. M. Intramolecular [4 + 2] Cycloadditions of 1,3-Enynes or Arylalkynes with Alkenes with Highly Reactive Cationic Phosphine Au(I) Complexes. J. Am. Chem. Soc. 2005, 127, 61786179,  DOI: 10.1021/ja042257t
    (b) Nieto-Oberhuber, C.; Pérez-Galán, P.; Herrero-Gómez, E.; Lauterbach, T.; Rodríguez, C.; López, S.; Bour, C.; Rosellón, A.; Cárdenas, D. J.; Echavarren, A. M. Gold(I)-Catalyzed Intramolecular [4 + 2] Cycloadditions of Arylalkynes or 1,3-Enynes with Alkenes: Scope and Mechanism. J. Am. Chem. Soc. 2008, 130, 269279,  DOI: 10.1021/ja075794x
  27. 27

    Previous asymmetric versions, all based on chiral ligands for Au(I), lack either scope or high enantiocontrol:

    (a) Two examples only (44–99% yield, 92–93% ee):Chao, C.-M.; Vitale, M. R.; Toullec, P. Y.; Genêt, J.-P.; Michelet, V. Asymmetric Gold-Catalyzed Hydroarylation/Cyclization Reactions. Chem.─Eur. J. 2009, 15, 13191323,  DOI: 10.1002/chem.200802341
    (b) One further example (86% yield, 16% ee):Pradal, A.; Chao, C.-M.; Vitale, M. R.; Toullec, P. Y.; Michelet, V. Asymmetric Au-Catalyzed Domino Cyclization/Nucleophile Addition Reactions of Enynes in the Presence of Water, Methanol and Electron-rich Aromatic Derivatives. Tetrahedron 2011, 67, 43714377,  DOI: 10.1016/j.tet.2011.03.071
    (c) Five examples (70–95% yield, 73–88% ee):Delpont, N.; Escofet, I.; Pérez-Galán, P.; Spiegl, D.; Raducan, M.; Bour, C.; Sinisi, R.; Echavarren, A. M. Modular Chiral Gold(I) Phosphite Complexes. Catal. Sci. Technol. 2013, 3, 30073012,  DOI: 10.1039/c3cy00250k
    (d) One example only (99% yield, 91% ee):Aillard, P.; Dova, D.; Magné, V.; Retailleau, P.; Cauteruccio, S.; Licandro, E.; Voituriez, A.; Marinetti, A. The Synthesis of Substituted Phosphathiahelicenes via Regioselective Bromination of a Preformed Helical Scaffold: A New Approach to Modular Ligands for Enantioselective Gold-Catalysis. Chem. Commun. 2016, 52, 1098410987,  DOI: 10.1039/C6CC04765C
    (e) 17 examples (61–99% yield, 58–92% ee):Zuccarello, G.; Mayans, J. G.; Escofet, I.; Scharnagel, D.; Kirillova, M. S.; Pérez-Jimeno, A. H.; Calleja, P.; Boothe, J. R.; Echavarren, A. M. Enantioselective Folding of Enynes by Gold(I) Catalysts with a Remote C2-Chiral Element. J. Am. Chem. Soc. 2019, 141, 1185811863,  DOI: 10.1021/jacs.9b06326
    (f) One example only (92% yield, 94% ee):Magné, V.; Sanogo, Y.; Demmer, C. S.; Retailleau, P.; Marinetti, A.; Guinchard, X.; Voituriez, A. Chiral Phosphathiahelicenes: Improved Synthetic Approach and Uses in Enantioselective Gold(I)-Catalyzed [2 + 2] Cycloadditions of N-Homoallenyl Tryptamines. ACS Catal. 2020, 10, 81418148,  DOI: 10.1021/acscatal.0c01819
  28. 28

    N-Triflyl phosphoric amide D1 has a pKa value 0.9 units lower than the corresponding phosphoric acid TRIP-H (4.2 vs 3.3, both values in DMSO at 25 °C):

    Christ, P.; Lindsay, A. G.; Vormittag, S. S.; Neudörfl, J.-M.; Berkessel, A.; O’Donoghue, A. C. pKa Values of Chiral Brønsted Acid Catalysts: Phosphoric Acids/Amides, Sulfonyl/Sulfuryl Imides, and Perfluorinated TADDOLs (TEFDDOLs). Chem.─Eur. J. 2011, 17, 85248528,  DOI: 10.1002/chem.201101157
  29. 29
    Raubenheimer, H. G.; Schmidbaur, H. Gold Chemistry Guided by the Isolobality Concept. Organometallics 2012, 31, 25072522,  DOI: 10.1021/om2010113
  30. 30
    Reid, J. P.; Goodman, J. M. Goldilocks Catalysts: Computational Insights into the Role of the 3,3′ Substituents on the Selectivity of BINOL-Derived Phosphoric Acid Catalysts. J. Am. Chem. Soc. 2016, 138, 79107917,  DOI: 10.1021/jacs.6b02825
  31. 31
    (a) Kepp, K. P. A Quantitative Scale of Oxophilicity and Thiophilicity. Inorg. Chem. 2016, 55, 94619470,  DOI: 10.1021/acs.inorgchem.6b01702
    (b) Izaga, A.; Herrera, R. P.; Gimeno, M. C. Gold(I)-Mediated Thiourea Organocatalyst Activation: A Synergic Effect for Asymmetric Catalysis. ChemCatChem 2017, 9, 13131321,  DOI: 10.1002/cctc.201601527
  32. 32
    Reichardt, C. Solvatochromic Dyes as Solvent Polarity Indicators. Chem. Rev. 1994, 94, 23192358,  DOI: 10.1021/cr00032a005
  33. 33

    All gold complexes used so far for the enantioselective cyclization of O-tethered 1,6-enynes bear chiral ligands. For a complete overview, see the following:

    (a) Mato, M.; Franchino, A.; García-Morales, C.; Echavarren, A. M. Gold-Catalyzed Synthesis of Small Rings. Chem. Rev. 2021, 121, 86138684,  DOI: 10.1021/acs.chemrev.0c00697

    For seminal examples, see the following:

    (b) Chao, C.-M.; Beltrami, D.; Toullec, P. Y.; Michelet, V. Asymmetric Au(I)-Catalyzed Synthesis of Bicyclo[4.1.0]heptene Derivatives via a Cycloisomerization Process of 1,6-Enynes. Chem. Commun. 2009, 69886990,  DOI: 10.1039/b913554e
    (c) Teller, H.; Corbet, M.; Mantilli, L.; Gopakumar, G.; Goddard, R.; Thiel, W.; Fürstner, A. One-Point Binding Ligands for Asymmetric Gold Catalysis: Phosphoramidites with a TADDOL-Related but Acyclic Backbone. J. Am. Chem. Soc. 2012, 134, 1533115342,  DOI: 10.1021/ja303641p
  34. 34
    Sanjuán, A. M.; Martínez, A.; García-García, P.; Fernández-Rodríguez, M. A.; Sanz, R. Gold(I)-catalyzed 6-endo hydroxycyclization of 7-substituted-1,6-enynes. Beilstein J. Org. Chem. 2013, 9, 22422249,  DOI: 10.3762/bjoc.9.263
  35. 35

    See Supporting Information for discussion on other nucleophiles and reaction side products.

  36. 36
    (a) Pfeifer, L.; Engle, K. M.; Pidgeon, G. W.; Sparkes, H. A.; Thompson, A. L.; Brown, J. M.; Gouverneur, V. Hydrogen-Bonded Homoleptic Fluoride-Diarylurea Complexes: Structure, Reactivity, and Coordinating Power. J. Am. Chem. Soc. 2016, 138, 1331413325,  DOI: 10.1021/jacs.6b07501
    (b) Ibba, F.; Pupo, G.; Thompson, A. L.; Brown, J. M.; Claridge, T. D. W.; Gouverneur, V. Impact of Multiple Hydrogen Bonds with Fluoride on Catalysis: Insight from NMR Spectroscopy. J. Am. Chem. Soc. 2020, 142, 1973119744,  DOI: 10.1021/jacs.0c09832
  37. 37
    D’Abrosca, B.; Fiorentino, A.; Golino, A.; Monaco, P.; Oriano, P.; Pacifico, S. Carexanes: Prenyl Stilbenoid Derivatives from Carex distachya. Tetrahedron Lett. 2005, 46, 52695272,  DOI: 10.1016/j.tetlet.2005.06.036
  38. 38
    Martín-Torres, I.; Ogalla, G.; Yang, J.-M.; Rinaldi, A.; Echavarren, A. M. Enantioselective Alkoxycyclization of 1,6-Enynes with Gold(I)-Cavitands: Total Synthesis of Mafaicheenamine C. Angew. Chem., Int. Ed. 2021, 60, 93399344,  DOI: 10.1002/anie.202017035
  39. 39
    Witham, C. A.; Mauleón, P.; Shapiro, N. D.; Sherry, B. D.; Toste, F. D. Gold(I)-Catalyzed Oxidative Rearrangements. J. Am. Chem. Soc. 2007, 129, 58385839,  DOI: 10.1021/ja071231+
  40. 40
    Wang, W.; Yang, J.; Wang, F.; Shi, M. Axially Chiral N-Heterocyclic Carbene Gold(I) Complex Catalyzed Asymmetric Cycloisomerization of 1,6-Enynes. Organometallics 2011, 30, 38593869,  DOI: 10.1021/om2004404
  41. 41
    (a) Yao, T.; Zhang, X.; Larock, R. C. AuCl3-Catalyzed Synthesis of Highly Substituted Furans from 2-(1-Alkynyl)-2-alken-1-ones. J. Am. Chem. Soc. 2004, 126, 1116411165,  DOI: 10.1021/ja0466964
    (b) Rauniyar, V.; Wang, Z. J.; Burks, H. E.; Toste, F. D. Enantioselective Synthesis of Highly Substituted Furans by a Copper(II)-Catalyzed Cycloisomerization-Indole Addition Reaction. J. Am. Chem. Soc. 2011, 133, 84868489,  DOI: 10.1021/ja202959n
    (c) Force, G.; Ki, Y. L. T.; Isaac, K.; Retailleau, P.; Marinetti, A.; Betzer, J.-F. Paracyclophane-based Silver Phosphates as Catalysts for Enantioselective Cycloisomerization/Addition Reactions: Synthesis of Bicyclic Furans. Adv. Synth. Catal. 2018, 360, 33563366,  DOI: 10.1002/adsc.201800587
  42. 42

    31P{1H} and 19F{1H} NMR spectra remained unchanged, confirming that the chloride was not scavenged. See Supporting Information for details.

  43. 43
    (a) Burés, J. A Simple Graphical Method to Determine the Order in Catalyst. Angew. Chem., Int. Ed. 2016, 55, 20282031,  DOI: 10.1002/anie.201508983
    (b) Burés, J. Variable Time Normalization Analysis: General Graphical Elucidation of Reaction Orders from Concentration Profiles. Angew. Chem., Int. Ed. 2016, 55, 1608416087,  DOI: 10.1002/anie.201609757
    (c) Nielsen, C. D.-T.; Burés, J. Visual Kinetic Analysis. Chem. Sci. 2019, 10, 348353,  DOI: 10.1039/C8SC04698K
  44. 44
    Briggs, G. E.; Haldane, J. B. S. A Note on the Kinetics of Enzyme Action. Biochem. J. 1925, 19, 338339,  DOI: 10.1042/bj0190338

    Application of this model rests on a series of underlying assumptions, all satisfied in the present case (see Supporting Information).

  45. 45
    (a) Michaelis, L.; Menten, M. L. Die Kinetik der Invertinwirkung. Biochem. Z. 1913, 49, 333369
    (b) Johnson, K. A.; Goody, R. S. The Original Michaelis Constant: Translation of the 1913 Michaelis−Menten Paper. Biochemistry 2011, 50, 82648269,  DOI: 10.1021/bi201284u
  46. 46
    Blackmond, D. G. Reaction Progress Kinetic Analysis: A Powerful Methodology for Mechanistic Studies of Complex Catalytic Reactions. Angew. Chem., Int. Ed. 2005, 44, 43024320,  DOI: 10.1002/anie.200462544
    Correction:Blackmond, D. G. Angew. Chem., Int. Ed. 2006, 45, 21622162,  DOI: 10.1002/anie.200690050
  47. 47
    Franchino, A.; Montesinos-Magraner, M.; Echavarren, A. M. Silver-Free Catalysis with Gold(I) Chloride Complexes. Bull. Chem. Soc. Jpn. 2021, 94, 10991117,  DOI: 10.1246/bcsj.20200358
  48. 48
    Lineweaver, H.; Burk, D. The Determination of Enzyme Dissociation Constants. J. Am. Chem. Soc. 1934, 56, 658666,  DOI: 10.1021/ja01318a036
  49. 49

    The uncertainty refers not to experimental variation but only to the mathematical error of the linear regression, as determined by Excel LINEST routine. If all data from time 0 to 10 h are used in the Lineawer–Burk plot, a KM value of 63 ± 4 is obtained, leading to identical conclusions regarding the partial order in substrate (see Supporting Information).

  50. 50
    Burés, J. What is the Order of a Reaction?. Top. Catal. 2017, 60, 631633,  DOI: 10.1007/s11244-017-0735-y
  51. 51

    Geometry optimizations were carried out using Gaussian 09 at the B3LYP-D3/6-31G(d,p) level of theory in toluene (SMD). Single point energy calculations were performed on the resulting structures employing B3LYP-D3/6-311+G(d,p)/SMD(toluene). See the Supporting Information for an overview of all computed structures. Refer to ref (12c) for an excellent discussion of alternative computational methods (functionals, basis sets, solvent models) in the context of chiral phosphate–iminium ion pairs.

  52. 52

    This step is assumed to be enantiodetermining on the basis of previous DFT calculations for the same reactions (refs (26) and (27e)).

Cited By

Click to copy section linkSection link copied!
Citation Statements
Explore this article's citation statements on scite.ai

This article is cited by 52 publications.

  1. Rui Wu, Zurong Xu, Dong Zhu, Shifa Zhu. Dirhodium-Catalyzed Asymmetric Transformations of Alkynes via Carbene Intermediates. Accounts of Chemical Research 2025, 58 (6) , 799-811. https://doi.org/10.1021/acs.accounts.4c00715
  2. Hannah K. Adams, Max Kadarauch, Nicholas J. Hodson, Arthur R. Lit, Robert J. Phipps. Design Approaches That Utilize Ionic Interactions to Control Selectivity in Transition Metal Catalysis. Chemical Reviews 2025, 125 (5) , 2846-2907. https://doi.org/10.1021/acs.chemrev.4c00849
  3. Pablo Mora, Imma Escofet, Maria Besora, Federica Cester Bonati, Antonio M. Echavarren. Practical Synthesis of Chiral Ferrocenenylphosphino-Gold(I) Catalysts and NEST Analysis of the Enantioinduction. ACS Catalysis 2025, 15 (3) , 2342-2350. https://doi.org/10.1021/acscatal.4c07495
  4. Yi-Bo Wang, Wei Liu, Ting Li, Yazhu Lu, Yi-Tian Yu, Hai-Tao Liu, Meiwen Liu, Pengfei Li, Peng-Cheng Qian, Hao Tang, Jia Guan, Long-Wu Ye, Long Li. Gold/HNTf2-Cocatalyzed Asymmetric Annulation of Diazo-Alkynes: Divergent Construction of Atropisomeric Biaryls and Arylquinones. Journal of the American Chemical Society 2024, 146 (49) , 33804-33816. https://doi.org/10.1021/jacs.4c12063
  5. Hao Xu, Zhenhao Zhang, Angela Marinetti, Xavier Guinchard. Enantioselective Au(I)-Catalyzed Cycloisomerization/Addition of Oxygen Nucleophiles to 2-Alkynylenones by the Tethered Counterion-Directed Catalysis Strategy. Organic Letters 2024, 26 (44) , 9525-9530. https://doi.org/10.1021/acs.orglett.4c03521
  6. Weiping Zhou, Renjie Wu, Jinchan Li, Dapeng Zhu, Biao Yu. A Ligand-Controlled Approach Enabling Gold(I)-Catalyzed Stereoinvertive Glycosylation with Primal Glycosyl ortho-Alkynylbenzoate Donors. Journal of the American Chemical Society 2024, 146 (40) , 27915-27924. https://doi.org/10.1021/jacs.4c10698
  7. Xu Wang, Junhao Yin, Wangqin Mao, Zhenyu Wang, Shuang Wu, Yang’en You. Cs2CO3 Promoted [4 + 2] Cycloaddition of 1,6-Enynes: An Approach to Tetrahydro-1H-benzo-f-isoindole Isomers. Organic Letters 2024, 26 (36) , 7757-7762. https://doi.org/10.1021/acs.orglett.4c02927
  8. Pilar Elías-Rodríguez, Manuel Benítez, Javier Iglesias-Sigüenza, Elena Díez, Rosario Fernández, José M. Lassaletta, David Monge. Hydrogen-Bonding Activation of Gold(I) Chloride Complexes: Enantioselective Synthesis of 3(2H)-Furanones by a Cycloisomerization-Addition Cascade. Organic Letters 2024, 26 (28) , 5995-6000. https://doi.org/10.1021/acs.orglett.4c02091
  9. Riccardo Pedrazzani, Sofia Kiriakidi, Magda Monari, Irene Lazzarini, Giulio Bertuzzi, Carlos Silva López, Marco Bandini. Fluorinated Biphenyl Phosphine Ligands for Accelerated [Au(I)]-Catalysis. ACS Catalysis 2024, 14 (8) , 6128-6136. https://doi.org/10.1021/acscatal.4c00593
  10. Kaylaa L. Gutman, Carlos D. Quintanilla, Liming Zhang. Catalytic Enantioselective Protonation of Gold Enolates Enabled by Cooperative Gold(I) Catalysis. Journal of the American Chemical Society 2024, 146 (6) , 3598-3602. https://doi.org/10.1021/jacs.3c11919
  11. Xuan Wu, Ke Zhao, Carlos D. Quintanilla, Liming Zhang. Chiral Bifunctional Phosphine Ligand Enables Asymmetric Trapping of Catalytic Vinyl Gold Carbene Species. Journal of the American Chemical Society 2024, 146 (4) , 2308-2312. https://doi.org/10.1021/jacs.3c10865
  12. Maurice P. Schrick, G. Kabelo Ramollo, Cristina-Maria Susanne Hirschbiegel, Manuel Fernandes, Andreas Lemmerer, Christoph Förster, Daniela I. Bezuidenhout, Katja Heinze. Redox Activation of Acyclic (Aryl)(amino)carbene Gold(I) Complexes. Organometallics 2024, 43 (2) , 69-84. https://doi.org/10.1021/acs.organomet.3c00395
  13. Shingo Harada, Shumpei Hirose, Mizuki Takamura, Maika Furutani, Yuna Hayashi, Tetsuhiro Nemoto. Silver(I)/Dirhodium(II) Catalytic Platform for Asymmetric N–H Insertion Reaction of Heteroaromatics. Journal of the American Chemical Society 2024, 146 (1) , 733-741. https://doi.org/10.1021/jacs.3c10596
  14. Jin-Bo Lu, Xiao-Qiu Xu, Zi-Sheng Ruan, Kai Liu, Ren-Xiao Liang, Yi-Xia Jia. Pd-Catalyzed Intramolecular Dearomative [4 + 2] Cycloaddition of Naphthalenes with Arylalkynes. Organic Letters 2023, 25 (45) , 8139-8144. https://doi.org/10.1021/acs.orglett.3c03240
  15. Alba Martínez-Bascuñana, José Luis Núñez-Rico, Lucas Carreras, Anton Vidal-Ferran. Counterion Variation: A Useful Lever for Maximizing the Regioselectivity in the Hydroboration of Terminal Alkynes. ACS Catalysis 2023, 13 (15) , 10447-10456. https://doi.org/10.1021/acscatal.3c02213
  16. Àlex Martí, Gala Ogalla, Antonio M. Echavarren. Hydrogen-Bonded Matched Ion Pair Gold(I) Catalysis. ACS Catalysis 2023, 13 (15) , 10217-10223. https://doi.org/10.1021/acscatal.3c02638
  17. Giuseppe Zuccarello, Leonardo J. Nannini, Ana Arroyo-Bondía, Nicolás Fincias, Isabel Arranz, Alba H. Pérez-Jimeno, Matthias Peeters, Inmaculada Martín-Torres, Anna Sadurní, Víctor García-Vázquez, Yufei Wang, Mariia S. Kirillova, Marc Montesinos-Magraner, Ulysse Caniparoli, Gonzalo D. Núñez, Feliu Maseras, Maria Besora, Imma Escofet, Antonio M. Echavarren. Enantioselective Catalysis with Pyrrolidinyl Gold(I) Complexes: DFT and NEST Analysis of the Chiral Binding Pocket. JACS Au 2023, 3 (6) , 1742-1754. https://doi.org/10.1021/jacsau.3c00159
  18. Jin-Ming Xie, Yun-Long Zhu, Yan-Ming Fu, Cheng-Feng Zhu, Lan-Jun Cheng, Yang-En You, Xiang Wu, You-Gui Li. Gold-Catalyzed Cascade Reaction of 2-Alkynyl Aryl Azides with Enecarbamates for Direct Synthesis of α-(3-Indolyl)ketones. Organic Letters 2023, 25 (2) , 421-425. https://doi.org/10.1021/acs.orglett.2c04147
  19. Zhantao Yang, Shenyin Hou, Yunfan Cheng, Li Sun, Chun-Hua Yang. Co-Catalyzed Reductive Cyclization of Acrylate-Containing 1,6-Enynes. The Journal of Organic Chemistry 2022, 87 (19) , 13339-13345. https://doi.org/10.1021/acs.joc.2c01345
  20. Yinchao Zhang, Wenxiu Xu, Tongtong Gao, Mengjuan Guo, Chun-Hua Yang, Hua Xie, Xiangtao Kong, Zhantao Yang, Junbiao Chang. Pd-Catalyzed Borylsilylative Cyclization of 1,6-Allenynes. Organic Letters 2022, 24 (38) , 7021-7025. https://doi.org/10.1021/acs.orglett.2c02878
  21. Lei Yu, Wenhai Li, Anyawan Tapdara, Sara Helen Kyne, Mandeep Harode, Rasool Babaahmadi, Alireza Ariafard, Philip Wai Hong Chan. Chiral Gold Complex Catalyzed Cycloisomerization/Regio- and Enantioselective Nitroso-Diels–Alder Reaction of 1,6-Diyne Esters with Nitrosobenzenes. ACS Catalysis 2022, 12 (12) , 7288-7299. https://doi.org/10.1021/acscatal.2c01680
  22. Jinhong Gao, Heng Du, Zhenhong Zhang, Qunpeng Duan, Libo Yuan, Bingchao Duan, Hongyan Yang, Kui Lu. Design and Characterization of Peptide-Based Self-Assembling Microgel for Encapsulation of Sesaminol. Foods 2025, 14 (6) , 971. https://doi.org/10.3390/foods14060971
  23. Robin Heinrich, Giovany Marie‐Rose, Christophe Gourlaouen, Patrick Pale, Eric Brenner, Aurélien Blanc. Chiral N ‐Alkylfluorenyl‐Substituted N‐Heterocyclic Carbenes in the Gold(I)‐Catalyzed Enantioselective Cycloisomerization of 1,6‐Enynes. Chemistry – A European Journal 2025, https://doi.org/10.1002/chem.202404446
  24. Zhanshuai Xiao, Yin Wei, Min Shi. Gold(I)‐Catalyzed Regioselective Cycloisomerization of Bis(Indol‐3‐Yl)‐Ynamides to Access Five‐Membered Ring Linked Bisindole Derivatives. Advanced Synthesis & Catalysis 2025, 367 (2) https://doi.org/10.1002/adsc.202401052
  25. Lipeng Li, Xiuzhi Xu, Jin Wang, Jianjun Kang, Yan Chen, Fang Ke. Heterogeneous visible-light photoredox catalysis with NiCl2@g-C3N4 for induced C(sp2)-H/N H cross-dehydrogenative coupling. Molecular Catalysis 2024, 569 , 114633. https://doi.org/10.1016/j.mcat.2024.114633
  26. Xiaoqing Guo, Xinyuan Zhang, Shaojun Hu, Lipeng Zhou, Qingfu Sun. Stereo-control on Lanthanide Triple-stranded Helicates Toward Enhanced Enantioselective Sensing. Chemical Research in Chinese Universities 2024, 40 (5) , 842-848. https://doi.org/10.1007/s40242-024-3287-2
  27. Banruo Huang, Binh Khanh Mai, Ulrike Warzok, Peng Liu, F. Dean Toste. On the gold(I)-catalyzed enantioselective addition of indole to diphenylallene via anion-binding catalysis. Tetrahedron Letters 2024, 149 , 155247. https://doi.org/10.1016/j.tetlet.2024.155247
  28. Bijin Lin, Ye Xiao, Tilong Yang, Gen-Qiang Chen, Xumu Zhang, Chi-Ming Che. Gold-catalyzed highly enantioselective cycloadditions of 1,6-enynes and 1,6-diynes assisted by remote hydrogen bonding interaction. iScience 2024, 27 (10) , 110876. https://doi.org/10.1016/j.isci.2024.110876
  29. Xiaobo He, Bendu Pan, Yaqi Zhang, Bin Chen, Long Jiang, Liqin Qiu. A Class of Chiral‐Bridged Biphenyl Monophosphine Ligands with Benzofuran Moiety and The Application in Diastereo‐ and Enantioselective Au(I)‐Catalyzed Cycloaddition. Advanced Synthesis & Catalysis 2024, 366 (14) , 3105-3110. https://doi.org/10.1002/adsc.202400433
  30. Yuan‐Yuan Zhao, Zhong‐Qiu Li, Zhong‐Liang Gong, Stefan Bernhard, Yu‐Wu Zhong. Endowing Metal‐Organic Coordination Materials with Chiroptical Activity by a Chiral Anion Strategy. Chemistry – A European Journal 2024, 30 (28) https://doi.org/10.1002/chem.202400685
  31. Chayanika Pegu, Bidisha Paroi, Nitin T. Patil. Enantioselective merged gold/organocatalysis. Chemical Communications 2024, 60 (27) , 3607-3623. https://doi.org/10.1039/D4CC00114A
  32. Amol B. Gade, Urvashi, Nitin T. Patil. Asymmetric gold catalysis enabled by specially designed ligands. Organic Chemistry Frontiers 2024, 11 (6) , 1858-1895. https://doi.org/10.1039/D4QO00057A
  33. Kaylaa Gutman, Liming Zhang. Gold-Catalyzed Enantioselective Reactions. 2024, 442-465. https://doi.org/10.1016/B978-0-32-390644-9.00095-0
  34. Andrea Cataffo, Miguel Peña‐López, Riccardo Pedrazzani, Antonio M. Echavarren. Chiral Auxiliary Approach for Gold(I)‐Catalyzed Cyclizations. Angewandte Chemie 2023, 135 (49) https://doi.org/10.1002/ange.202312874
  35. Andrea Cataffo, Miguel Peña‐López, Riccardo Pedrazzani, Antonio M. Echavarren. Chiral Auxiliary Approach for Gold(I)‐Catalyzed Cyclizations. Angewandte Chemie International Edition 2023, 62 (49) https://doi.org/10.1002/anie.202312874
  36. Ekata Saha, Ajijur Rahaman, Sukalyan Bhadra, Joyee Mitra. Exploring amine-rich supramolecular silver( i ) metallogels for autonomous self-healing and as catalysts for a three component coupling reaction. Dalton Transactions 2023, 52 (42) , 15530-15538. https://doi.org/10.1039/D3DT01654D
  37. Takumi Inoue, Moe Ota, Yui Amijima, Haru Takahashi, Shohei Hamada, Seikou Nakamura, Yusuke Kobayashi, Takahiro Sasamori, Takumi Furuta. Dual Chalcogen‐Bonding Interactions for the Conformational Control of Urea**. Chemistry – A European Journal 2023, 29 (60) https://doi.org/10.1002/chem.202302139
  38. Yuhang Yao, Jiyun Hu, Guiyu Liu, Yin-Shan Meng, Song Gao, Jun-Long Zhang. Fluoride counterions boost gold( i ) catalysis: case studies for hydrodefluorination and CO 2 hydrosilylation. Inorganic Chemistry Frontiers 2023, 10 (19) , 5573-5583. https://doi.org/10.1039/D3QI00452J
  39. Chenqi Zhang, Xinshuang Gao, Hanbo Li, Yinglu Ji, Rui Cai, Zhijian Hu, Xiaochun Wu. Regulation of Chirality Transfer and Amplification from Chiral Cysteine to Gold Nanorod Assemblies using Nonchiral Surface Ligands. Advanced Optical Materials 2023, 11 (18) https://doi.org/10.1002/adom.202202804
  40. Paul Teixeira, Stéphanie Bastin, Vincent César. Fused Polycyclic NHC Ligands in Gold Catalysis: Recent Advances. Israel Journal of Chemistry 2023, 63 (9) https://doi.org/10.1002/ijch.202200051
  41. Yufeng Wu, Hui Yang, Haojie Gao, Xiaoyi Huang, Liyuan Geng, Rui Zhang. Advances in Versatile Chiral Ligands for Asymmetric Gold Catalysis. Catalysts 2023, 13 (9) , 1294. https://doi.org/10.3390/catal13091294
  42. Jingwen Wei, Yangyang Xing, Xiaohan Ye, Bao Nguyen, Lukasz Wojtas, Xin Hong, Xiaodong Shi. Gold‐Catalyzed Amine Cascade Addition to Diyne‐Ene: Enantioselective Synthesis of 1,2‐Dihydropyridines. Angewandte Chemie 2023, 135 (31) https://doi.org/10.1002/ange.202305409
  43. Jingwen Wei, Yangyang Xing, Xiaohan Ye, Bao Nguyen, Lukasz Wojtas, Xin Hong, Xiaodong Shi. Gold‐Catalyzed Amine Cascade Addition to Diyne‐Ene: Enantioselective Synthesis of 1,2‐Dihydropyridines. Angewandte Chemie International Edition 2023, 62 (31) https://doi.org/10.1002/anie.202305409
  44. Lorenzo Carli, Anyawan Tapdara, Jianwen Jin, Yichao Zhao, Philip Wai Hong Chan. Chiral counteranion controlled chemoselectivity in gold catalysed hydroamination/enantioselective formal 1,3-allylic alcohol isomerisation of β-amino-1,4-enynols. Tetrahedron Chem 2023, 7 , 100043. https://doi.org/10.1016/j.tchem.2023.100043
  45. Chandrasekar Praveen, Sławomir Szafert. Homogeneous Gold Catalysis for Regioselective Carbocyclization of Alkynyl Precursors. ChemPlusChem 2023, 88 (7) https://doi.org/10.1002/cplu.202300202
  46. Yunliang Yu, Nazarii Sabat, Meriem Daghmoum, Zhenhao Zhang, Pascal Retailleau, Gilles Frison, Angela Marinetti, Xavier Guinchard. Enantioselective Au( i )-catalyzed tandem reactions between 2-alkynyl enones and naphthols by the tethered counterion-directed catalysis strategy. Organic Chemistry Frontiers 2023, 10 (12) , 2936-2942. https://doi.org/10.1039/D3QO00415E
  47. John M. Ovian, Petra Vojáčková, Eric N. Jacobsen. Enantioselective transition-metal catalysis via an anion-binding approach. Nature 2023, 616 (7955) , 84-89. https://doi.org/10.1038/s41586-023-05804-3
  48. Bo‐Han Zhu, Sheng‐Bing Ye, Min‐Ling Nie, Zhong‐Yang Xie, Yi‐Bo Wang, Peng‐Cheng Qian, Qing Sun, Long‐Wu Ye, Long Li. I 2 ‐Catalyzed Cycloisomerization of Ynamides: Chemoselective and Divergent Access to Indole Derivatives. Angewandte Chemie 2023, 135 (8) https://doi.org/10.1002/ange.202215616
  49. Bo‐Han Zhu, Sheng‐Bing Ye, Min‐Ling Nie, Zhong‐Yang Xie, Yi‐Bo Wang, Peng‐Cheng Qian, Qing Sun, Long‐Wu Ye, Long Li. I 2 ‐Catalyzed Cycloisomerization of Ynamides: Chemoselective and Divergent Access to Indole Derivatives. Angewandte Chemie International Edition 2023, 62 (8) https://doi.org/10.1002/anie.202215616
  50. Àlex Martí, Marc Montesinos‐Magraner, Antonio M. Echavarren, Allegra Franchino. H‐Bonded Counterion‐Directed Catalysis: Enantioselective Gold(I)‐Catalyzed Addition to 2‐Alkynyl Enones as a Case Study. European Journal of Organic Chemistry 2022, 2022 (38) https://doi.org/10.1002/ejoc.202200518
  51. Xun‐Shen Liu, Zhiqiong Tang, Zhi‐Yao Si, Zhikun Zhang, Lei Zhao, Lu Liu. Enantioselective para ‐C(sp 2 )−H Functionalization of Alkyl Benzene Derivatives via Cooperative Catalysis of Gold/Chiral Brønsted Acid**. Angewandte Chemie 2022, 134 (40) https://doi.org/10.1002/ange.202208874
  52. Xun‐Shen Liu, Zhiqiong Tang, Zhi‐Yao Si, Zhikun Zhang, Lei Zhao, Lu Liu. Enantioselective para ‐C(sp 2 )−H Functionalization of Alkyl Benzene Derivatives via Cooperative Catalysis of Gold/Chiral Brønsted Acid**. Angewandte Chemie International Edition 2022, 61 (40) https://doi.org/10.1002/anie.202208874

Journal of the American Chemical Society

Cite this: J. Am. Chem. Soc. 2022, 144, 8, 3497–3509
Click to copy citationCitation copied!
https://doi.org/10.1021/jacs.1c11978
Published February 9, 2022

Copyright © 2022 The Authors. Published by American Chemical Society. This publication is licensed under

CC-BY-NC-ND 4.0 .

Article Views

13k

Altmetric

-

Citations

Learn about these metrics

Article Views are the COUNTER-compliant sum of full text article downloads since November 2008 (both PDF and HTML) across all institutions and individuals. These metrics are regularly updated to reflect usage leading up to the last few days.

Citations are the number of other articles citing this article, calculated by Crossref and updated daily. Find more information about Crossref citation counts.

The Altmetric Attention Score is a quantitative measure of the attention that a research article has received online. Clicking on the donut icon will load a page at altmetric.com with additional details about the score and the social media presence for the given article. Find more information on the Altmetric Attention Score and how the score is calculated.

  • Abstract

    Scheme 1

    Scheme 1. Asymmetric Counterion-Directed Gold Catalysis

    Scheme 2

    Scheme 2. Design for Asymmetric H-Bonded Counterion-Directed Au(I) Catalysis

    Scheme 3

    Scheme 3. Synthesis of Phosphino(thio)urea Au(I) Chloride Complexes Au113a

    aSelected X-ray structures displayed with ORTEP ellipsoid at 50% probability level; solvent molecules and selected H atoms omitted for clarity.

    Scheme 4

    Scheme 4. Synthesis of Ag(I), Na(I), and Cu(II) Chiral Saltsa

    aX-ray structure displayed with ORTEP ellipsoid at 50% probability level; binaphthol scaffold and toluene in wireframe; selected solvent molecules and all H atoms omitted for clarity.

    Scheme 5

    Scheme 5. Enantioselective Formal [4 + 2] Cycloadditions of 1,6-Enynes Based on 5-exo-dig Cyclizationsa

    aReactions performed under Ar or N2 in anhydrous toluene (0.1 or 0.2 M), unless otherwise stated. Yields of material isolated after purification, er determined by HPLC or SFC on chiral stationary phase.

    bCarried out at 2 mmol scale, with 1 mol % Au10 and 1 mol % Ag6, in technical-grade toluene (0.6 M) under air for 24 h.

    cAt 23 °C.

    dAt 0 °C.

    eWith 10 mol % Au10 and 10 mol % Ag6.

    fIncluding 5% of inseparable 6-endo-isomer.

    gReaction time: 96 h.

    Scheme 6

    Scheme 6. Enantioselective 6-endo-dig Cyclizations of 1,6-Enynes without (A) or with (B) Nucleophile Additiona

    aYields of material isolated after purification, er determined by HPLC or SFC on chiral stationary phase.

    bFor 48 h.

    cAt 23 °C.

    Scheme 7

    Scheme 7. Derivatization and Scope Extension

    aPrepared from a batch of 6f with 93:7 er (see Supporting Information).

    Scheme 8

    Scheme 8. Enantioselective Cycloisomerization–Indole Addition to 2-Alkynyl Enonesa

    aYields of material isolated after purification, er determined by HPLC or SFC on chiral stationary phase.

    Scheme 9

    Scheme 9. Synthesis of Complexes with Ureaphosphine L10

    Scheme 10

    Scheme 10. Mechanistic Investigations: (A) Enantiofacial Selectivity, (B) Study of Nonlinear Effects, (C) Use of Complex Au4 as Au10 Surrogate, (D) 1H NMR Titration of Au4 with (R)-Na6 (298 K, CD2Cl2), (E–J) Kinetic Studies

    Scheme 11

    Scheme 11. Proposed Mechanism for the Cyclization of Enyne 1 under H-Bonded Counterion-Directed Au(I) Catalysis
  • References


    This article references 52 other publications.

    1. 1
      Hamilton, G. L.; Kang, E. J.; Mba, M.; Toste, F. D. A Powerful Chiral Counterion Strategy for Asymmetric Transition Metal Catalysis. Science 2007, 317, 496499,  DOI: 10.1126/science.1145229
    2. 2
      (a) Mayer, S.; List, B. Asymmetric Counteranion-Directed Catalysis. Angew. Chem. Int. Ed. 2006, 45, 41934195,  DOI: 10.1002/anie.200600512
      (b) Mukherjee, S.; List, B. Chiral Counteranions in Asymmetric Transition-Metal Catalysis: Highly Enantioselective Pd/Brønsted Acid-Catalyzed Direct α-Allylation of Aldehydes. J. Am. Chem. Soc. 2007, 129, 1133611337,  DOI: 10.1021/ja074678r
      (c) Phipps, R. J.; Hamilton, G. L.; Toste, F. D. The Progression of Chiral Anions from Concepts to Applications in Asymmetric Catalysis. Nat. Chem. 2012, 4, 603614,  DOI: 10.1038/nchem.1405
      (d) Mahlau, M.; List, B. Asymmetric Counteranion-Directed Catalysis: Concept, Definition, and Applications. Angew. Chem., Int. Ed. 2013, 52, 518533,  DOI: 10.1002/anie.201205343
      (e) Brak, K.; Jacobsen, E. N. Asymmetric Ion-Pairing Catalysis. Angew. Chem., Int. Ed. 2013, 52, 534561,  DOI: 10.1002/anie.201205449
      (f) Shirakawa, S.; Maruoka, K. Recent Developments in Asymmetric Phase-Transfer Reactions. Angew. Chem., Int. Ed. 2013, 52, 43124348,  DOI: 10.1002/anie.201206835
    3. 3

      For work on the combination of achiral Au(I) complexes and chiral anions, see the following:

      (a) Reference (1).
      (b) LaLonde, R. L.; Wang, Z. J.; Mba, M.; Lackner, A. D.; Toste, F. D. Gold(I)-Catalyzed Enantioselective Synthesis of Pyrazolidines, Isoxazolidines, and Tetrahydrooxazines. Angew. Chem., Int. Ed. 2010, 49, 598601,  DOI: 10.1002/anie.200905000
      (c) Zi, W.; Toste, F. D. Gold(I)-Catalyzed Enantioselective Desymmetrization of 1,3-Diols through Intramolecular Hydroalkoxylation of Allenes. Angew. Chem., Int. Ed. 2015, 54, 1444714451,  DOI: 10.1002/anie.201508331
      (d) Pedrazzani, R.; An, J.; Monari, M.; Bandini, M. New Chiral BINOL-Based Phosphates for Enantioselective [Au(I)]-Catalyzed Dearomatization of β-Naphthols with Allenamides. Eur. J. Org. Chem. 2021, 2021, 17321736,  DOI: 10.1002/ejoc.202100166
    4. 4

      For work on the combination of chiral Au(I) complexes and chiral anions, see the following:

      (a) Aikawa, K.; Kojima, M.; Mikami, K. Axial Chirality Control of Gold(biphep) Complexes by Chiral Anions: Application to Asymmetric Catalysis. Angew. Chem., Int. Ed. 2009, 48, 60736077,  DOI: 10.1002/anie.200902084
      (b) Aikawa, K.; Kojima, M.; Mikami, K. Synergistic Effect: Hydroalkoxylation of Allenes through Combination of Enantiopure BIPHEP-Gold Complexes and Chiral Anions. Adv. Synth. Catal. 2010, 352, 31313135,  DOI: 10.1002/adsc.201000672
      Corrigendum:Aikawa, K.; Kojima, M.; Mikami, K. Adv. Synth. Catal. 2011, 353, 28822883,  DOI: 10.1002/adsc.201100838
      (c) Barreiro, E. M.; Broggini, D. F. D.; Adrio, L. A.; White, A. J. P.; Schwenk, R.; Togni, A.; Hii, K. K. Gold(I) Complexes of Conformationally Constricted Chiral Ferrocenyl Phosphines. Organometallics 2012, 31, 37453754,  DOI: 10.1021/om300222k
      (d) Miles, D. H.; Veguillas, M.; Toste, F. D. Gold(I)-Catalyzed Enantioselective Bromocyclization Reactions of Allenes. Chem. Sci. 2013, 4, 34273431,  DOI: 10.1039/c3sc50811k
      (e) Handa, S.; Lippincott, D. J.; Aue, D. H.; Lipshutz, B. H. Asymmetric Gold-Catalyzed Lactonizations in Water at Room Temperature. Angew. Chem., Int. Ed. 2014, 53, 1065810662,  DOI: 10.1002/anie.201404729
    5. 5
      (a) Mourad, A. K.; Leutzow, J.; Czekelius, C. Anion-Induced Enantioselective Cyclization of Diynamides to Pyrrolidines Catalyzed by Cationic Gold Complexes. Angew. Chem., Int. Ed. 2012, 51, 1114911152,  DOI: 10.1002/anie.201205416
      (b) Spittler, M.; Lutsenko, K.; Czekelius, C. Total Synthesis of (+)-Mesembrine Applying Asymmetric Gold Catalysis. J. Org. Chem. 2016, 81, 61006105,  DOI: 10.1021/acs.joc.6b00985
    6. 6

      For a review on the combination of Au(I) methyl complexes and excess Brønsted acid for tandem reactions, see the following:

      (a) Inamdar, S. M.; Konala, A.; Patil, N. T. When gold meets chiral Brønsted acid catalysis: extending the boundaries of enantioselective gold catalysis. Chem. Commun. 2014, 50, 1512415135,  DOI: 10.1039/C4CC04633A

      For selected examples, see the following:

      (b) Han, Z.-Y.; Xiao, H.; Chen, X.-H.; Gong, L.-Z. Consecutive Intramolecular Hydroamination/Asymmetric Transfer Hydrogenation under Relay Catalysis of an Achiral Gold Complex/Chiral Brønsted Acid Binary System. J. Am. Chem. Soc. 2009, 131, 91829183,  DOI: 10.1021/ja903547q
      (c) Muratore, M. E.; Holloway, C. A.; Pilling, A. W.; Storer, R. I.; Trevitt, G.; Dixon, D. J. Enantioselective Brønsted Acid-Catalyzed N-Acyliminium Cyclization Cascades. J. Am. Chem. Soc. 2009, 131, 1079610797,  DOI: 10.1021/ja9024885
      (d) Liu, X.-Y.; Che, C.-M. Highly Enantioselective Synthesis of Chiral Secondary Amines by Gold(I)/Chiral Brønsted Acid Catalyzed Tandem Intermolecular Hydroamination and Transfer Hydrogenation Reactions. Org. Lett. 2009, 11, 42044207,  DOI: 10.1021/ol901443b
      (e) Tu, X.-F.; Gong, L.-Z. Highly Enantioselective Transfer Hydrogenation of Quinolines Catalyzed by Gold Phosphates: Achiral Ligand Tuning and Chiral-Anion Control of Stereoselectivity. Angew. Chem., Int. Ed. 2012, 51, 1134611349,  DOI: 10.1002/anie.201204179
      (f) Cala, L.; Mendoza, A.; Fañanás, F. J.; Rodríguez, F. A Catalytic Multicomponent Coupling Reaction for the Enantioselective Synthesis of Spiroacetals. Chem. Commun. 2013, 49, 27152717,  DOI: 10.1039/c3cc00118k
      (g) Zhou, S.; Li, Y.; Liu, X.; Hu, W.; Ke, Z.; Xu, X. Enantioselective Oxidative Multi-Functionalization of Terminal Alkynes with Nitrones and Alcohols for Expeditious Assembly of Chiral α-Alkoxy-β-amino-ketones. J. Am. Chem. Soc. 2021, 143, 1470314711,  DOI: 10.1021/jacs.1c06178
    7. 7
      (a) Dorel, R.; Echavarren, A. M. Gold(I)-Catalyzed Activation of Alkynes for the Construction of Molecular Complexity. Chem. Rev. 2015, 115, 90289072,  DOI: 10.1021/cr500691k
      (b) Zuccarello, G.; Escofet, I.; Caniparoli, U.; Echavarren, A. M. New-Generation Ligand Design for the Gold Catalyzed Asymmetric Activation of Alkynes. ChemPlusChem 2021, 86, 12831296,  DOI: 10.1002/cplu.202100232
      (c) Barbazanges, M.; Augé, M.; Moussa, J.; Amouri, H.; Aubert, C.; Desmarets, C.; Fensterbank, L.; Gandon, V.; Malacria, M.; Ollivier, C. Enantioselective IrI-Catalyzed Carbocyclization of 1,6-Enynes by the Chiral Counterion Strategy. Chem.─Eur. J. 2011, 17, 1378913794,  DOI: 10.1002/chem.201102723
    8. 8
      Raducan, M.; Moreno, M.; Bour, C.; Echavarren, A. M. Phosphate Ligands in the Gold(I)-Catalysed Activation of Enynes. Chem. Commun. 2012, 48, 5254,  DOI: 10.1039/C1CC15739F
    9. 9

      For reviews, see the following:

      (a) Jia, M.; Bandini, M. Counterion Effects in Homogeneous Gold Catalysis. ACS Catal. 2015, 5, 16381652,  DOI: 10.1021/cs501902v
      (b) Zuccaccia, D.; Del Zotto, A.; Baratta, W. The Pivotal Role of the Counterion in Gold Catalyzed Hydration and Alkoxylation of Alkynes. Coord. Chem. Rev. 2019, 396, 103116,  DOI: 10.1016/j.ccr.2019.06.007
      (c) Lu, Z.; Li, T.; Mudshinge, S. R.; Xu, B.; Hammond, G. B. Optimization of Catalysts and Conditions in Gold(I) Catalysis─Counterion and Additive Effects. Chem. Rev. 2021, 121, 84528477,  DOI: 10.1021/acs.chemrev.0c00713
    10. 10

      For selected articles, see the following:

      (a) Zuccaccia, D.; Belpassi, L.; Tarantelli, F.; Macchioni, A. Ion Pairing in Cationic Olefin-Gold(I) Complexes. J. Am. Chem. Soc. 2009, 131, 31703171,  DOI: 10.1021/ja809998y
      (b) Zuccaccia, D.; Belpassi, L.; Rocchigiani, L.; Tarantelli, F.; Macchioni, A. A Phosphine Gold(I) π-Alkyne Complex: Tuning the Metal-Alkyne Bond Character and Counterion Position by the Choice of the Ancillary Ligand. Inorg. Chem. 2010, 49, 30803082,  DOI: 10.1021/ic100093n
      (c) Bandini, M.; Bottoni, A.; Chiarucci, M.; Cera, G.; Miscione, G. P. Mechanistic Insights into Enantioselective Gold-Catalyzed Allylation of Indoles with Alcohols: The Counterion Effect. J. Am. Chem. Soc. 2012, 134, 2069020700,  DOI: 10.1021/ja3086774
      (d) Zhdanko, A.; Maier, M. E. Explanation of Counterion Effects in Gold(I)-Catalyzed Hydroalkoxylation of Alkynes. ACS Catal. 2014, 4, 27702775,  DOI: 10.1021/cs500446d
      (e) Biasiolo, L.; Del Zotto, A.; Zuccaccia, D. Toward Optimizing the Performance of Homogeneous L-Au-X Catalysts through Appropriate Matching of the Ligand (L) and Counterion (X). Organometallics 2015, 34, 17591765,  DOI: 10.1021/acs.organomet.5b00308
      (f) Lu, Z.; Han, J.; Okoromoba, O. E.; Shimizu, N.; Amii, H.; Tormena, C. F.; Hammond, G. B.; Xu, B. Predicting Counterion Effects Using a Gold Affinity Index and a Hydrogen Bonding Basicity Index. Org. Lett. 2017, 19, 58485851,  DOI: 10.1021/acs.orglett.7b02829
      (g) Schießl, J.; Schulmeister, J.; Doppiu, A.; Wörner, E.; Rudolph, M.; Karch, R.; Hashmi, A. S. K. An Industrial Perspective on Counter Anions in Gold Catalysis: Underestimated with Respect to “Ligand Effects”. Adv. Synth. Catal. 2018, 360, 24932502,  DOI: 10.1002/adsc.201800233
    11. 11
      (a) Neel, A. J.; Hilton, M. J.; Sigman, M. S.; Toste, F. D. Exploiting Non-covalent π Interactions for Catalyst Design. Nature 2017, 543, 637646,  DOI: 10.1038/nature21701
      (b) Fanourakis, A.; Docherty, P. J.; Chuentragool, P.; Phipps, R. J. Recent Developments in Enantioselective Transition Metal Catalysis Featuring Attractive Noncovalent Interactions between Ligand and Substrate. ACS Catal. 2020, 10, 1067210714,  DOI: 10.1021/acscatal.0c02957
    12. 12
      (a) Duarte, F.; Paton, R. S. Molecular Recognition in Asymmetric Counteranion Catalysis: Understanding Chiral Phosphate-Mediated Desymmetrization. J. Am. Chem. Soc. 2017, 139, 88868896,  DOI: 10.1021/jacs.7b02468
      (b) Orlandi, M.; Coelho, J. A. S.; Hilton, M. J.; Toste, F. D.; Sigman, M. S. Parametrization of Non-covalent Interactions for Transition State Interrogation Applied to Asymmetric Catalysis. J. Am. Chem. Soc. 2017, 139, 68036806,  DOI: 10.1021/jacs.7b02311
      (c) Shoja, A.; Reid, J. P. Computational Insights into Privileged Stereocontrolling Interactions Involving Chiral Phosphates and Iminium Intermediates. J. Am. Chem. Soc. 2021, 143, 72097215,  DOI: 10.1021/jacs.1c03829
    13. 13
      Zhang, Z.; Smal, V.; Retailleau, P.; Voituriez, A.; Frison, G.; Marinetti, A.; Guinchard, X. Tethered Counterion-Directed Catalysis: Merging the Chiral Ion-Pairing and Bifunctional Ligand Strategies in Enantioselective Gold(I) Catalysis. J. Am. Chem. Soc. 2020, 142, 37973805,  DOI: 10.1021/jacs.9b11154
    14. 14
      Michelet, V. Noble Metal-Catalyzed Enyne Cycloisomerizations and Related Reactions. Comprehensive Organic Synthesis, 2nd ed.; Elsevier, 2014; Vol. 5, pp 14831536.
    15. 15

      For recent reviews on asymmetric gold catalysis, see the following:

      (a) Zi, W.; Toste, F. D. Recent Advances in Enantioselective Gold Catalysis. Chem. Soc. Rev. 2016, 45, 45674589,  DOI: 10.1039/C5CS00929D
      (b) Li, Y.; Li, W.; Zhang, J. Gold-Catalyzed Enantioselective Annulations. Chem.─Eur. J. 2017, 23, 467512,  DOI: 10.1002/chem.201602822
      (c) Jiang, J.-J.; Wong, M.-K. Recent Advances in the Development of Chiral Gold Complexes for Catalytic Asymmetric Catalysis. Chem.─Asian J. 2021, 16, 364377,  DOI: 10.1002/asia.202001375
      (d) Porcel García, S. Gold Catalyzed Asymmetric Transformations. IntechOpen, 2021;  DOI: 10.5772/intechopen.97519 .
      (e) Reference (7b).
    16. 16
      (a) Cheng, X.; Zhang, L. Designed Bifunctional Ligands in Cooperative Homogeneous Gold Catalysis. CCS Chem. 2021, 3, 19892002,  DOI: 10.31635/ccschem.020.202000454
      (b) Zuccarello, G.; Zanini, M.; Echavarren, A. M. Buchwald-Type Ligands on Gold(I) Catalysis. Isr. J. Chem. 2020, 60, 360372,  DOI: 10.1002/ijch.201900179
    17. 17
      (a) Schreiner, P. R. Metal-Free Organocatalysis through Explicit Hydrogen Bonding Interactions. Chem. Soc. Rev. 2003, 32, 289296,  DOI: 10.1039/b107298f
      (b) Doyle, A. G.; Jacobsen, E. N. Small-Molecule H-Bond Donors in Asymmetric Catalysis. Chem. Rev. 2007, 107, 57135743,  DOI: 10.1021/cr068373r
    18. 18

      For the only previous examples of phosphino(thio)urea Au(I) complexes, see the following:

      (a) Campbell, M. J.; Toste, F. D. Enantioselective Synthesis of Cyclic Carbamimidates via a Three-Component Reaction of Imines, Terminal Alkynes, and p-Toluenesulfonylisocyanate Using a Monophosphine Gold(I) Catalyst. Chem. Sci. 2011, 2, 13691378,  DOI: 10.1039/c1sc00160d
      (b) Franchino, A.; Martí, À.; Nejrotti, S.; Echavarren, A. M. Silver-Free Au(I) Catalysis Enabled by Bifunctional Urea- and Squaramide-Phosphine Ligands via H-Bonding. Chem.─Eur. J. 2021, 27, 1198911996,  DOI: 10.1002/chem.202101751
    19. 19
      (a) Akiyama, T. Stronger Brønsted Acids. Chem. Rev. 2007, 107, 57445758,  DOI: 10.1021/cr068374j
      (b) Terada, M. Binaphthol-derived Phosphoric Acid as a Versatile Catalyst for Enantioselective Carbon-Carbon Bond Forming Reactions. Chem. Commun. 2008, 40974112,  DOI: 10.1039/b807577h
      (c) Akiyama, T.; Mori, K. Stronger Brønsted Acids: Recent Progress. Chem. Rev. 2015, 115, 92779306,  DOI: 10.1021/acs.chemrev.5b00041
      (d) Parmar, D.; Sugiono, E.; Raja, S.; Rueping, M. Complete Field Guide to Asymmetric BINOL-Phosphate Derived Brønsted Acid and Metal Catalysis: History and Classification by Mode of Activation; Brønsted Acidity, Hydrogen Bonding, Ion Pairing, and Metal Phosphates. Chem. Rev. 2014, 114, 90479153,  DOI: 10.1021/cr5001496
      Correction:Parmar, D.; Sugiono, E.; Raja, S.; Rueping, M. Chem. Rev. 2017, 117, 1060810620,  DOI: 10.1021/acs.chemrev.7b00197
    20. 20
      Wang, Y.; Wang, Z.; Li, Y.; Wu, G.; Cao, Z.; Zhang, L. A General Ligand Design for Gold Catalysis Allowing Ligand-Directed anti-Nucleophilic Attack of Alkynes. Nat. Commun. 2014, 5, 3470,  DOI: 10.1038/ncomms4470
    21. 21
      (a) Zhang, Z.; Schreiner, P. R. (Thio)urea Organocatalysis─What Can Be Learnt from Anion Recognition?. Chem. Soc. Rev. 2009, 38, 11871198,  DOI: 10.1039/b801793j
      (b) Amendola, V.; Fabbrizzi, L.; Mosca, L. Anion Recognition by Hydrogen Bonding: Urea-Based Receptors. Chem. Soc. Rev. 2010, 39, 38893915,  DOI: 10.1039/b822552b
      (c) Bregović, V. B.; Basarić, N.; Mlinarić-Majerski, K. Anion Binding with Urea and Thiourea Derivatives. Coord. Chem. Rev. 2015, 295, 80124,  DOI: 10.1016/j.ccr.2015.03.011
    22. 22

      For reviews on supramolecular ligands for asymmetric metal catalysis, see the following:

      (a) Meeuwissen, J.; Reek, J. N. H. Supramolecular Catalysis Beyond Enzyme Mimics. Nat. Chem. 2010, 2, 615621,  DOI: 10.1038/nchem.744
      (b) Carboni, S.; Gennari, C.; Pignataro, L.; Piarulli, U. Supramolecular Ligand-Ligand and Ligand-Substrate Interactions for Highly Selective Transition Metal Catalysis. Dalton Trans. 2011, 40, 43554373,  DOI: 10.1039/c0dt01517b
      (c) Bellini, R.; van der Vlugt, J. I.; Reek, J. N. H. Supramolecular Self-Assembled Ligands in Asymmetric Transition Metal Catalysis. Isr. J. Chem. 2012, 52, 613629,  DOI: 10.1002/ijch.201200002
      (d) Raynal, M.; Ballester, P.; Vidal-Ferran, A.; van Leeuwen, P. W. N. M. Supramolecular Catalysis. Part 1: Non-Covalent Interactions as a Tool for Building and Modifying Homogeneous Catalysts. Chem. Soc. Rev. 2014, 43, 16601733,  DOI: 10.1039/C3CS60027K
      (e) Ohmatsu, K.; Ooi, T. Design of Supramolecular Chiral Ligands for Asymmetric Metal Catalysis. Tetrahedron Lett. 2015, 56, 20432048,  DOI: 10.1016/j.tetlet.2015.02.096
      (f) Trouvé, J.; Gramage-Doria, R. Beyond Hydrogen Bonding: Recent Trends of Outer Sphere Interactions in Transition Metal Catalysis. Chem. Soc. Rev. 2021, 50, 35653584,  DOI: 10.1039/D0CS01339K

      For a seminal example involving ion-pairing interactions between an achiral ligand and a chiral anion, see the following:

      (g) Ohmatsu, K.; Ito, M.; Kunieda, T.; Ooi, T. Ion-Paired Chiral Ligands for Asymmetric Palladium Catalysis. Nat. Chem. 2012, 4, 473477,  DOI: 10.1038/nchem.1311
    23. 23
      Murata, M.; Buchwald, S. L. A General and Efficient Method for the Palladium-Catalyzed Cross-Coupling of Thiols and Secondary Phosphines. Tetrahedron 2004, 60, 73977403,  DOI: 10.1016/j.tet.2004.05.044
    24. 24
      Luchini, G.; Ascough, D. M. H.; Alegre-Requena, J. V.; Gouverneur, V.; Paton, R. S. Data-Mining the Diaryl(thio)urea Conformational Landscape: Understanding the Contrasting Behavior of Ureas and Thioureas with Quantum Chemistry. Tetrahedron 2019, 75, 697702,  DOI: 10.1016/j.tet.2018.12.033
    25. 25
      Nakashima, D.; Yamamoto, H. Design of Chiral N-Triflyl Phosphoramide as a Strong Chiral Brønsted Acid and Its Application to Asymmetric Diels-Alder Reaction. J. Am. Chem. Soc. 2006, 128, 96269627,  DOI: 10.1021/ja062508t
    26. 26
      (a) Nieto-Oberhuber, C.; López, S.; Echavarren, A. M. Intramolecular [4 + 2] Cycloadditions of 1,3-Enynes or Arylalkynes with Alkenes with Highly Reactive Cationic Phosphine Au(I) Complexes. J. Am. Chem. Soc. 2005, 127, 61786179,  DOI: 10.1021/ja042257t
      (b) Nieto-Oberhuber, C.; Pérez-Galán, P.; Herrero-Gómez, E.; Lauterbach, T.; Rodríguez, C.; López, S.; Bour, C.; Rosellón, A.; Cárdenas, D. J.; Echavarren, A. M. Gold(I)-Catalyzed Intramolecular [4 + 2] Cycloadditions of Arylalkynes or 1,3-Enynes with Alkenes: Scope and Mechanism. J. Am. Chem. Soc. 2008, 130, 269279,  DOI: 10.1021/ja075794x
    27. 27

      Previous asymmetric versions, all based on chiral ligands for Au(I), lack either scope or high enantiocontrol:

      (a) Two examples only (44–99% yield, 92–93% ee):Chao, C.-M.; Vitale, M. R.; Toullec, P. Y.; Genêt, J.-P.; Michelet, V. Asymmetric Gold-Catalyzed Hydroarylation/Cyclization Reactions. Chem.─Eur. J. 2009, 15, 13191323,  DOI: 10.1002/chem.200802341
      (b) One further example (86% yield, 16% ee):Pradal, A.; Chao, C.-M.; Vitale, M. R.; Toullec, P. Y.; Michelet, V. Asymmetric Au-Catalyzed Domino Cyclization/Nucleophile Addition Reactions of Enynes in the Presence of Water, Methanol and Electron-rich Aromatic Derivatives. Tetrahedron 2011, 67, 43714377,  DOI: 10.1016/j.tet.2011.03.071
      (c) Five examples (70–95% yield, 73–88% ee):Delpont, N.; Escofet, I.; Pérez-Galán, P.; Spiegl, D.; Raducan, M.; Bour, C.; Sinisi, R.; Echavarren, A. M. Modular Chiral Gold(I) Phosphite Complexes. Catal. Sci. Technol. 2013, 3, 30073012,  DOI: 10.1039/c3cy00250k
      (d) One example only (99% yield, 91% ee):Aillard, P.; Dova, D.; Magné, V.; Retailleau, P.; Cauteruccio, S.; Licandro, E.; Voituriez, A.; Marinetti, A. The Synthesis of Substituted Phosphathiahelicenes via Regioselective Bromination of a Preformed Helical Scaffold: A New Approach to Modular Ligands for Enantioselective Gold-Catalysis. Chem. Commun. 2016, 52, 1098410987,  DOI: 10.1039/C6CC04765C
      (e) 17 examples (61–99% yield, 58–92% ee):Zuccarello, G.; Mayans, J. G.; Escofet, I.; Scharnagel, D.; Kirillova, M. S.; Pérez-Jimeno, A. H.; Calleja, P.; Boothe, J. R.; Echavarren, A. M. Enantioselective Folding of Enynes by Gold(I) Catalysts with a Remote C2-Chiral Element. J. Am. Chem. Soc. 2019, 141, 1185811863,  DOI: 10.1021/jacs.9b06326
      (f) One example only (92% yield, 94% ee):Magné, V.; Sanogo, Y.; Demmer, C. S.; Retailleau, P.; Marinetti, A.; Guinchard, X.; Voituriez, A. Chiral Phosphathiahelicenes: Improved Synthetic Approach and Uses in Enantioselective Gold(I)-Catalyzed [2 + 2] Cycloadditions of N-Homoallenyl Tryptamines. ACS Catal. 2020, 10, 81418148,  DOI: 10.1021/acscatal.0c01819
    28. 28

      N-Triflyl phosphoric amide D1 has a pKa value 0.9 units lower than the corresponding phosphoric acid TRIP-H (4.2 vs 3.3, both values in DMSO at 25 °C):

      Christ, P.; Lindsay, A. G.; Vormittag, S. S.; Neudörfl, J.-M.; Berkessel, A.; O’Donoghue, A. C. pKa Values of Chiral Brønsted Acid Catalysts: Phosphoric Acids/Amides, Sulfonyl/Sulfuryl Imides, and Perfluorinated TADDOLs (TEFDDOLs). Chem.─Eur. J. 2011, 17, 85248528,  DOI: 10.1002/chem.201101157
    29. 29
      Raubenheimer, H. G.; Schmidbaur, H. Gold Chemistry Guided by the Isolobality Concept. Organometallics 2012, 31, 25072522,  DOI: 10.1021/om2010113
    30. 30
      Reid, J. P.; Goodman, J. M. Goldilocks Catalysts: Computational Insights into the Role of the 3,3′ Substituents on the Selectivity of BINOL-Derived Phosphoric Acid Catalysts. J. Am. Chem. Soc. 2016, 138, 79107917,  DOI: 10.1021/jacs.6b02825
    31. 31
      (a) Kepp, K. P. A Quantitative Scale of Oxophilicity and Thiophilicity. Inorg. Chem. 2016, 55, 94619470,  DOI: 10.1021/acs.inorgchem.6b01702
      (b) Izaga, A.; Herrera, R. P.; Gimeno, M. C. Gold(I)-Mediated Thiourea Organocatalyst Activation: A Synergic Effect for Asymmetric Catalysis. ChemCatChem 2017, 9, 13131321,  DOI: 10.1002/cctc.201601527
    32. 32
      Reichardt, C. Solvatochromic Dyes as Solvent Polarity Indicators. Chem. Rev. 1994, 94, 23192358,  DOI: 10.1021/cr00032a005
    33. 33

      All gold complexes used so far for the enantioselective cyclization of O-tethered 1,6-enynes bear chiral ligands. For a complete overview, see the following:

      (a) Mato, M.; Franchino, A.; García-Morales, C.; Echavarren, A. M. Gold-Catalyzed Synthesis of Small Rings. Chem. Rev. 2021, 121, 86138684,  DOI: 10.1021/acs.chemrev.0c00697

      For seminal examples, see the following:

      (b) Chao, C.-M.; Beltrami, D.; Toullec, P. Y.; Michelet, V. Asymmetric Au(I)-Catalyzed Synthesis of Bicyclo[4.1.0]heptene Derivatives via a Cycloisomerization Process of 1,6-Enynes. Chem. Commun. 2009, 69886990,  DOI: 10.1039/b913554e
      (c) Teller, H.; Corbet, M.; Mantilli, L.; Gopakumar, G.; Goddard, R.; Thiel, W.; Fürstner, A. One-Point Binding Ligands for Asymmetric Gold Catalysis: Phosphoramidites with a TADDOL-Related but Acyclic Backbone. J. Am. Chem. Soc. 2012, 134, 1533115342,  DOI: 10.1021/ja303641p
    34. 34
      Sanjuán, A. M.; Martínez, A.; García-García, P.; Fernández-Rodríguez, M. A.; Sanz, R. Gold(I)-catalyzed 6-endo hydroxycyclization of 7-substituted-1,6-enynes. Beilstein J. Org. Chem. 2013, 9, 22422249,  DOI: 10.3762/bjoc.9.263
    35. 35

      See Supporting Information for discussion on other nucleophiles and reaction side products.

    36. 36
      (a) Pfeifer, L.; Engle, K. M.; Pidgeon, G. W.; Sparkes, H. A.; Thompson, A. L.; Brown, J. M.; Gouverneur, V. Hydrogen-Bonded Homoleptic Fluoride-Diarylurea Complexes: Structure, Reactivity, and Coordinating Power. J. Am. Chem. Soc. 2016, 138, 1331413325,  DOI: 10.1021/jacs.6b07501
      (b) Ibba, F.; Pupo, G.; Thompson, A. L.; Brown, J. M.; Claridge, T. D. W.; Gouverneur, V. Impact of Multiple Hydrogen Bonds with Fluoride on Catalysis: Insight from NMR Spectroscopy. J. Am. Chem. Soc. 2020, 142, 1973119744,  DOI: 10.1021/jacs.0c09832
    37. 37
      D’Abrosca, B.; Fiorentino, A.; Golino, A.; Monaco, P.; Oriano, P.; Pacifico, S. Carexanes: Prenyl Stilbenoid Derivatives from Carex distachya. Tetrahedron Lett. 2005, 46, 52695272,  DOI: 10.1016/j.tetlet.2005.06.036
    38. 38
      Martín-Torres, I.; Ogalla, G.; Yang, J.-M.; Rinaldi, A.; Echavarren, A. M. Enantioselective Alkoxycyclization of 1,6-Enynes with Gold(I)-Cavitands: Total Synthesis of Mafaicheenamine C. Angew. Chem., Int. Ed. 2021, 60, 93399344,  DOI: 10.1002/anie.202017035
    39. 39
      Witham, C. A.; Mauleón, P.; Shapiro, N. D.; Sherry, B. D.; Toste, F. D. Gold(I)-Catalyzed Oxidative Rearrangements. J. Am. Chem. Soc. 2007, 129, 58385839,  DOI: 10.1021/ja071231+
    40. 40
      Wang, W.; Yang, J.; Wang, F.; Shi, M. Axially Chiral N-Heterocyclic Carbene Gold(I) Complex Catalyzed Asymmetric Cycloisomerization of 1,6-Enynes. Organometallics 2011, 30, 38593869,  DOI: 10.1021/om2004404
    41. 41
      (a) Yao, T.; Zhang, X.; Larock, R. C. AuCl3-Catalyzed Synthesis of Highly Substituted Furans from 2-(1-Alkynyl)-2-alken-1-ones. J. Am. Chem. Soc. 2004, 126, 1116411165,  DOI: 10.1021/ja0466964
      (b) Rauniyar, V.; Wang, Z. J.; Burks, H. E.; Toste, F. D. Enantioselective Synthesis of Highly Substituted Furans by a Copper(II)-Catalyzed Cycloisomerization-Indole Addition Reaction. J. Am. Chem. Soc. 2011, 133, 84868489,  DOI: 10.1021/ja202959n
      (c) Force, G.; Ki, Y. L. T.; Isaac, K.; Retailleau, P.; Marinetti, A.; Betzer, J.-F. Paracyclophane-based Silver Phosphates as Catalysts for Enantioselective Cycloisomerization/Addition Reactions: Synthesis of Bicyclic Furans. Adv. Synth. Catal. 2018, 360, 33563366,  DOI: 10.1002/adsc.201800587
    42. 42

      31P{1H} and 19F{1H} NMR spectra remained unchanged, confirming that the chloride was not scavenged. See Supporting Information for details.

    43. 43
      (a) Burés, J. A Simple Graphical Method to Determine the Order in Catalyst. Angew. Chem., Int. Ed. 2016, 55, 20282031,  DOI: 10.1002/anie.201508983
      (b) Burés, J. Variable Time Normalization Analysis: General Graphical Elucidation of Reaction Orders from Concentration Profiles. Angew. Chem., Int. Ed. 2016, 55, 1608416087,  DOI: 10.1002/anie.201609757
      (c) Nielsen, C. D.-T.; Burés, J. Visual Kinetic Analysis. Chem. Sci. 2019, 10, 348353,  DOI: 10.1039/C8SC04698K
    44. 44
      Briggs, G. E.; Haldane, J. B. S. A Note on the Kinetics of Enzyme Action. Biochem. J. 1925, 19, 338339,  DOI: 10.1042/bj0190338

      Application of this model rests on a series of underlying assumptions, all satisfied in the present case (see Supporting Information).

    45. 45
      (a) Michaelis, L.; Menten, M. L. Die Kinetik der Invertinwirkung. Biochem. Z. 1913, 49, 333369
      (b) Johnson, K. A.; Goody, R. S. The Original Michaelis Constant: Translation of the 1913 Michaelis−Menten Paper. Biochemistry 2011, 50, 82648269,  DOI: 10.1021/bi201284u
    46. 46
      Blackmond, D. G. Reaction Progress Kinetic Analysis: A Powerful Methodology for Mechanistic Studies of Complex Catalytic Reactions. Angew. Chem., Int. Ed. 2005, 44, 43024320,  DOI: 10.1002/anie.200462544
      Correction:Blackmond, D. G. Angew. Chem., Int. Ed. 2006, 45, 21622162,  DOI: 10.1002/anie.200690050
    47. 47
      Franchino, A.; Montesinos-Magraner, M.; Echavarren, A. M. Silver-Free Catalysis with Gold(I) Chloride Complexes. Bull. Chem. Soc. Jpn. 2021, 94, 10991117,  DOI: 10.1246/bcsj.20200358
    48. 48
      Lineweaver, H.; Burk, D. The Determination of Enzyme Dissociation Constants. J. Am. Chem. Soc. 1934, 56, 658666,  DOI: 10.1021/ja01318a036
    49. 49

      The uncertainty refers not to experimental variation but only to the mathematical error of the linear regression, as determined by Excel LINEST routine. If all data from time 0 to 10 h are used in the Lineawer–Burk plot, a KM value of 63 ± 4 is obtained, leading to identical conclusions regarding the partial order in substrate (see Supporting Information).

    50. 50
      Burés, J. What is the Order of a Reaction?. Top. Catal. 2017, 60, 631633,  DOI: 10.1007/s11244-017-0735-y
    51. 51

      Geometry optimizations were carried out using Gaussian 09 at the B3LYP-D3/6-31G(d,p) level of theory in toluene (SMD). Single point energy calculations were performed on the resulting structures employing B3LYP-D3/6-311+G(d,p)/SMD(toluene). See the Supporting Information for an overview of all computed structures. Refer to ref (12c) for an excellent discussion of alternative computational methods (functionals, basis sets, solvent models) in the context of chiral phosphate–iminium ion pairs.

    52. 52

      This step is assumed to be enantiodetermining on the basis of previous DFT calculations for the same reactions (refs (26) and (27e)).

  • Supporting Information

    Supporting Information


    The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/jacs.1c11978.

    • Optimization tables, procedures, characterization, kinetic data, NMR spectra, SFC and HPLC traces, DFT computations, crystallographic data (PDF)

    Accession Codes

    CCDC 21077462107758 contain the supplementary crystallographic data for this paper. These data can be obtained free of charge via www.ccdc.cam.ac.uk/data_request/cif, or by emailing [email protected], or by contacting The Cambridge Crystallographic Data Centre, 12 Union Road, Cambridge CB2 1EZ, UK; fax: +44 1223 336033.


    Terms & Conditions

    Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.