You are on page 1of 11

pubs.acs.

org/acscatalysis Research Article

Halogen Bond Activation in Gold Catalysis


Helgi Freyr Jónsson, Daniel Sethio, Julian Wolf, Stefan M. Huber, Anne Fiksdahl,* and Mate Erdelyi*
Cite This: ACS Catal. 2022, 12, 7210−7220 Read Online

ACCESS Metrics & More Article Recommendations *


sı Supporting Information

ABSTRACT: Gold catalysis has, over the past decades, provided


innovative organic transformations under mild conditions with
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

high chemoselectivities. It receives steadily growing attention


thanks to its wide synthetic applicability. The catalytically active
form, [Ln-Au]+, of ligated gold complexes, [Ln-Au-Cl], is formed
via halide abstraction. This is typically achieved by anion exchange
upon the addition of an appropriate silver salt to the reaction
Downloaded via 185.183.146.239 on June 1, 2022 at 21:57:17 (UTC).

mixture. Herein, an alternative silver-free route for gold activation


is presented, making use of halogen bonding to promote halide
abstraction. We demonstrate that a catalytic amount of a strong
halogen bond donor efficiently activates both gold(I) and gold(III)
catalysts. Following the reaction, both the catalyst and the activator
are easily recovered. Importantly, this not only reduces the metal
waste in a gold-catalyzed process but also enables its upscaling, possibly opening new avenues for its use in industrial settings. Gold is
an expensive and limited resource, and its recyclability is of supreme importance. Based on systematic reaction kinetics, NMR
spectroscopic, and computational investigations, we describe the mechanism of halogen bond-activated gold(I/III) catalysis using
cyclopropanation as a model reaction. Our discovery paves the way for the development of gold-mediated transformations that allow
recycling of the precious gold catalyst and that may thereby become useful also for the large-scale generation of complex molecules.
KEYWORDS: gold, catalysis, halogen bond, bond activation, halide abstraction, recyclability

■ INTRODUCTION
Homogeneous gold catalysis has become a powerful tool for
Several approaches for silver-free gold catalysis have been
developed.24 These include chloride abstraction by use of
the synthesis of complex organic molecules.1−13 Offering ambiphilic ligands containing a hydrogen bond donor25 and an
internal Lewis acid26 as well as halide abstraction by alternative
outstanding chemoselectivities under mild conditions, it has
metal salts.27
found a wide range of applications in a variety of research
Halogen bond28-mediated catalysis is a novel approach in
fields.14−18 In most transformations, gold’s Lewis acidity is
synthesis.29−31 Halogen bond donors have recently been used
used for the activation of carbon−carbon multiple bonds to
to activate carbonyl groups to promote Diels−Alder reactions
promote the formation of a new carbon−carbon bond.
and Michael additions, for instance, and to assist the
Initially, gold chloride salts were used as catalysts, whereas
dissociation of halides to facilitate SN1-type processes.
later protocols apply ligands to modulate gold’s reactivity and
Activation of a metal−halogen bond by a halogen bond
stability.19 A common feature of the majority of gold-catalyzed
donor to enable catalytic reactions has recently been reported,
processes, ligated or non-ligated, is that the precatalyst
without a detailed mechanistic discussion however, by Huber
possesses a gold−chloride bond. Usually, this is activated
and co-workers, using the example of the gold(I)-catalyzed
through chloride abstraction by anion exchange using a silver cyclization of a propargylic amide and the cycloisomerization
salt with a weakly coordinating anion, generating the active of 1,6-enyne.32
catalyst, [Ln-Au]+, along with silver chloride. The potential The interaction of halogen bond donors with gold(III)
influences of silver20,21 and its counterion22 on gold catalysis complexes has not yet been assessed and that with gold(I) has
have been a matter of debate over the past decades. Silver is of only been scarcely explored.32 Halogen bond catalysis was
concern in ecosystems as it is one of the most toxic metals to
aquatic organisms.23 Recent efforts in synthesis target not only
the development of new processes but also the improvement of Received: April 16, 2022
the sustainability of existing transformations. Besides the Revised: May 14, 2022
decrease in catalyst loading and recycling, the substitution of
silver salts with alternatives capable of activating gold
precatalysts is therefore of vast interest, both from a
mechanistic perspective and from a sustainability perspective.
© XXXX The Authors. Published by
American Chemical Society https://doi.org/10.1021/acscatal.2c01864
7210 ACS Catal. 2022, 12, 7210−7220
ACS Catalysis pubs.acs.org/acscatalysis Research Article

proposed based on intuition but without the support of


comprehensive experimental and computational proof. We,
therefore, set the aim to evaluate whether the halogen bond
activation of gold(I) and gold(III) species may be a generally
applicable alternative to activation with the traditionally used
silver salts (Figure 1). In addition to disclosing a sustainable

Figure 1. Activation (a) with a silver salt is the common route in gold
catalysis (X− is a weakly coordinating anion). Herein, (b) gold
activation using a halogen bond donor (XBD) is explored as an
alternative sustainable route.

alternative synthetic route for gold(I/III) activation, we


provide a detailed mechanistic understanding of the halogen Figure 3. Halogen bond donors 7−11 were explored for their ability
bond-activated gold-catalyzed process.


to activate gold(I) and gold(III) catalysts.

RESULTS AND DISCUSSION


Design. To demonstrate the real-life applicability of a bond donor 7 or 8 (Scheme 1). For the dinuclear complex 2,
halogen bond activation strategy, we explored the most 10 mol % was used as this equals 20 mol % of gold(I) as
commonly used gold(I) complexes, namely, the gold(I) compared to the substrate. The conversion to product 14 was
JohnPhos complex 1, the chiral dinuclear BINAP gold(I) monitored by 1H NMR. Most importantly, the reaction
complex 2, the gold(I) N-heterocyclic carbene complexes progressed when a halogen bond donor was introduced to
possessing IMes (3) and IPr (5) ligands, and the respective the mixture. The fastest reaction was observed using the
gold(III) complexes 4 and 6 (Figure 2). Compounds 7−11 combination of IPrAuCl3 6 and halogen bond donor 7,
reaching completion in 18 h, whereas most other halogen
bond-activated reactions proceeded but did not give full
conversions within 60 h (Table 1, Figures 4 and 5). The
observed relative reaction rates suggest that the identities of
both the gold complex and the halogen bond donor play a role,
and simple conclusions, such as the stronger the halogen bond
donor, the quicker the reaction rate, cannot be drawn, but
factors other than being purely electronic might be at play as
well. Hence, halogen bond donor 8, which has a strong
electron-withdrawing CF3 group, is not always a stronger
activator than the somewhat less electron-poor 7 and gave
greater reaction rates only in combination with gold catalysts 2,
4, and 5 (Table 1). The high reactivity of dimesityl-substituted
iodoimidazolium salts may be due to the aryl groups’
preference to adopt conformations that minimize steric
repulsions, which breaks the conjugation with the imidazolium
ring.38 Accordingly, the positive charge is localized close to the
halogen bond donor iodine in these complexes.
Figure 2. Common gold complexes 1−6 were explored in the
We performed several control experiments to confirm that
investigation of halogen bond activation in gold catalysis. the activation truly originates from halogen bonding. Using 20
mol % gold complex 3 without any activator, we observed no
conversion after 24 h. This confirms that the gold catalyst does
(Figure 3) were chosen as halogen bond donors to be explored not show any background reactivity. No conversion was seen
for their halide abstraction ability. Due to their monodentate using 20 mol % of halogen bond donor 7 without a gold
nature, compounds 7 and 8 are weak halogen bond donors. catalyst. Hidden acid catalysis was ruled out by detection of no
The cationic iodolium complex 9 and the bidentate 10 and 11 conversion upon the addition of small amounts of HCl along
are strong halogen bond donors. We chose to use one of the with 20 mol % of gold complex 3. There was no conversion
most well-studied gold-mediated synthetic procedures, the either using 1,3-dimesitylimidazolium chloride, the nonhalogen
cyclopropanation of propargyl acetate 12 with the vinyl bond donor analogue of 7, in combination with catalyst 3,
derivative styrene 13 to cyclopropane 14 (Scheme 1) as a which corroborates our hypothesis that the halogen bond
model reaction,33 which has previously been used for the donor iodine of 7 is crucial for the activation of the gold
evaluation of the catalytic ability of novel gold(I) and gold(III) catalyst.
catalysts.34−37 Subsequently, we turned our attention to 9, which,
Reaction Rates. Reactions were initially performed by possessing a charged iodolium ion, is expected to be a stronger
mixing 1 equiv of propargyl acetate 12 and 2 equiv of styrene halogen bond donor than 7 and 8. Indeed, using as little as 0.5
13 with 20 mol % each of gold complexes 1−6 and halogen mol % 9 and 0.5 mol % gold catalyst 3, full conversion was
7211 https://doi.org/10.1021/acscatal.2c01864
ACS Catal. 2022, 12, 7210−7220
ACS Catalysis pubs.acs.org/acscatalysis Research Article

Scheme 1. Gold-Catalyzed (1−6) Cyclopropanation of Propargyl Acetate 12 Using the XBDs 7−11 as Activatorsa

a
In the reactions using 7 and 8 as activators, 20 mol % gold catalyst (1−6) and 20 mol % 7 or 8 were applied. In the reactions using 9−11 as the
activators, 0.5 mol % gold catalyst and 0.5 mol % chosen activator were used.

Table 1. Reactivity of Gold Complexes (Au) 1−6 upon


Activation with Halogen Bond Donors (Activator) 7 and 8
in Cyclopropanationa
activator Au conversionb krelc
7 1 31 3.2
7 2 38 3.9
7 3 60 5
7 4 25 2.6
7 5 10 1
7 6 84 8.8
8 1 16 1.7 Figure 4. Conversions for the cyclopropanation of propargyl acetate
8 2 70 7.3 12 to 14 (Scheme 1 and Table 1) as a function of reaction time using
8 3 43 4.5 7 (triangles) or 8 (circles) to activate gold(I) catalysts 1 (green), 2
8 4 29 3.0 (red), 3 (turquoise), and 5 (blue). Reactions were run at 25 °C,
8 5 63 6.6 followed by 1H NMR. The runs with the same gold catalyst are shown
8 6 45 4.7 with the same color. The control experiments, using 20 mol % gold
none 3 0d catalyst 3 without a halogen bond donor, using 20 mol % halogen
bond donor 7 without a gold catalyst, using 20 mol % each 1,3-
7 none 0d
dimesitylimidazolium chloride and gold catalyst 3, and finally, using
IMes-Cl 3 0d
20 mol % gold catalyst 3 with a small amount of HCl, are all shown as
HCl 3 0d black squares. None of the controls gave conversion.
a
Reactions were performed with 20 mol % of activators 7 and 8 and
gold complexes 1−6 in CD2Cl2 at 25 °C, [12] = 0.075 M, [Au] =
[XBD] = 0.015 M. Both cis and trans isomers of 14 are formed, and
the reported conversions are given as the sum of those of the isomers.
The isomerization is a slow background reaction that takes place in
the presence of gold complexes.34 bConversion. cRelative reaction
rates measured by 1H NMR at 12 h reaction time (before
isomerization). dControl experiment. Under “classical” activation
conditions, using 5 mol % AgSbF6 to activate the 5 mol % gold
complex, the reactions reach full conversion within 5 min, regardless
of the type of gold catalyst used.

obtained within 2.5 h (see Table 2 and Figure 6). Iodolium ion
9 was significantly better at activating gold(I) as compared to Figure 5. Conversions observed for the cyclopropanation of propargyl
gold(III) catalysts. The relative rate of cyclopropanation using acetate 12 to 14 (Scheme 1 and Table 1) as a function of reaction
activator 9 in combination with catalyst 6 is higher than that time using 7 (triangles) or 8 (circles) to activate gold(III) catalysts 4
using 4 as a catalyst (Table 2). Thus, the former only reached (bright green) and 6 (dark red). Reactions were run at 25 °C, and
60% conversion while the latter 89% within 24 h. This may conversion was followed by 1H NMR.
indicate the decomposition of 6 during the reaction.
Next, gold activation using the bidentate halogen bond Similar to the reactions using activator 9, the gold(III)-
donors 10 and 11 was explored. As the iodine of these can mediated reactions ran with halogen bond donors 10 and 11
simultaneously form two halogen bonds with the same were faster when using catalyst 4 (IMesAuCl3) as compared to
chlorine, they are highly efficient in facilitating the cleavage those using 6 (IPrAuCl3). It should be underlined that
of a Au−Cl bond and thereby activating the catalysts. activators 10 and 11 provided comparable reaction rates to
Accordingly, full conversion in the cyclopropanation shown that using NaBArF as the activator. Halogen bond donor 11
in Scheme 1 was reached within 10 min when using gold(I) gave faster initial reaction rates than NaBArF for both gold
catalysts. To obtain reliable reaction rates, we ran these complexes studied. The fact that the NaBArF-activated
reactions at −20 °C (Table 3). Similar to the reactions using reactions reached completion earlier might indicate an
activator 9, those using 10 and 11 were significantly slower induction period for these reactions.
when a gold(III) catalyst was applied and were, therefore, run Overall, the halogen bond donor strength positively
at 25 °C. The relative rates shown in Table 3 and Figure 7 for correlates with the reaction rate. Accordingly, the monodentate
gold(I)- and Figure 8 for gold(III)-catalyzed reactions are imidazolium-type 7 and 8 are weak gold(I/III) activators, even
therefore not directly comparable. at catalyst loadings as high as 20 mol %. The stronger
7212 https://doi.org/10.1021/acscatal.2c01864
ACS Catal. 2022, 12, 7210−7220
ACS Catalysis pubs.acs.org/acscatalysis Research Article

Table 2. Conversion of Alkyne 12 and Styrene 13 to Table 3. Conversion of Alkyne 12 and Styrene 13 to
Cyclopropane 14 (Scheme 1) Using Gold Complexes 1−6 Cyclopropane 14 (Scheme 1) Using Gold Catalysts 1−6
(Au) in Combination with Activator 9 and the (Au) in Combination with Halogen Bond Donor Activators
Corresponding Relative Reaction Rates (krel)a 10 and 11, and the Corresponding Relative Reaction Rates
(krel)a
Au conversionb krelc
1 80 10 activator Au conversionb krelc
2 48 6 With Gold(I) Catalysts, Reactions Were Run at −20 °C
3 88 11 10 1 20 2.2
4 8 1 10 2 46 4.9
5 51 6.4 10 3 41 4.4
6 18 2.3 10 5 9 1
a 11 1 37 4.0
Reactions were performed with 0.5 mol % halogen bond donor 9
and gold complexes 1−6 (0.0015 M each), 0.30 M 12, and 0.60 M 13 11 2 61 6.4
in CD2Cl2 at 25 °C, monitoring the conversions by 1H NMR. 11 3 80 8.5
b
Conversions. cRelative reaction rates (krel) following 60 min reaction 11 5 24 2.6
time are shown. As a consequence of the short reaction times, no cis- NaBArF 1 77 7.4
to-trans isomerization was observed. NaBArF 2 89 9.5
NaBArF 3 91 9.7
NaBArF 5 45 4.8
With Gold(III) Catalysts, Reactions Were Run at 25 °C
10 4 25 2.5
10 6 10 1
11 4 86 8.6
11 6 24 2.4
NaBArF 4 78 7.8
NaBArF 6 20 2
a
Reactions were performed with 0.5 mol % halogen bond donor 10 or
11 and gold complex 1−6 (0.0015 M each), 0.30 M 12 and 0.60 M
13 in CD 2 Cl 2 , monitoring the conversions by 1 H NMR.
b
Conversions. cRelative reaction rates (krel) following 30 min reaction
Figure 6. Conversions observed for the cyclopropanation of propargyl time are shown. Due to the short reaction times, no cis-to-trans
acetate 12 (Table 2) to 14 as a function of reaction time using isomerization was observed.
halogen bond donor 9 for activation of gold(I) catalysts 1 (green), 2
(red), 3 (turquoise), and 5 (blue) and gold(III) complexes 4 (bright
green) and 6 (dark red).

iodolium-type 9 and the bidentate 10 and 11 halogen bond


donors provide quicker conversions at very low catalyst
loadings of 0.5 mol %. Gold(III)-catalyzed reactions are
generally slower than the gold(I)-activated ones, even when
activated by a reagent other than a halogen bond donor
(Figure 9).
Catalyst and Activator Recycling. As reagent recycling is
an important aspect of sustainability, we evaluated the
recyclability of the gold catalyst and the activator. NMR
Figure 7. Conversions observed for the cyclopropanation of propargyl
monitoring indicated that the gold precatalysts remained acetate 12 (Table 2) to 14 as a function of reaction time using 10
unchanged throughout all the above reactions (Tables 1−3 (triangles), 11 (circles), or NaBArF (squares) for activation of gold(I)
and Figures 4−7), and hence, the intact precatalysts were catalysts 1 (green), 2 (red), 3 (turquoise), or 5 (blue). Reactions were
observed in all crude product mixtures. To evaluate run at −20 °C, and conversions were followed by 1H NMR.
recyclability, we have used the cyclopropanation reaction
mediated by gold precatalyst 1 and activator 9 as a model
system and performed the reaction on a ∼100 mg scale of catalyst decomposition is common, but even if it would not be
propargyl acetate 12 (93% yield). Following the reaction, one the case, the gold catalyst still gets decomposed in the
single flash column enabled the isolation of the target subsequent quenching process. As gold is an expensive and
cyclopropane 14 (93%), gold precatalyst 1 (88%), and rare element, the decomposition of its catalytic complexes
activator 9 (75%). The possibility to easily recover both the poses a serious limitation, particularly for large-scale
precatalyst and the activator is highly significant as the gold applications. Accordingly, an alternative route allowing smooth
precatalyst is not recoverable from reactions using anion recovery of the gold precatalyst, and of the activator, is a great
exchange when using a silver(I) salt for activation. Under such leap toward sustainability, especially for large-scale synthetic
classical conditions, the active gold catalyst is quantitatively applications.
prepared from a precatalyst, either in advance or in situ, with a The above recyclability implies a dynamic catalyst activation,
silver salt that is converted into a AgCl precipitate. The latter where the cationic active gold species is in equilibrium with the
cannot be directly reused. Throughout the reaction, gold chloride-ligated inactive complex throughout the reaction. This
7213 https://doi.org/10.1021/acscatal.2c01864
ACS Catal. 2022, 12, 7210−7220
ACS Catalysis pubs.acs.org/acscatalysis Research Article

Figure 10. The formation of a halogen bond between the Lewis basic
chlorine (halogen bond acceptor, blue) of the gold catalyst and the
Lewis acidic iodine (halogen bond donor, purple) was detected on the
NMR-active nuclei that are closest to the interacting atoms (in bold,
marked with a star) using 13C and 31P NMR spectroscopy. The
coordination shifts observed upon complex formation are given in
Figure 8. Conversions observed for the cyclopropanation of propargyl Tables 4 and 5.
acetate 12 (Table 2) to 14 as a function of reaction time using
halogen bond donors 10 (triangles) or 11 (circles) or NaBArF Table 4. 31P and 13C NMR Coordination Shifts (Δδ, ppm)
(square) for the activation of gold(III) catalysts 4 (green) and 6 for Gold Complexes 1−6 with Activators 7−11a
(red). Reactions were run at 25 °C, and the conversions were
followed by 1H NMR. catalyst 31
P NMR 13
C NMR
activator 1 2 3 4 5 6
7 0.045 0.45 −0.18 −0.36 −0.41 −0.46
8 0.060 0.51 −0.31 −0.39 −0.47 −0.51
9 0.96 7.15 −2.79 −2.56 −3.21 −3.65
10 0.74 5.27 −1.71 −2.10 −2.43 −2.75
11 0.96 6.26 −2.63 −2.85 −3.15 −3.63
a
NMR spectra were acquired in CD2Cl2 at 25 °C with a 0.03 M
concentration of the gold complex and the activator. For the accurate
determination of 31P NMR chemical shifts, PPh3 was used as the
internal standard.

bound to halogen bond donor iodine (Table 5). This


observation is in line with the formation of a halogen bond,

Table 5. 13C NMR Coordination Shifts (Δδ, ppm) of C−I


for XBDs 7−11 with Gold Complexes 1−6a
activator
catalyst 7 8 9 10 11 6
1 1.65 1.33 0.39 3.56 4.93 1.65
2 0.38 0.57 0.41 1.58 2.72 0.38
Figure 9. Normalized relative reaction rates observed for the
cyclopropanation of propargyl acetate 12 (Table 2) to 14 using 3 0.31 0.61 0.32 2.33 3.95 0.31
halogen bond donors 7−11 for activation of gold(I) and gold(III) 4 0.57 0.42 0.38 1.08 2.52 0.57
catalysts 1−6. For comparison, reactions ran at different conditions 5 0.98 1.07 0.24 2.79 3.94 0.98
were normalized to 0.5% catalyst concentration and 25 °C. For 6 0.33 0.48 0.39 1.35 2.94 0.33
details, see Table S1 in the Supporting Information. a
NMR spectra were acquired in CD2Cl2 at 25 °C with a 0.03 M
concentration of the gold complex and the activator.
is different from silver(I)-activated reactions that generate the
cationic gold species irreversibly. The dynamic gold activation corroborating our hypothesis. Following literature conven-
of the halogen bond-activated route is expected to be beneficial tions,42 the coordination shift is defined as the chemical shift
for the stability of the gold catalyst since catalyst decom- difference between the halogen bonded and the free ligand,
position often takes place during the off-cycle.39 Similarly, the that is, δ31Pcoord = δ31Pcomplex − δ31Pligand, when expressed for
31
slow release of cationic, catalytically active gold has been P NMR chemical shifts. Upon halogen bonding, the 31P
reported by using Cu(OTf)2 as the activating agent40 or by chemical shifts of 1 and 2 increased (Table 4). While the
applying thiolate-bridged, dinuclear NHC−Au(I) complexes.41 chemical shift change is small upon interacting with weak
NMR Chemical Shift Perturbation. To gain under- halogen bond donors 7 and 8, it dramatically increases when
standing of the mechanism of gold activation by halogen bond the strong halogen bond donors 9−11 are involved. The
donors, we measured the interaction-induced 13C and 31P magnitude of the chemical shift change shows a clear
NMR chemical shift changes. These were largest for the atoms dependence on the type of gold catalyst (Table 4). Similar
closest to the presumptive interaction sites of the gold to that observed for gold complexes 1 and 2, the coordination
complexes 1−6 and activators 7−11 (Figure 10). Thus, shifts induced by the strong halogen bond donors 9−11 on 3−
upon mixing the gold complexes with the activators (1:1), the 6 were a magnitude larger than those induced by the weak
largest chemical shift changes were observed for carbon or halogen bond donors 7 and 8. The 31P NMR chemical shifts
phosphorous bound to gold (Table 4) and carbon directly show an inverse trend as compared to that observed by 13C
7214 https://doi.org/10.1021/acscatal.2c01864
ACS Catal. 2022, 12, 7210−7220
ACS Catalysis pubs.acs.org/acscatalysis Research Article

NMR for 3−6, in line with the previous literature.43 Besides a coefficients is not just modulated by the molecular volume but
few exceptions, most literature studies report the inverse also by the interactionhere, the halogen bond strength.
correlation of 31P NMR chemical shifts to the electron density. Presuming a comparable volume for each catalyst and
This has been rationalized by electron-withdrawing substitu- activator, which is simplistic yet not unrealistic, a greater
ents (here, charge transfer via halogen bonding) decreasing the decrease in diffusion coefficient is indicative of a larger
electron density around phosphorus and thereby contracting population of the halogen-bonded complex and hence a
its d-orbitals; this is suggested to lead to an increased π back- stronger halogen bond. In line with this presumption, we
bonding with neighboring atoms that, in turn, results in detected a greater decrease in diffusion coefficient for all gold
increased shielding. We further observed that the signal of the complexes when interacting with the stronger halogen bond
carbenic carbon of gold catalyst 5 split into two upon its donors 9−11 as compared to the weak donors 7 and 8 (Figure
interaction with 9 (Figure S93, Supporting Information), 11). Activators 10 and 11 are, however, larger than 9, which in
whereas a single signal was observed for its complexes with 7− turn is also smaller than 7 and 8 (Scheme 1). Taking this into
8 and 10−11. The cause of this signal doubling is not entirely consideration, the decrease in diffusion coefficient suggests 9
clear; however, 7−8 and 10−11 are structurally closely related to be as strong halogen bond donor as 10−11, which agrees
2-iodoimidazolium salts, whereas 9 is an iodolium ion. The with the observed chemical shift perturbations shown in Table
latter is the only halogen bond donor among 7−11 capable of 4. Accordingly, the largest decrease in the diffusion coefficient
forming both mono- and dicoordinated complexes with of 7−11 upon complexation with 1−6 was observed for
halogen bond acceptors, which may explain the above iodolium complex 9 (Figure 12). The relative alteration in
observation. between the coefficients of 7−11 is, however, difficult to
Diffusion NMR. The coordination of gold catalysts with interpret.
halogen bond donor activators was further corroborated by the NMR Titration. NMR titration experiments were per-
observation of the consistent decrease in translational diffusion formed to provide further proof that halogen bonding is the
coefficient (D) of 1−6 and 7−11 upon their mixing (Figures dominant interaction between the gold catalysts and halogen
11 and 12; for tabulated values, see Supporting Information, bond donors. Halogen bond donor 7 was titrated with 1−10
equiv of gold catalyst 1 while keeping its concentration
constant and monitoring the chemical shift change of all nuclei
of the donor (Figure 13). Upon increasing the concentration

Figure 11. Alteration of the translational diffusion coefficients of gold


complexes 1−6 upon addition of an equivalent of activators 7 (blue),
8 (red), 9 (green), 10 (purple), and 11 (cyan). NMR spectra were
acquired in CD2Cl2 at 25 °C with [Au] = [activator] = 0.03 M.

Figure 13. 13C NMR chemical shift change of nuclei C-1 (orange), C-
2 (blue), and C-3 (pink) of halogen bond donor 7 when titrated with
gold catalyst 1 in 2.5 equiv increments. NMR spectra were acquired in
CD2Cl2 at 25 °C with concentrations [7] = 0.03 M and [1] = 0.03−
0.30 M.

of the gold catalyst, the fraction of complexed halogen bond


donor in solution increases, and the chemical shifts are
Figure 12. Alteration of the translational diffusion coefficients of expected to approach those of the halogen-bonded complex.
halogen bond donors 7−11 upon addition of an equivalent of gold The greatest chemical shift increase of the activator throughout
complexes 1 (blue), 2 (red), 3 (green), 4 (purple), 5 (cyan), and 6 the titration was observed on carbon neighboring iodine, C-1
(orange). NMR spectra were acquired in CD2Cl2 at 25 °C with [Au]
= [XBD] = 0.03 M.
(1.67 to 5.61 ppm), while other nuclei showed an order of
magnitude smaller shift alterations. This is in line with the
formation of the expected C−I···Cl−Au halogen bond. While
Tables S4 and S5). This indicates an increase in their Stokes binding curves were obtained in the NMR titrations, the data
radii, which is expected upon complexation. The diffusion were judged to not be complete enough to allow reliable
coefficient of the component of noncovalent complexes is the determination of binding constants but only to estimate that
population-weighted average of that of the free and of the the binding affinity of the halogen bond donor to the gold
bound forms. Hence, the magnitude of the observed diffusion catalyst is expected to be in the low millimolar range.
7215 https://doi.org/10.1021/acscatal.2c01864
ACS Catal. 2022, 12, 7210−7220
ACS Catalysis pubs.acs.org/acscatalysis Research Article

Scheme 2. Free Energy Profile Computed for the Cyclopropanation of Propargyl Acetate 12 with Styrene 13 Using a Halogen
Bond Donor to Activate the Gold(III) (Blue) or Gold(I) (Green) Catalysta

a
Simplified models of the gold(I/III) catalysts 3/4 were used due to the time limitations of DFT calculations. Following Int4, the reaction route
leading to the trans product is shown in red whereas that to the cis product in blue. Relative stabilities are given in kcal/mol, with respect to the
energy of the isolated fragments of the reactants (Tables S8 and S9). The full free energy diagram for the gold(I)-mediated process is given in
Figure S109 in the Supporting Information to avoid an overcrowded graph. Computations were done at the B3LYP-D3BJ/Def2-TZVP; for further
details, see the Supporting Information.

Computations. The above observations give a strong Supporting Information, Figure S112). This observation is in
experimental indication that gold catalysts are activated by good agreement with the literature studies.21,50−57 In agree-
halogen bond donors. The latter is expected to lower the ment with the experimentally observed trend of reaction rates,
energy barrier for the Au−Cl bond cleavage and thereby TS1 of the gold(I)-mediated process has a somewhat lower
promote the formation of the active gold catalyst. To gain an energy than that of the analogous gold(III)-mediated reaction
insight into the mechanism of halogen bond activation, we route (Scheme 2). The fully computed free energy profile for
carried out density functional theory (DFT) calculations for the gold(III)-catalyzed process is shown in Scheme 2, whereas
the envisioned reaction intermediates and related transition the corresponding free energy diagram for the gold(I)-
states of the gold-catalyzed cyclopropanation (Scheme 1). To catalyzed reaction is shown in Figure S109 of the Supporting
avoid computational limitations posed by the size of the Information. The gold complex TS1 is converted to Int2 upon
studied system, a truncated model (Scheme 1) was used, and completion of the chloride abstraction by the activator,
the computational method was carefully optimized (see Table adjoined with a simultaneous strengthening of the π-
S7 in the Supporting Information). Geometries were optimized coordination of propargylacetate to gold. This leads to
using B3LYP44-D3BJ45,46/Def2-SVP47 /CPCM (CH2Cl2), activation of the unsaturated bond of the substrate for
hence employing Grimme’s correction for dispersion effects. cyclization. The chloride complex of the halogen bond donor
Gibbs free energies were corrected using the quasi-harmonic leaving the gold(I/III) complex is slightly endothermic and
approximation48 and standard state correction.49 leads to the formation of Int3. At this point, the halogen bond
The mechanism of gold activation (Scheme 2) is expected to donor has completed its role as the activator. The subsequent
begin with the iodine of the activator forming a halogen bond 1,2-acyl shift takes place by rearrangement of Int3 via TS2,
with chlorine of the gold(I) or gold(III) catalyst. This leads to which involves the carboxylate group and leads to the
the formation of intermediate Int1, whereas the formation of formation of the five-membered ring of Int4. The subsequent
binary adducts of the two reactants is higher in energy (see ring opening to give the 2-acyl intermediate may proceed via
Tables S8 and S9, Supporting Information). According to two possible paths through a higher energy cis (TS3c) (blue)
DFT, a Au−Cl···I → Au···Cl−I reorganization takes place or a lower energy trans (TS3t) (red) configured transition
simultaneously to coordinate propargyl 12 with gold(I) or state, yielding Int5c and Int5t, respectively. Formation of a
gold(III), leading to the transition state TS1. Here, the styrene (13) π···π interaction stabilizes the Int6c complex,
association of alkyne is facilitated by the concerted weakening whereas no significant energy is gained for the trans-configured
of the Au−Cl covalent bond upon strengthening the Cl···I intermediate Int6t. Formation of the cyclopropane ring from
halogen bond. Since the free energies of TS1 are predicted to these intermediates passes intermediates of significantly
be 26.12 and 20.76 kcal/mol compared to those of the different activation energies, TS4t and TS4c, and leads to the
reactants for the gold(III)- and gold(I)-mediated processes, cis- and trans-configured products Pc and Pt, respectively.
respectively, the formation of TS1 is the rate-limiting step. Overall, both transformations leading to the cis- and trans-
This energy is in good agreement with the experimentally configured products are feasible, which is in agreement with
observed reaction rates for the gold(I)-mediated process, the experimental observations. Following dissociation of the
leading to a full conversion in ca. 60 h reaction time. The active cationic gold catalyst, [Ln-Au]+, from the product, the
alternative reaction route analogous to an SN1-type halide neutral precatalyst, [Ln-Au−Cl], is regenerated by chloride
abstraction followed by substrate coordination would have a transfer from the halogen bond donor−chloride complex. The
>50 kcal/mol energy barrier and is hence not realistic (see proposed mechanism is energetically feasible.
7216 https://doi.org/10.1021/acscatal.2c01864
ACS Catal. 2022, 12, 7210−7220
ACS Catalysis pubs.acs.org/acscatalysis Research Article

We further evaluated the influence of the halogen bond Hence, the stronger halogen bond the activator forms with
strength on the reaction rate. The electron densities at the the catalyst, the faster the reaction.
Au−Cl and I···Cl bond critical points ρAu−Cl and ρI···Cl, Principal component analysis (PCA) of the experimental
respectively, were estimated as these can be taken as indicators observables and computational parameters for gold(I) complex
for the interaction strengths. Computations (DFT; for details, 3 and gold(III) complex 4 indicates that the observed reaction
see the Supporting Information) confirm that the rate of the rate correlates with the experimentally observed chemical shift
gold(I)-mediated reaction increases as the electron density at of carbon C1 that is covalently bound to gold(I) in 3 (Figure
the I···Cl bond critical point increases, and the electron density 15) and gold(III) in 4 (Figure 16). The rates of the gold(I)-
at the Au−Cl bond critical point simultaneously decreases, that
is, as a halogen bond is formed. As the halogen bond strength
of the activator increases, the I···Cl interaction strengthens, and
the Au−Cl bond weakens, which is accordingly manifested in
the alteration of these bond lengths. The noteworthy
correlation (R2 0.962, Figure 14) of the relative reaction rate

Figure 15. PCA of the data of IMesAuCl (3). The rate is proportional
to the chemical shift of C1 (carbene) of the gold catalyst and is
inversely proportional to the nitrogen chemical shift. Here, ρ
represents the electron density at the bond critical point; C and N
Figure 14. Experimental reaction rates of gold(I) complex 3 as a are the chemical shifts of C and N atoms with the index indicating the
function of the 13C NMR coordination shifts (ppm) using activators position of the atoms, respectively; r is the bond length; Hc/ρc is the
7−11. Apart from one outlier, all data points suggest the correlation bond energy expressed as the energy density divided by the electron
of the chemical shifts of carbon directly bound to gold(I) and the density at the bond critical point; and log10 rate is the natural
reaction rates (see Table S20 in the Supporting Information). logarithm of the reaction rate.

with the 13C NMR coordination shift of carbon directly bound


to gold(I) corroborates the formation of a Cl···I halogen bond and gold(III)-catalyzed reactions are inversely proportional to
(Figure 14). Unexpectedly, the correlation of the reaction rate the chemical shift of nitrogen two bonds away from gold. The
to the electron density (Figures S114 and S115, Supporting inverse correlation is expected and is explained by the 15N
Information) and to the 13C NMR coordination shift of the NMR chemical shifts of aromatic heterocycles, in contrast to
halogen bond donor (Figure S116, Supporting Information) the 1H and 13C NMR chemical shifts, being dominantly
for gold(III) complexes is poor. Topological analysis of the influenced by the paramagnetic term instead of the electron
electron density suggests that the Au(III)−Cl bond is density-related diamagnetic contribution to shielding.58,59 Not
predominantly covalent, while the I···Cl interaction is unexpectedly, the reaction rate also correlates to the computed
dominantly electrostatic. Au−Cl bond length for both the gold(I)- and gold(III)-
The gold catalyst−activator complexes that form the mediated reactions. Furthermore, a somewhat weaker
strongest halogen bonds are expected to provide the highest correlation is seen for the halogen bond energy, expressed as
reaction rates because these promote the most efficient the energy density divided by the electron density at the bond
chloride abstraction. To confirm this hypothesis, we calculated critical point (Hc/ρc). These observations are in good
(at the M06-2X/Def2-SVP level of theory, Figure S104 in the agreement with the fact that halogen bonding plays an
activating role in the gold-mediated process.


Supporting Information) the molecular electrostatic potential
of halogen bond donors 7−11. In line with our expectations,
the halogen bond donors that provided the highest CONCLUSIONS
experimental reaction rates also showed the most positive Recyclable, homogeneous gold catalyst systems are so far
computed molecular electrostatic potentials, that is, 11 > 10 > scarce, and their design is typically based on the use of water-
9 ≫ 8 > 7 (Table S13, Supporting Information). soluble gold complexes and the reuse of the aqueous
The correlation of the experimental reaction rates and the solutions.60−63 Ionic liquids have also been used in a similar
sum of the computed molecular electrostatic potentials of the manner.64,65 These techniques do not allow for the isolation of
activators and the catalysts (Figures S105 and S106) confirms the catalyst and the activator as pure chemicals and catalysis in
that the halogen bond strength modulates the rate of the commonly applied organic solvents. Halogen bond
transformation of the studied cyclopropanation reaction. activation introduced herein as a resource-friendly alternative
7217 https://doi.org/10.1021/acscatal.2c01864
ACS Catal. 2022, 12, 7210−7220
ACS Catalysis pubs.acs.org/acscatalysis Research Article

bis(benzimidazolium)-type halogen bond donors, whereas


the weaker halogen bond monodentate imidazolium-type
donors are less strong activators of gold. DFT computations
estimate a ∼21 kcal/mol energy barrier for gold activation
through halogen bonding for a simplified model system, which
is in good agreement with the experimental reaction times.
In our experiments, activation by NaBArF along with
activator 11 provided the greatest catalytic activities. The
recycling of the activator in the halogen bond-activated process
also allows for the reuse of the expensive BArF− counterion.
Halogen bond activation is herein demonstrated as a resource-
friendly strategy for both gold(I)- and gold(III)-mediated
reactions. The first fragments of understanding of the
mechanism of halogen bond activation are given based on
NMR spectroscopic and computational data. Our results
provide the basis for the development of new, silver-free, and
ecological gold-catalyzed processes. The possibility of simple
and effective regeneration of both the gold catalyst and the
halogen bond donor activator may allow even future large-scale
Figure 16. PCA of the data of IMesAuCl3 (4). The rate is applications.


proportional to the C1 chemical shift of the gold complex and to
the Au−Cl bond length and is inversely proportional to the electron ASSOCIATED CONTENT
density at the Au−Cl bond critical point. Here, ρ represents the
electron density at the bond critical point; C and N are the chemical *
sı Supporting Information
shifts of C and N atoms with the index indicating the position of the The Supporting Information is available free of charge at
atoms, respectively; r is the bond length; Hc/ρc is the bond energy https://pubs.acs.org/doi/10.1021/acscatal.2c01864.
expressed as the energy density divided by the electron density at the
bond critical point; and log10 rate is the natural logarithm of the Details on the synthesis and spectroscopic data for
reaction rate. compound identification and details on the NMR,
computational, and kinetic investigations (PDF)


to the activation of gold(I) and gold(III) chloride precatalysts
by silver salts compensates for the above weaknesses. The
preparation of halogen bond activators 10 and 11 requires AUTHOR INFORMATION
significant synthetic effort. The synthesis of the strong activator Corresponding Authors
9 is, however, straightforward, and it may therefore be a Anne Fiksdahl − Department of Chemistry, Norwegian
promising candidate for future applications. It should be University of Science and Technology, Trondheim 7491,
emphasized that activators 7−11 were used here as model Norway; orcid.org/0000-0003-2577-2421;
systems to prove a principle and are not proposed to be the Email: anne.fiksdahl@ntnu.no
most optimal alternatives but rather starting points for future Mate Erdelyi − Department of ChemistryBMC, Uppsala
optimization. University, Uppsala SE-751 23, Sweden; orcid.org/0000-
The formation of a halogen bond between the activator and 0003-0359-5970; Email: mate.erdelyi@kemi.uu.se
the gold catalyst is indicated by chemical shift perturbation,
diffusion NMR, and chemical shift titration data along with Authors
PCA. The activation is shown to take place via an SN2-type Helgi Freyr Jónsson − Department of Chemistry, Norwegian
mechanism, and thus, there is a concerted transfer of chlorine University of Science and Technology, Trondheim 7491,
from gold and the coordination of the unsaturated bond of the Norway; orcid.org/0000-0002-1630-598X
substrate to gold. The original gold catalyst remains intact Daniel Sethio − Department of ChemistryBMC, Uppsala
during the reaction. No reaction takes place without an University, Uppsala SE-751 23, Sweden; orcid.org/0000-
activator. Both the gold catalyst and the halogen bond donor 0002-8075-1482
activator can be efficiently and easily regenerated on a Julian Wolf − Faculty of Chemistry and Biochemistry, Organic
preparative scale. This is an important advantage, in addition Chemistry I, Ruhr-Universität Bochum, Bochum 44801,
to avoiding the use of silver for activation, paving the way for Germany
upscaling and for the development of gold-catalyzed ecological Stefan M. Huber − Faculty of Chemistry and Biochemistry,
industrial processes. The variety of gold catalysts and halogen Organic Chemistry I, Ruhr-Universität Bochum, Bochum
bond donors tested in this work suggests that this concept may 44801, Germany; orcid.org/0000-0002-4125-159X
be applicable to a number of additional gold-catalyzed Complete contact information is available at:
processes. https://pubs.acs.org/10.1021/acscatal.2c01864
The rate of the gold-mediated cyclopropanation is
determined by the halogen bond donor strength of the
Funding
activator, as demonstrated by the correlation of the relative
reaction rates and the coordination shifts of carbon directly Research Council of Norway (226244/F50), Swedish Research
bound to gold and that to iodine in the catalyst and the Council (2020-03431).
activator, respectively. An iodolium ion was observed to be a Notes
strong activator as the bidentate bis(imidazolium)- and The authors declare no competing financial interest.
7218 https://doi.org/10.1021/acscatal.2c01864
ACS Catal. 2022, 12, 7210−7220
ACS Catalysis


pubs.acs.org/acscatalysis Research Article

ACKNOWLEDGMENTS (16) Arcadi, A.; Giuseppe, S. Recent applications of gold catalysis in


organic synthesis. Curr. Org. Chem. 2004, 8, 795−812.
We gratefully acknowledge the Norwegian University of (17) Alshammari, A. S. Heterogeneous Gold Catalysis: From
Science and Technology for a Ph.D. studentship for Helgi Discovery to Applications. Catalysts 2019, 9, 402.
Freyr Jónsson. This work was partly supported by the Research (18) Pflästerer, D.; Hashmi, A. S. K. Gold catalysis in total
Council of Norway through the Norwegian NMR Platform, synthesisrecent achievements. Chem. Soc. Rev. 2016, 45, 1331−
NNP (226244/F50). A/S.M.H. Lundgreens Enkes Fond is 1367.
thanked for their generous support in covering NMR (19) Collado, A.; Nelson, D. J.; Nolan, S. P. Optimizing Catalyst and
instrument costs. We thank Stefano Battaglia (Uppsala Reaction Conditions in Gold(I) Catalysis-Ligand Development.
University) and Oriana Brea Noriega (KTH Royal Institute Chem. Rev. 2021, 121, 8559−8612.
of Technology) for helpful discussions. This project made use (20) Wang, D.; Cai, R.; Sharma, S.; Jirak, J.; Thummanapelli, S. K.;
of the NMR Uppsala infrastructure, which is funded by the Akhmedov, N. G.; Zhang, H.; Liu, X.; Petersen, J. L.; Shi, X. “Silver
Department of Chemistry, BMC and the Disciplinary Domain Effect” in Gold(I) Catalysis: An Overlooked Important Factor. J. Am.
of Medicine and Pharmacy. We also thank Gaston Courtade Chem. Soc. 2012, 134, 9012−9019.
and Nebojsa Simic (Norwegian University of Science and (21) Ranieri, B.; Escofet, I.; Echavarren, A. M. Anatomy of gold
Technology) for their assistance with diffusion NMR. The catalysts: facts and myths. Org. Biomol. Chem. 2015, 13, 7103−7118.
(22) Jia, M.; Bandini, M. Counterion Effects in Homogeneous Gold
computations were enabled by resources provided by the
Catalysis. ACS Catal. 2015, 5, 1638−1652.
Swedish National Infrastructure for Computing (SNIC) at (23) Eisler, R. Silver Hazards to Fish, Wildlife, and Invertebrates: A
Tetralith and partially funded by the Swedish Research Synoptic Review; Contaminant Hazard Biological Report; US Depart-
Council through grant agreement no. 2018-05973, under ment of the Interior, National Biological Service: Laurel, MD, USA,
project numbers 2020/5-435, 2021/5-359, and 2021/22-350. 1996.
J.W. and S.M.H. acknowledge financial support by the (24) Franchino, A.; Montesinos-Magraner, M.; Echavarren, A. M.
Deutsche Forschungsgemeinschaft (DFG; grant number HU Silver-Free Catalysis with Gold(I) Chloride Complexes. Bull. Chem.
1782/6-1). Soc. Jpn. 2021, 94, 1099−1117.


(25) Sen, S.; Gabbaï, F. P. An ambiphilic phosphine/H-bond donor
REFERENCES ligand and its application to the gold mediated cyclization of
propargylamides. Chem. Commun. 2017, 53, 13356−13358.
(1) Hashmi, A. S. K.; Hutchings, G. J. Gold catalysis. Angew. Chem., (26) Devillard, M.; Nicolas, E.; Appelt, C.; Backs, J.; Mallet-Ladeira,
Int. Ed. Engl. 2006, 45, 7896−7936.
S.; Bouhadir, G.; Slootweg, J. C.; Uhl, W.; Bourissou, D. Novel
(2) Gorin, D. J.; Toste, F. D. Relativistic effects in homogeneous
zwitterionic complexes arising from the coordination of an ambiphilic
gold catalysis. Nature 2007, 446, 395−403.
(3) Fürstner, A. Gold and platinum catalysisa convenient tool for phosphorus-aluminum ligand to gold. Chem. Commun. 2014, 50,
generating molecular complexity. Chem. Soc. Rev. 2009, 38, 3208− 14805−14808.
3221. (27) Fang, W.; Presset, M.; Guérinot, A.; Bour, C.; Bezzenine-
(4) Krause, N.; Winter, C. Gold-Catalyzed Nucleophilic Cyclization Lafollée, S.; Gandon, V. Silver- Free Two- Component Approach in
of Functionalized Allenes: A Powerful Access to Carbo- and Gold Catalysis: Activation of [ LAuCl] Complexes with Derivatives of
Heterocycles. Chem. Rev. 2011, 111, 1994−2009. Copper, Zinc, Indium, Bismuth, and Other Lewis Acids. Chem.Eur.
(5) Bandini, M. Gold-catalyzed decorations of arenes and J. 2014, 20, 5439−5446.
heteroarenes with C-C multiple bonds. Chem. Soc. Rev. 2011, 40, (28) Desiraju, G. R.; Ho, P. S.; Kloo, L.; Legon, A. C.; Marquardt,
1358−1367. R.; Metrangolo, P.; Politzer, P.; Resnati, G.; Rissanen, K. Definition of
(6) Dorel, R.; Echavarren, A. M. Gold(I)-Catalyzed Activation of the halogen bond (IUPAC Recommendations 2013). Pure Appl.
Alkynes for the Construction of Molecular Complexity. Chem. Rev. Chem. 2013, 85, 1711−1713.
2015, 115, 9028−9072. (29) Bulfield, D.; Huber, S. M. Halogen Bonding in Organic
(7) Hopkinson, M. N.; Tlahuext-Aca, A.; Glorius, F. Merging Visible Synthesis and Organocatalysis. Chem.Eur. J. 2016, 22, 14434−
Light Photoredox and Gold Catalysis. Acc. Chem. Res. 2016, 49, 14450.
2261−2272. (30) Bamberger, J.; Ostler, F.; Mancheño, O. G. Frontiers in
(8) Witzel, S.; Hashmi, A. S. K.; Xie, J. Light in Gold Catalysis. Halogen and Chalcogen-Bond Donor Organocatalysis. Chemcatchem
Chem. Rev. 2021, 121, 8868−8925. 2019, 11, 5198−5211.
(9) Chintawar, C. C.; Yadav, A. K.; Kumar, A.; Sancheti, S. P.; Patil, (31) Sutar, R. L.; Huber, S. M. Catalysis of Organic Reactions
N. T. Divergent Gold Catalysis: Unlocking Molecular Diversity through Halogen Bonding. ACS Catal. 2019, 9, 9622−9639.
through Catalyst Control. Chem. Rev. 2021, 121, 8478−8558. (32) Wolf, J.; Huber, F.; Erochok, N.; Heinen, F.; Guérin, V.;
(10) Li, D.; Zang, W.; Bird, M. J.; Hyland, C. J. T.; Shi, M. Gold- Legault, C. Y.; Kirsch, S. F.; Huber, S. M. Activation of a Metal-
Catalyzed Conversion of Highly Strained Compounds. Chem. Rev. Halogen Bond by Halogen Bonding. Angew. Chem., Int. Ed. Engl.
2021, 121, 8685−8755.
2020, 59, 16496−16500.
(11) Rocchigiani, L.; Bochmann, M. Recent Advances in Gold(III)
(33) Johansson, M. J.; Gorin, D. J.; Staben, S. T.; Toste, F. D.
Chemistry: Structure, Bonding, Reactivity, and Role in Homogeneous
Catalysis. Chem. Rev. 2021, 121, 8364−8451. Gold(I)-catalyzed stereoselective olefin cyclopropanation. J. Am.
(12) Wang, T.; Hashmi, A. S. K. 1,2-Migrations onto Gold Carbene Chem. Soc. 2005, 127, 18002−18003.
Centers. Chem. Rev. 2021, 121, 8948−8978. (34) Reiersølmoen, A. C.; Østrem, E.; Fiksdahl, A. Gold(III)-
(13) Herrera, R. P.; Gimeno, M. C. Main Avenues in Gold Catalysed cis-to-trans Cyclopropyl Isomerization. Eur. J. Org. Chem.
Coordination Chemistry. Chem. Rev. 2021, 121, 8311−8363. 2018, 2018, 3317−3325.
(14) Webster, S.; O’Rourke, K. M.; Fletcher, C.; Pimlott, S. L.; (35) Solvi, T. N.; Reiersølmoen, A. C.; Orthaber, A.; Fiksdahl, A.
Sutherland, A.; Lee, A.-L. Rapid Iododeboronation with and without Studies towards Pyridine-Based N,N,O-Gold(III) Complexes: Syn-
Gold Catalysis: Application to Radiolabelling of Arenes. Chem.Eur. thesis, Characterization and Application. Eur. J. Org. Chem. 2020,
J. 2018, 24, 937−943. 2020, 7062−7068.
(15) Ciriminna, R.; Falletta, E.; Pina, C. D.; Teles, J. H.; Pagliaro, M. (36) Reiersølmoen, A. C.; Solvi, T. N.; Fiksdahl, A. Au(III)
Industrial Applications of Gold Catalysis. Angew. Chem., Int. Ed. Engl. complexes with tetradentate-cyclam-based ligands. Beilstein J. Org.
2016, 55, 14210−14217. Chem. 2021, 17, 186−192.

7219 https://doi.org/10.1021/acscatal.2c01864
ACS Catal. 2022, 12, 7210−7220
ACS Catalysis pubs.acs.org/acscatalysis Research Article

(37) Reiersølmoen, A. C.; Fiksdahl, A. Pyridine- and Quinoline- Cycloisomerization of 7-Aryl-1,6-Enynes. Organometallics 2014, 33,
Based Gold(III) Complexes: Synthesis, Characterization, and 6466−6473.
Application. Eur. J. Org. Chem. 2020, 2020, 2867−2877. (57) Couce-Rios, A.; Kovács, G.; Ujaque, G.; Lledós, A. Hydro-
(38) Haraguchi, R.; Hoshino, S.; Sakai, M.; Tanazawa, S.-g.; Morita, amination of C-C Multiple Bonds with Hydrazine Catalyzed by N-
Y.; Komatsu, T.; Fukuzawa, S.-i. Bulky iodotriazolium tetrafluor- Heterocyclic Carbene-Gold(I) Complexes: Substrate and Ligand
oborates as highly active halogen-bonding-donor catalysts. Chem. Effects. ACS Catal. 2015, 5, 815−829.
Commun. 2018, 54, 10320−10323. (58) Pazderski, L. 15N NMR coordination shifts in Pd(II), Pt(II),
(39) Kumar, M.; Jasinski, J.; Hammond, G. B.; Xu, B. Alkyne/ Au(III), Co(III), Rh(III), Ir(III), Pd(IV), and Pt(IV) complexes with
alkene/allene-induced disproportionation of cationic gold(I) catalyst. pyridine, 2,2′-bipyridine, 1,10-phenanthroline, quinoline, isoquino-
Chem.Eur. J. 2014, 20, 3113−3119. line, 2,2′-biquinoline, 2,2′:6′, 2′-terpyridine and their alkyl or aryl.
(40) Guérinot, A.; Fang, W.; Sircoglou, M.; Bour, C.; Bezzenine- Magn. Reson. Chem. 2008, 46, S3−S15.
Lafollée, S.; Gandon, V. Copper Salts as Additives in Gold(I)- (59) Solum, M. S.; Altmann, K. L.; Strohmeier, M.; Berges, D. A.;
Catalyzed Reactions. Angew. Chem., Int. Ed. Engl. 2013, 52, 5848− Zhang, Y.; Facelli, J. C.; Pugmire, R. J.; Grant, D. M. 15N Chemical
5852. Shift Principal Values in Nitrogen Heterocycles. J. Am. Chem. Soc.
(41) Mora, M.; Gimeno, M. C.; Visbal, R. Recent advances in gold- 1997, 119, 9804−9809.
NHC complexes with biological properties. Chem. Soc. Rev. 2019, 48, (60) Tomás-Mendivil, E.; Toullec, P. Y.; Borge, J.; Conejero, S.;
447−462. Michelet, V.; Cadierno, V. Water-Soluble Gold(I) and Gold(III)
(42) Carlsson, A.-C. C.; Mehmeti, K.; Uhrbom, M.; Karim, A.; Complexes with Sulfonated N-Heterocyclic Carbene Ligands: Syn-
Bedin, M.; Puttreddy, R.; Kleinmaier, R.; Neverov, A. A.; thesis, Characterization, and Application in the Catalytic Cyclo-
Nekoueishahraki, B.; Gräfenstein, J.; Rissanen, K.; Erdélyi, M. isomerization of gamma-Alkynoic Acids into Enol-Lactones. ACS
Substituent Effects on the [N-I-N]+ Halogen Bond. J. Am. Chem. Catal. 2013, 3, 3086−3098.
Soc. 2016, 138, 9853−9863. (61) Sak, H.; Mawick, M.; Krause, N. Sustainable Gold Catalysis in
(43) Dunn, E. J.; Purdon, J. G.; Bannard, R. A. B.; Albright, K.; Water Using Cyclodextrin-tagged NHC-Gold Complexes. Chem-
Buncel, E. Correlation of P-31 Nuclear Magnetic-Resonance CatChem 2019, 11, 5821−5829.
Chemical-Shifts in Aryl Phosphinates with Hammett Substituent (62) Fernandez, G. A.; Dorn, V. B.; Chopa, A. B.; Silbestri, G. F.
ConstantsInductive Versus Resonance Interactions and Relevance Sulphonated NHC-gold(I) complexes in aqueous catalysis. An
to P-Pi-D-Pi Bonding. Can. J. Chem. 1988, 66, 3137−3142. experimental and DFT computational analysis. 19th Int. Electronic
(44) Becke, A. D. Density-functional exchange-energy approxima- Conf. on Synth. Org. Chem., 2015.
tion with correct asymptotic behavior. Phys. Rev. A: At., Mol., Opt. (63) Minkler, S. R. K.; Lipshutz, B. H.; Krause, N. Gold Catalysis in
Phys. 1988, 38, 3098−3100. Micellar Systems. Angew. Chem., Int. Ed. Engl. 2011, 50, 7820−7823.
(45) Grimme, S.; Ehrlich, S.; Goerigk, L. Effect of the damping (64) Baron, M.; Biffis, A. Gold(I) Complexes in Ionic Liquids: An
function in dispersion corrected density functional theory. J. Comput. Efficient Catalytic System for the C-H Functionalization of Arenes
Chem. 2011, 32, 1456−1465. and Heteroarenes under Mild Conditions. Eur. J. Org. Chem. 2019,
(46) Lee, C.; Yang, W.; Parr, R. G. Development of the Colle- 2019, 3687−3693.
Salvetti correlation-energy formula into a functional of the electron (65) Marinelli, F.; Ambrogio, I.; Arcadi, A.; Cacchi, S.; Fabrizi, G.
density. Phys. Rev. B: Condens. Matter Mater. Phys. 1988, 37, 785−789. Gold-catalyzed synthesis of 2-substituted, 2,3-Disubstituted and 1,2,3-
(47) Weigend, F.; Ahlrichs, R. Balanced basis sets of split valence, trisubstituted Indoles in [Bmim]BF4. Synlett 2007, 2007, 1775−1779.
triple zeta valence and quadruple zeta valence quality for H to Rn:
Design and assessment of accuracy. Phys. Chem. Chem. Phys. 2005, 7,
3297−3305.
(48) Grimme, S. Supramolecular binding thermodynamics by
dispersion-corrected density functional theory. Chem.Eur. J. 2012,
18, 9955−9964.
(49) Jensen, J. H. Predicting accurate absolute binding energies in
aqueous solution: thermodynamic considerations for electronic
structure methods. Phys. Chem. Chem. Phys. 2015, 17, 12441−12451.
(50) Komiya, S.; Albright, T. A.; Hoffmann, R.; Kochi, J. K.
Reductive elimination and isomerization of organogold complexes.
Theoretical studies of trialkylgold species as reactive intermediates. J.
Am. Chem. Soc. 1976, 98, 7255−7265.
(51) Colburn, C. B.; Hill, W. E.; Mcauliffe, C. A.; Parish, R. V. P-31
Nmr-Study of Tertiary Phosphine Complexes of Gold(I). Chem.
Commun. 1979, 218−219.
(52) Dickson, P. N.; Wehrli, A.; Geier, G. Coordination Chemistry
of Gold(I) with Cyanide and 1-Methylpyridine-2-ThioneKinetics
and Thermodynamics of Ligand-Exchange at Gold(I) in Aqueous-
Solution. Inorg. Chem. 1988, 27, 2921−2925.
(53) Komiya, S.; Iwata, M.; Sone, T.; Fukuoka, A. Isolation of
Highly Nucleophilic Gold(I) Alkoxides Having a Tertiary Phosphine
Ligand. Chem. Commun. 1992, 16, 1109−1110.
(54) Zhdanko, A.; Maier, M. E. Quantitative Evaluation of the
Stability of gem-Diaurated Species in Reactions with Nucleophiles.
Organometallics 2013, 32, 2000−2006.
(55) Xi, Y.; Wang, Q.; Su, Y.; Li, M.; Shi, X. Quantitative kinetic
investigation of triazole-gold(I) complex catalyzed [3,3]-rearrange-
ment of propargyl ester. Chem. Commun. 2014, 50, 2158−2160.
(56) Brooner, R. E. M.; Robertson, B. D.; Widenhoefer, R. A.
Mechanism of 1,3-Hydrogen Migration in a Gold Bicyclo[3.2.0]-
heptene Complex: The Role of Bronsted Acid in the Gold-Catalyzed

7220 https://doi.org/10.1021/acscatal.2c01864
ACS Catal. 2022, 12, 7210−7220

You might also like