Carbonyl-Olefin Metathesis Catalyzed by ... - ACS Publications

tion mode via molecular iodine and iodonium ion that could change the previously established perception of catalyst and substrate design for the carbo...
0 downloads 0 Views 1MB Size
Subscriber access provided by YORK UNIV

Article

Carbonyl-Olefin Metathesis Catalyzed by Molecular Iodine Uyen Phuoc Nhat Tran, Giulia Oss, Martin Breugst, Eric Detmar, Domenic Paul Pace, Kevin Liyanto, and Thanh Vinh Nguyen ACS Catal., Just Accepted Manuscript • DOI: 10.1021/acscatal.8b03769 • Publication Date (Web): 14 Dec 2018 Downloaded from http://pubs.acs.org on December 15, 2018

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 9 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

Carbonyl-Olefin Metathesis Catalyzed by Molecular Iodine Uyen P. N. Tran,a Giulia Oss,a Martin Breugst,b Eric Detmar,b Domenic P. Pace,a Kevin Liyantoa and Thanh V. Nguyena,* a b

School of Chemistry, University of New South Wales, Sydney, Australia Department für Chemie, Universität zu Köln, Köln, Germany

Supporting Information Placeholder ABSTRACT: The carbonyl-olefin metathesis reaction

could facilitate rapid functional group interconversion and allow construction of complicated organic structures. Herein we demonstrate that elemental iodine, a very simple catalyst, can efficiently promote this chemical transformation under mild reaction conditions. Our mechanistic studies revealed intriguing aspects of activation mode via molecular iodine and iodonium ion that could change the previously established perception of catalyst and substrate design for the carbonyl-olefin metathesis reaction.

Catalytic Carbonyl-Olefin Metathesis (COM) Reaction R2 R1

27 examples 21-96% practical on gram scale broad substrate scope (all three COM reaction types)

Previously: Lambert’s hydrazine catalyst NH NH

The carbonyl-olefin metathesis reaction (COM),1 an analogue of the olefin-olefin metathesis reaction,2 is considered to be very useful for direct carbon-carbon bond forming transformations and functional group interconversions. There had been a number of developments on experimental protocols3 and catalysts4 for the COM reaction. Over the last five years, it has become increasingly attractive after interesting studies from the Lambert group using hydrazine5 and the Franzen group using tritylium salts6 as organocatalysts for this reaction (Scheme 1). More recently, the Schindler group7 and the Li group8 reported elegant studies in which they utilized salts of iron(III) and gallium(III) to promote COM reactions. Earlier this year, the Tiefenbacher group also published an interesting report on the cooperative effect of hexameric resorcinarene assembly and Brønsted acid to catalyze COM reactions.9 A recent study from our laboratory showed that tropylium ion10 could also be used as an organocatalytic promoter for the COM reaction (Scheme 1).11 The COM reaction is, however, still much less investigated than the olefin-olefin metathesis,12 most likely due to a lack of practicability and general knowledge about rational catalyst and substrate design for the reaction. Herein we demonstrate that elemental iodine, a very simple catalyst, can efficiently promote the COM reaction on a broad range of substrates under

.

(ref 5)

(ref 7a-d and 8)

intramolecular COM only GaCl3

Franzen’s tritylium catalyst Ar

2HCl

Schindler’s and Li’s catalyst FeCl3

R2

O

rt, 24 h via iodonium catalysis

ring-opening COM of cyclopropene only

INTRODUCTION

R1

This work: [cat.] = I2

O

Ar

Ar

X

(ref 6)

intermolecular COM only

Tiefenbacher’s catalyst HCl inside hexameric (ref 9) resorcinarene host intramolecular COM only

(ref 7e)

ring-opening COM only

Our tropylium catalyst

BF4 (ref 11)

all three types: intra/inter-molecular and ring-opening

Scheme 1. Catalytic carbonyl-olefin metathesis reaction mild conditions with excellent outcomes. The inspiration for our use of iodine as a promoter for the COM reaction came from previous reports of iodine being able to activate carbonyl substrates for a range of chemical reactions via halogen bonding interaction.13 Elemental iodine can also serve as a pre-catalyst for iodonium-activation of alkenes14 or alkynes15 in isomerization and cyclization reactions. We predicted that iodine could either activate the carbonyl moiety or the alkene moiety, or cooperatively activate both functional groups for the carbonyl-olefin metathesis reaction. Thus, we set out to test our hypothesis by using I2 catalyst to promote the intramolecular COM reaction of substrate 1a1 (Scheme 2).

ACS Paragon Plus Environment

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

R1

O

R3

R4

2a’

Ph

O

CO2Me

CO2Et

200 mg, 93% (2a : 2a’ = 11 : 1) 1.60 g, 89% (2a : 2a’ = 38 : 1)

1a1

Products

Me

CO2Et

2a

2l, 23%

1l

O

Me CO2Et

CO2Et

2a

1a2

2m 1m

Ph

2m’

94% (2m : 2m’ = 2.5 : 1)

27% (2a : 2a’ = 11 : 1)

O

AND

H 2C

Me

H CO2Et

CO2Et

2a

1a3

Me

O

CO2Et

2a’

Me

CO2Et

Me

Me

AND CO2iPr

CO2

2c

iPr

Me

MeO

AND

MeO

Me

H

2p, 47% Me

O

MeO CO2Et

2d

1d

Me Me

1p

CO2Et

CO2Et

N

O

AND O

Ph 1e

N

O

2e

Ph

2g’

Me Allyl

Allyl

CO2Et

O

Br

Me Allyl

Me

Me N

Me Ts

Me CO2Et 1k

CO Et Me 2 2k, 54%

2t (*), 78%

N

N Ts

Me Ts Me

1v

2u (*), 86%

Me N

Me Ts

Ph

Scheme 2. Iodine-catalyzed intramolecular carbonyl-olefin metathesis reactions16 ACS Paragon Plus Environment

Br

1u

O

Me

O

N Ts

Me

2j, 58%

1j

Me

Me 2s, 54%

Me Allyl

N Ts

1t

O

Me

Me

Ts

2i, 46%

1i

O

2s 104 mg, 92% 1.51 g, 91%

H

1s2

2h, 59%

CO2Et

N Me

Me CO2Bn

Me

Ph

O

Me

O

Me

1s1

Ph

Me

N Ts

Me Ts

Me

Me

CO2Bn 1h

N

Me

42% (2g : 2g’ = 1 : 3)

O

Me

O

AND

2g

1g

2r, 96%

2e’

Ph

Ph

Me

1r

60% (2e : 2e’ = 1 : 3)

Me

N Ts

Me Ts

Me

Ph

Me

O

2q (*), 82% Me

Me Me

N Ts

Me Ts

1q

2d’ 83% (2d : 2d’ = 4 : 1)

O

2o’, 32%

2c’

58% (2c : 2c’ = 2.1 : 1)

O

Me

1o

O

Me

1c

Me

2b’ 76% (2b : 2b’ = 2 : 1)

CO2iPr

Me

O

CO2Et

2b

O

2n’ 54% (2n : 2n’ = 1 : 1)

AND

Me

1b

Me

2n

1n

Me

CO2Et

O

AND

Me

48% (2a : 2a’ = 11 : 1)

O

Me Me

AND

Me

Me

2a’

H 2C

Me

Me

AND

CO2Et

3

Me

O

CO2Et

CO2Et

O

Substrate

AND

Me Me

R2

2’ (isomerized product)

2 (COM product)

Products

R1

( )n R4

R4

(*) = using 25 mol% cat. I2

Substrate

R3

( )n

neat, rt, 24 h

1 (see pages S5-S17 in the SI)

O

R3

10 mol% I2

( )n R2

Page 2 of 9

N Ts Ph 2v, 76%

Page 3 of 9 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

RESULTS AND DISCUSSIONS Gratifyingly, our preliminary investigation immediately showed very promising outcomes: within 24 hours at 50 ºC using 5 mol% of I2 as catalyst, substrate 1a1 was converted to the cyclized COM product 2a and its isomerized product 2a’ in 62% total yield (~ 90% conversion).16 Subsequent optimization studies (see page S3 in the SI) revealed that the reaction could be carried out smoothly and cleanly at ambient temperature and atmosphere in solvent-free conditions using 10 mol% I2 catalyst, giving the products 2a and 2a’ in 93% total yield after purification.16 The optimal conditions were effective for a wide range of substrates (Scheme 2). Most of products 2 were obtained in moderate to excellent yields from generally clean intramolecular COM reactions when substrates 1 bear the isopropylidene moiety (i.e. when acetone is formed as a by-product, Scheme 2, see page S5-S17 in the SI for the specific structures of COM precursors 1).16 Substrates with active olefin moieties bearing different substituents (such as Ph or H) also cyclized via the COM reactions, albeit with lower yields (Scheme 2, comparison of entries 1a1 to 1a2-1a3 and entries 1s1 to 1s2). This phenomenon has been previously observed with transition metal Lewis acid catalysts7a-d,8 and our own tropylium catalytic system. This new protocol can be practically carried out on gram scale with very high efficiency, as demonstrated for substrates 1a1 and 1s1 to form 1.60 g of 2a and 1.51 g of 2s respectively (Scheme 2).16 The most notable feature of this protocol is that it only required short reaction time at ambient temperature to achieve similar results to the COM reactions promoted by other catalytic systems7a-e,8-9,11 at harsher conditions. As expected with the benign nature of the iodine catalyst and the mild reaction conditions, this COM protocol tolerates several types of functional groups. Most of the non-nitrogen containing substrates (1a-1p) went through the iodine-catalyzed COM reactions to give a mixture of two regioisomeric products. When the functional group at the α-position of the newly formed C-C double bond is a alkoxycarbonyl group, the expected COM products (2a-2d and 2h) were produced as major components. When this functional group is an acyl group, phenyl ring or methyl group, the isomerized olefins were formed as the major products (2e’, 2g’ and 2o’). The formation of these isomerized products can be attributed to isomerization of the C-C double bond, generated after the COM reactions, presumably due to traces amounts of Brønsted acids present in the reaction mixture (Table 1, vide infra). The Tiefenbacher group also reported similar observations in their Brønsted-acid catalyzed COM reactions.9 These regioisomeric ratios changed with modifications of reaction conditions (Scheme 2) and we observed the isomerization (2a to 2a’, see page S31 in the SI for more details) when subjecting the pure product 2a to the same I2-catalyzed COM reaction conditions.

Substrates bearing Brønsted-basic nitrogen-centers8 did not undergo this type of isomerization (Scheme 2, 2q-2v). They worked smoothly to give cyclized COM products exclusively in high to excellent yields (2q-2v), although substrates without substitution on the carbon between the carbonyl and the nitrogen centres required higher catalyst loading for efficient reactions (Scheme 2, see products 2q, 2t and 2u, 25 mol% I2 was used).16 Apart from the driving force due to the formation of acetone as a by-product, stabilizing effect from the conjugation of the aromatic ring to the carbonyl group is also essential for this iodine-catalyzed COM reaction, as non-aromatic substrate 1l was only converted to bicyclic product 2l in poor yield (see page S12 in the SI for more unproductive non-aromatic substrates).16 When the distance between the carbonyl and the olefin moieties was varied with different carbon chain linkers (1’, Scheme 3a), the reaction outcomes changed vastly. γ,δ-Olefin ketone (n = 1) gave the 3,4-dihydro-2H-pyran 4, presumably via a 6-endo-trig cyclization process similar to what were reported by Schindler and coworkers,17 while δ,ε-olefin ketone (n = 2) gave the normal COM products (2a and 2a’). ε,ζ-Olefin ketone (n = 3), on the other hand, did not convert to any cyclized products (see page S19 in the SI for more unproductive ε,ζ-olefin ketone substrates).16 presumably via 6-endo-trig cyclization of 4’ [see ref 17]

(a)

I

O

with n = 1

O

CO2Et

CO2Et

4’

4, 89%

O

with n = 2 COM reaction

10 mol% I2

n

AND CO2Et

neat, rt, 24 h

CO2Et

H

CO2Et

2a 93% (2a : 2a’ = 11 : 1) 2a’

1’ no COM reaction no cyclization reaction ~ 80% recovered SM 1’ (see the SI page S19 for more details)

with n = 3

(b)

O

5

neat, rt, 24-72 h

6

Me

7a, 50%

O

7

R

Me

R

Ar

10 mol% I2

H

Ar

nBu

CO2Me 7b, 34%

3a

7c, 39%

MeO

Ph 7d, 31%

with 6 as ring-opening COM reaction

7e, 50%

O

7f, 21%

O

Scheme 3. (a) COM reactions with different chain lengths; (b) Intermolecular COM reactions

ACS Paragon Plus Environment

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

This newly developed I2-catalyzed protocol was subsequently tested on the intermolecular COM reactions between aromatic aldehydes 5 and isopropylidenyl olefins 6 (Scheme 3b). This type of reaction were reported previously by the Franzen group and our group using tritylium6a and tropylium11 salts as catalysts, respectively. In both cases, only low to moderate efficiencies were obtained, most probably because the intermolecular COM reaction is not as entropically favoured as the intramolecular reaction. Our reactions afforded 31-50% of the intermolecular COM adducts (7a-7d, Scheme 3b), which were comparable results to previous studies.6a,11 The intermolecular ring-opening COM reactions with 1methylcyclopentene, identified as the best substrate for this type of reaction in our previous study,11 also gave low product yields (7e and 7f). Nevertheless, these reaction outcomes are comparable with similar reactions promoted by our tropylium catalyst11 or Schindler’s gallium(III) catalyst,7e which potentially paves the way for future development in this field based on the simplicity of the iodine catalytic system. At this stage, it became very interesting to us to understand how such a simple catalytic system like iodine can promote this previously considered challenging COM reaction. In principle, there are four potential pathways18 how this iodine-promoted reaction can proceed through:13a,13b,13e (i) Brønsted-acid catalyzed9 pathway with HI formed in situ from I2;13e (ii) Radical-initiated pathway from homolytic cleavage of I2 (by light or air);19 (iii) Halogen-bonding assisted pathway with molecular iodine I-I;18a (iv) Lewis-acid catalyzed pathway with iodonium ion (I+) formed in situ by the heterolytic cleavage of I2.18a A halogen-bond activation was recently proposed for the iodine-catalyzed Michael addition to α,β-unsaturated carbonyls as both Brønsted-acid and iodonium-ion catalysis were unlikely under the reaction conditions.18 Thus, we carried out a range of mechanistic studies to get further insights into the mode of activation by iodine (Table 1, see page S24 in the SI for more details).16 Replacement of elemental iodine with HI or KI resulted in non-productive reactions (entries 2-3, Table 1), suggesting that iodide anion and Brønsted acid have no significant role in this reaction. Carrying the reaction with I2 catalyst in the dark (entry 4) or with a radical scavenger (BHT, see entry 8) still gave full conversions of 1a1 to the products while BHT itself has no effect on the reaction (entry 9), hinting that the radical pathway can also be ruled out. A series of reactions with I2 catalyst, where we excluded the presence of moisture and air (entry 5) or just air (entries 6-7), met with significant reductions in the conversions of 1a1. These demonstrated that air did participate in promoting the reaction, most likely through the oxidation of molecular iodine or iodide ion to higher oxidation states. However, when we tried to recreate this oxidative environment in a controlled manner using m-CPBA (entry 10) or hydrogen peroxide

(entry 11), they only resulted in low conversion of 1a1 to the product. Two other iodonium sources, namely NIS and ICl, were also tested as catalysts for this COM reactions (entries 12 and 15). These reactions gave very positive outcomes, supporting the iodonium-catalyzed pathway with I2 catalyst. That it was not the succinimide component catalyzing the reaction is confirmed by entry 14 where NBS could not convert any of 1a1 to the products. Interestingly, when we added DMSO (20 mol%, 2 equiv to NIS catalyst, entry 13, Table 1), a very good iodoniumbinding solvent,20 to the reaction, it completely turned off the catalytic activity of NIS. Similar results were observed with the addition of DMSO (20 mol%) to the original I2 (10 mol%) catalyzed reaction conditions (entry 16, compared to entry 1), which was not too surprising as DMSO can also coordinate to I2.21 The addition of potassium iodide (10 mol%), which could suppress the formation of iodonium ion (I+) and combine with molecular iodine to form triiodide (I3-), also negatively affected the reaction outcomes (entry 17). These experiments once again confirmed the possibility of the iodonium ion being the actual active catalyst in this COM reaction. Due to the presence of moisture, it is possible that some hypoiodous acid (HOI) could form from the disproportionation of the molecular iodine and play some roles in this reaction.22 However, under oxidative conditions such as in entry 11, which should favor the formation of HOI,23 the reaction outcomes were rather poor. We further examined this possibility by carrying out the COM reaction with 10 mol% I2 in the presence of 10 mol% KOH (entry 18, Table 1), which could form potassium hypoiodite (KOI) in situ. This reaction gave a messy mixture with only trace amounts of the wanted COM products, suggesting that a hypoiodous/hypoiodite pathway is unlikely. On the other hand, although we cannot completely rule out the possibility of activation by direct halogen-bonding interaction from molecular iodine based on the experimental investigations,18a entries 19-22 in Table 1 were indicative that halogenbonding activation is unlikely to be the driving force for this type of reaction. Indeed, the use of catalytic or stoichiometric amounts of pentafluorophenyl iodide (entry 19), a frequently used halogen-bond donor,24 did not result in any reaction. Similarly, using triphenylphosphine diiodine (entry 20) and Huber’s monodentate or bidentate iodoazolium salts25 (entries 21-22), all very good halogen-bonding donors but not iodonium sources, did not lead to formation of the COM products. We furthermore carried out kinetic studies to determine the reaction order with respect to the iodine catalyst. The kinetic studies were performed by 1H NMR spectroscopy at different catalyst loadings (8 to 14 mol%), as the reaction was too slow or too fast outside this range.16 For higher iodine loadings (>10 mol%) we

ACS Paragon Plus Environment

Page 4 of 9

Page 5 of 9 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

observed a good linear correlation with a slope of 1 in double-logarithmic plots. The data point for low catalyst loading (8 mol%) deviates slightly from the above and if it was taken into account, a slope close to 2 was obtained (see page S26 in the SI). These outcomes were reproduc-

ible from a triplicate. Based on the kinetic data, it is rather unlikely that I3- or any other higher iodine aggregate is involved in the rate-limiting transition state. Arguably, the reaction order in the catalyst is around 1 and

Table 1. Mechanistic studies of the I2-catalyzed COM reaction with substrate 1a116

O

cat. CO2Et

neat, rt, 24 h

1a1

Entry

a

+ CO2Et 2a (major)

+ CO2Et 2a’ (minor)

O

3a b

Catalyst (10 mol%)

Additive

Conditions

1

I2

-

air, lab light

100%

2

KI

-

air, lab light

no reaction (8%)d

3

HI

-

air, lab light

no reaction

4

I2

-

air, in the dark

100%

5

I2

-

N2, lab light

30% (100%)d

6

I2

water (10 mol%)

N2, lab light

10%

7

I2

water (10 mol%)

N2, in the dark

16%

8

I2

BHT (100 mol%)

air, lab light

100%

9

-

BHT (100 mol%)

air, lab light

no reaction

10

I2

m-CPBA (10 mol%)

N2, lab light

7%

11

Conversion

I2

H2O2 (10 mol%)

N2, lab light

10%

c

NIS

-

air, lab light

48% (100%)d

13 c

NIS

DMSO (20 mol%)

air, lab light

no reaction

14

NBS

-

air, lab light

no reaction

15

ICl

-

air, lab light

62% (100%)d

16

I2

DMSO (20 mol%)

air, lab light

3%

17

I2

KI (10 mol%)

air, lab light

12%

18

I2

KOH (10 mmol%)

air, lab light

messy reaction

-

air, lab light

no reaction

-

air, lab light

no reaction

-

air, lab light

no reaction

-

air, lab light

no reaction

12

F

19

F

I

F

F F

20



Ph3P-I2 nOct

N

21

I N Me



OTf 2 BArF4

22

N N nOct

N I

I

N nOct



a

Reaction conditions: substrate 1a (0.5 mmol) and catalyst/additive were stirred for 24 h at rt; bAir means the reaction was carried out without exclusion of air. Normal lab light means fumehood lighting (4 x 13W fluorescent lamps); cNIS was recrystallized; dConversion in the parentheses were recorded after 72 h.

ACS Paragon Plus Environment

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

the iodine dissociation probably occurs only ‘partially’, e.g., through the formation of very tight ion pairs. Based on these mechanistic studies, we hypothesized a mechanistic pathway in which the I+ ion acted as a Lewis acid to activate the carbonyl moiety and trigger the oxetane formation event (1a1 to 10, Scheme 4). Presumably, the oxetane intermediate 10 subsequently rearranges to the cyclized product 2a similar to how it works with the other acidic catalysts.6a,7a-d,8-9,11 To further probe the reaction mechanism, we additionally employed DFT calculation at the M06-2X/aug-ccpVTZ/IEFPCM//M06-2X/6-311+G(d,p)/IEFPCM with the aug-cc-pVTZ-PP for I] level to validate this proposed mechanism (Scheme 4). In line with the experimental observation that the reverse reaction of a mixture of 2a and 3a (in excess) did not proceed to form 1a1 under the same catalytic conditions (see page S31 in the SI),16, the calculations predict a substantial thermodynamic driving force for the COM reaction (Scheme 4, ΔG = –11.9 kcal mol–1). Furthermore, the computational investigations indicate that a putative halogen-bond activation by I2 requires an activation free energy of 51 kcal mol–1. Compared to the uncatalyzed reaction, this pathway is still favourable (ΔΔG‡ = 6.7 kcal mol–1) but very unlikely due to the high activation free energy (see the computational SI for more details). Our calculations on an iodonium-ion pathway are summarized in Scheme 4. While our calculations predict a very endergonic splitting of molecular iodine due to

Page 6 of 9

unfavorable charge separation (see the computational SI for more details), experimental data suggest equilibrium constants between 1.4 × 10–5 (ΔG = +6.5 kcal mol–1) and 1.8 × 10–10 (ΔG = +13.3 kcal mol–1) for this process.26 As different solvation models (IEFPCM,27a refined SMD,27b,c and CPCM27d–g) resulted in very similar energies, this deviation is most likely caused by charge separation and only to a small extent by an insufficient solvation model. Consequently, we decided to use the latter experimental value for our calculations. As a consequence, the calculated free energy differences will constitute an upper limit for these processes and the real number might even be smaller. Based on our calculations, the rate-limiting transition state is the nucleophilic attack of the alkene onto the activated carbonyl (TS1, Scheme 4), i.e., the first step of the [2+2]-cycloaddition. All subsequent transition states are significantly lower in energy. This energy profile is therefore slightly different to that obtained by Schindler and coworkers for the FeCl3-catalyzed reaction where all transition states were comparable in energy.7a Our calculations furthermore indicate that the addition of iodine to the double bond is thermodynamically unfavorable (ΔG = +6.6 kcal mol–1) and is consequently not observed under the reaction conditions. Therefore, the computational data support the picture of an iodonium-ion pathway as the origin of the catalytic activity of molecular iodine in COM.

I

O

I2

I

I

O

TS1

8

9

TS3

Ph

MeO2C

CO2Me

CO2Me

I O

TS2

Ph

Ph

Ph

I O

I O

13

TS4

O I

I2

3a +

+

Ph Ph

MeO2C

Ph

MeO2C 12

11

10

O

MeO2C

MeO2C

14

14

ΔG / kcal mol–1 TSuncat2 +57.6

+60 TSuncat1 +50.2

+50 +40

TS1 +29.8

+30 +20

10 +22.1 9 +13.3

+10 0 –10

TS2 +22.9

8 + I2 0.0

11 +10.1

TS4 +17.8

TS3 +16.1 12 +13.3

+2.7 13 + 14 –2.4 14 + 3a + I2 –11.9

–20

Scheme 4. Gibbs free-energy profile for the uncatalyzed and the iodonium-ion-catalyzed carbonyl-olefin metathesis.

ACS Paragon Plus Environment

Page 7 of 9 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

CONCLUSION We have developed a new practical protocol for the intramolecular carbonyl-olefin metathesis reactions using elemental iodine as catalyst. This method is easy to set up with very mild reaction conditions and involves an inexpensive and benign catalyst while offering comparable outcomes to previously reported procedures. Preliminary mechanistic studies revealed the important role of the in situ generated iodonium ion, which might promote the COM reaction via the activation of the carbonyl moiety. Further applications of this new method in organic synthesis are ongoing and will be reported shortly. AUTHOR INFORMATION Corresponding Author

T.V. Nguyen, email: [email protected] Funding Sources The authors declare no competing financial interests.

ASSOCIATED CONTENT Supporting Information Experimental procedures, characterization data, NMR spectra and details of the computational investigations are available free of charge on the ACS Publications website.

Keywords Carbonyl-olefin metathesis; olefination; iodine; iodonium; catalysis; metal-free

A CKNOWLEDGMENT This work is financially supported by Australian Research Council (grant DE150100517 to TVN). Additional support from the Fonds der chemischen Industrie (Liebig scholarship to MB) and the DFG (BR 5154/2-1) is gratefully acknowledged. We are grateful to the Regional Computing Center of the University of Cologne for providing computing time of the DFG-funded High Performance Computing (HPC) System CHEOPS as well as for their support.

REFERENCES (with titles) (1) (a) Lee, A.-L. Organocatalyzed Carbonyl–Olefin Metathesis, Angew. Chem. Int. Ed. 2013, 52, 4524; (b) Hennessy, E. T.; Jacobsen, E. N. Organometallic chemistry: A new metathesis, Nat Chem 2016, 8, 741; (c) Saá, C. Iron(III)-Catalyzed Ring-Closing Carbonyl–Olefin Metathesis, Angew. Chem. Int. Ed. 2016, 55, 10960. (2) (a) Grubbs, R. H.; Chang, S. Recent advances in olefin metathesis and its application in organic synthesis, Tetrahedron 1998, 54, 4413; (b) Fürstner, A. Olefin Metathesis and Beyond, Angew. Chem. Int. Ed. 2000, 39, 3012; (c) Chauvin, Y. Olefin Metathesis: The Early Days (Nobel Lecture), Angew. Chem. Int. Ed. 2006, 45, 3740; (d) Grubbs, R. H. OlefinMetathesis Catalysts for the Preparation of Molecules and Materials (Nobel Lecture), Angew. Chem. Int. Ed. 2006, 45, 3760; (e) Schrock, R. R. Multiple Metal–Carbon Bonds for Catalytic Metathesis Reactions (Nobel Lecture), Angew. Chem. Int. Ed. 2006, 45, 3748; (f) Hoveyda, A. H.; Zhugralin, A. R. The remarkable metal-catalysed olefin metathesis reaction, Nature 2007, 450, 243. (3) (a) Fu, G. C.; Grubbs, R. H. Synthesis of cycloalkenes via alkylidene-mediated olefin metathesis and carbonyl olefination, J. Am. Chem. Soc. 1993, 115, 3800; (b) Jackson, A. C.; Goldman, B. E.; Snider, B. B. Intramolecular and intermolecular Lewis acid catalyzed ene

reactions using ketones as enophiles, J. Org. Chem. 1984, 49, 3988; (c) Khripach, V. A.; Zhabinskii, V. N.; Kuchto, A. I.; Zhiburtovich, Y. Y.; Gromak, V. V.; Groen, M. B.; van der Louw, J.; de Groot, A. Intramolecular cycloaddition/cycloreversion of (E)-3β,17β-diacetoxy5,10-secoandrost-1(10)-en-5-one, Tetrahedron Lett. 2006, 47, 6715; (d) Schopov, I.; Jossifov, C. A carbonyl-olefin exchange reaction — new route to polyconjugated polymers, 1. A new synthesis of polyphenylacetylene, Makromol. Chem., Rapid Commun. 1983, 4, 659; (e) Soicke, A.; Slavov, N.; Neudörfl, J.-M.; Schmalz, H.-G. Metal-Free Intramolecular Carbonyl-Olefin Metathesis of ortho-Prenylaryl Ketones, Synlett 2011, 2487; (f) van Schaik, H.-P.; Vijn, R.-J.; Bickelhaupt, F. Acid-Catalyzed Olefination of Benzaldehyde, Angew. Chem. Int. Ed. 1994, 33, 1611. (4) (a) Stille, J. R.; Grubbs, R. H. Synthesis of (.+-.)-.DELTA.9,12capnellene using titanium reagents, J. Am. Chem. Soc. 1986, 108, 855; (b) Stille, J. R.; Santarsiero, B. D.; Grubbs, R. H. Rearrangement of bicyclo[2.2.1]heptane ring systems by titanocene alkylidene complexes to bicyclo[3.2.0]heptane enol ethers. Total synthesis of (.+-.)-.DELTA.9(12)capnellene, J. Org. Chem. 1990, 55, 843; (c) Nicolaou, K. C.; Postema, M. H. D.; Claiborne, C. F. Olefin Metathesis in Cyclic Ether Formation. Direct Conversion of Olefinic Esters to Cyclic Enol Ethers with TebbeType Reagents, J. Am. Chem. Soc. 1996, 118, 1565; (d) Rainier, J. D.; Allwein, S. P.; Cox, J. M. C-Glycosides to Fused Polycyclic Ethers. A Formal Synthesis of (±)-Hemibrevetoxin B, J. Org. Chem. 2001, 66, 1380; (e) Majumder, U.; Rainier, J. D. Olefinic-ester cyclizations using Takai– Utimoto reduced titanium alkylidenes, Tetrahedron Lett. 2005, 46, 7209; (f) Iyer, K.; Rainier, J. D. Olefinic Ester and Diene Ring-Closing Metathesis Using a Reduced Titanium Alkylidene, J. Am. Chem. Soc. 2007, 129, 12604; (g) Heller, S. T.; Kiho, T.; Narayan, A. R. H.; Sarpong, R. Protic-Solvent-Mediated Cycloisomerization of Quinoline and Isoquinoline Propargylic Alcohols: Syntheses of (±)-3Demethoxyerythratidinone and (±)-Cocculidine, Angew. Chem. Int. Ed. 2013, 52, 11129; (h) Hong, B.; Li, H.; Wu, J.; Zhang, J.; Lei, X. Total Syntheses of (−)-Huperzine Q and (+)-Lycopladines B and C, Angew. Chem. Int. Ed. 2015, 54, 1011. (5) (a) Griffith, A. K.; Vanos, C. M.; Lambert, T. H. Organocatalytic Carbonyl-Olefin Metathesis, J. Am. Chem. Soc. 2012, 134, 18581; (b) Hong, X.; Liang, Y.; Griffith, A. K.; Lambert, T. H.; Houk, K. N. Distortion-accelerated cycloadditions and strain-release-promoted cycloreversions in the organocatalytic carbonyl-olefin metathesis, Chem. Sci. 2014, 5, 471. (6) (a) Veluru Ramesh, N.; Bah, J.; Franzén, J. Direct Organocatalytic Oxo-Metathesis, a trans-Selective Carbocation-Catalyzed Olefination of Aldehydes, Eur. J. Org. Chem. 2015, 1834; (b) Ni, S.; Franzén, J. Carbocation catalysed ring closing aldehyde–olefin metathesis, Chem. Commun. 2018, 54, 12982. (7) (a) Ludwig, J. R.; Zimmerman, P. M.; Gianino, J. B.; Schindler, C. S. Iron(III)-catalysed carbonyl–olefin metathesis, Nature 2016, 533, 374; (b) Ludwig, J. R.; Phan, S.; McAtee, C. C.; Zimmerman, P. M.; Devery, J. J.; Schindler, C. S. Mechanistic Investigations of the Iron(III)-Catalyzed Carbonyl-Olefin Metathesis Reaction, J. Am. Chem. Soc. 2017, 139, 10832; (c) McAtee, C. C.; Riehl, P. S.; Schindler, C. S. Polycyclic Aromatic Hydrocarbons via Iron(III)-Catalyzed Carbonyl–Olefin Metathesis, J. Am. Chem. Soc. 2017, 139, 2960; (d) Groso, E. J.; Golonka, A. N.; Harding, R. A.; Alexander, B. W.; Sodano, T. M.; Schindler, C. S. 3-Aryl-2,5-Dihydropyrroles via Catalytic Carbonyl-Olefin Metathesis, ACS Catal. 2018, 8, 2006; (e) Albright, H.; Vonesh, H. L.; Becker, M. R.; Alexander, B. W.; Ludwig, J. R.; Wiscons, R. A.; Schindler, C. S. GaCl3Catalyzed Ring-Opening Carbonyl–Olefin Metathesis, Org. Lett. 2018, 20, 4954; (f) Ludwig, J. R.; Watson, R. B.; Nasrallah, D. J.; Gianino, J. B.; Zimmerman, P. M.; Wiscons, R. A.; Schindler, C. S. Interrupted carbonylolefin metathesis via oxygen atom transfer, Science 2018, 361, 1363. (8) Ma, L.; Li, W.; Xi, H.; Bai, X.; Ma, E.; Yan, X.; Li, Z. FeCl3Catalyzed Ring-Closing Carbonyl–Olefin Metathesis, Angew. Chem. Int. Ed. 2016, 55, 10410. (9) Catti, L.; Tiefenbacher, K. Brønsted Acid‐Catalyzed Carbonyl‐Olefin Metathesis inside a Self‐Assembled Supramolecular Host, Angew. Chem. Int. Ed. 2018, 57, 14589. (10) (a) Nguyen, T. V.; Bekensir, A. Aromatic Cation Activation: Nucleophilic Substitution of Alcohols and Carboxylic Acids, Org. Lett. 2014, 16, 1720; (b) Nguyen, T. V.; Hall, M. Activation of DMSO for Swern-type oxidation by 1,1-dichlorocycloheptatriene, Tetrahedron Lett. 2014, 55, 6895; (c) Nguyen, T. V.; Lyons, D. J. M. A novel aromatic carbocation-based coupling reagent for esterification and amidation reactions, Chem. Commun. 2015, 51, 3131; (d) Lyons, D. J. M.; Crocker,

ACS Paragon Plus Environment

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

R. D.; Blümel, M.; Nguyen, T. V. Promotion of Organic Reactions by Non-Benzenoid Carbocylic Aromatic Ions, Angew. Chem. Int. Ed. 2017, 56, 1466; (e) Lyons, D. J. M.; Crocker, R. D.; Enders, D.; Nguyen, T. V. Tropylium salts as efficient organic Lewis acid catalysts for acetalization and transacetalization reactions in batch and flow, Green Chem. 2017, 3993; (f) Oss, G.; de Vos, S. D.; Luc, K. N. H.; Harper, J. B.; Nguyen, T. V. Tropylium-Promoted Oxidative Functionalization of Tetrahydroisoquinolines, J. Org. Chem. 2018, 83, 1000; (g) Oss, G.; Ho, J.; Nguyen Thanh, V. Tropylium Ion Catalyzes Hydration Reactions of Alkynes, Eur. J. Org. Chem. 2018, 3974; (h) Lyons, D.; Crocker, R.; Nguyen Thanh, V. Stimuli‐Responsive Organic Dyes with Tropylium Chromophore, Chem. Eur. J. 2018, 24, 10959; (i) Hussein, M. A.; Huynh, V. T.; Hommelsheim, R.; Koenigs, R. M.; Nguyen, T. V. An efficient method for retro-Claisen-type C–C bond cleavage of diketones with tropylium catalyst, Chem. Commun. 2018, 54, 12970. (11) Tran, U. P. N.; Oss, G.; Pace, D. P.; Ho, J.; Nguyen, T. V. Tropylium-Promoted Carbonyl-Olefin Metathesis Reactions, Chem. Sci. 2018, 9, 5145. (12) Ludwig, J. R.; Schindler, C. S. Lewis Acid Catalyzed Carbonyl– Olefin Metathesis, Synlett 2017, 28, 1501. (13) (a) Jereb, M.; Vražič, D.; Zupan, M. Iodine-catalyzed transformation of molecules containing oxygen functional groups, Tetrahedron 2011, 67, 1355; (b) Ren, Y.-M.; Cai, C.; Yang, R.-C. Molecular iodine-catalyzed multicomponent reactions: an efficient catalyst for organic synthesis, RSC Adv. 2013, 3, 7182; (c) Schindler, S.; Huber, S. M. In Halogen Bonding II: Impact on Materials Chemistry and Life Sciences; Metrangolo, P., Resnati, G., Eds.; Springer International Publishing: Cham, 2015, p 167; (d) Bulfield, D.; Huber Stefan, M. Halogen Bonding in Organic Synthesis and Organocatalysis, Chem. Eur. J. 2016, 22, 14434; (e) Breugst, M.; von der Heiden, D. Mechanisms in Iodine Catalysis, Chem. Eur. J. 2018, 24, 9187. (14) Hepperle, S. S.; Li, Q.; East, A. L. L. Mechanism of Cis/Trans Equilibration of Alkenes via Iodine Catalysis, J. Phys. Chem. A 2005, 109, 10975. (15) Takeda, Y.; Kajihara, R.; Kobayashi, N.; Noguchi, K.; Saito, A. Molecular-Iodine-Catalyzed Cyclization of 2-Alkynylanilines via Iodocyclization–Protodeiodination Sequence, Org. Lett. 2017, 19, 6744. (16) See the Supporting Information for more details. (17) (a) Watson, R. B.; Golonka, A. N.; Schindler, C. S. Iron(III) Chloride Catalyzed Formation of 3,4-Dihydro-2H-pyrans from αAlkylated 1,3-Dicarbonyls. Selective Synthesis of α- and β-Lapachone, Org. Lett. 2016, 18, 1310; (b) Watson, R. B.; Schindler, C. S. IronCatalyzed Synthesis of Tetrahydronaphthalenes via 3,4-Dihydro-2H-pyran Intermediates, Org. Lett. 2018, 20, 68. (18) (a) von der Heiden, D.; Bozkus, S.; Klussmann, M.; Breugst, M. Reaction Mechanism of Iodine-Catalyzed Michael Additions, J. Org. Chem. 2017, 82, 4037; (b) Breugst, M.; Detmar, E.; von der Heiden, D. Origin of the Catalytic Effects of Molecular Iodine: A Computational Analysis, ACS Catal. 2016, 6, 3203. (19) (a) Gardner, J. M.; Abrahamsson, M.; Farnum, B. H.; Meyer, G. J. Visible Light Generation of Iodine Atoms and I−I Bonds: Sensitized I− Oxidation and I3− Photodissociation, J. Am. Chem. Soc. 2009, 131, 16206; (b) Saiz-Lopez, A.; Plane, J. M. C.; Baker, A. R.; Carpenter, L. J.; von Glasow, R.; Gómez Martín, J. C.; McFiggans, G.; Saunders, R. W. Atmospheric Chemistry of Iodine, Chem. Rev. 2012, 112, 1773. (20) Ashikari, Y.; Shimizu, A.; Nokami, T.; Yoshida, J.-i. Halogen and Chalcogen Cation Pools Stabilized by DMSO. Versatile Reagents for Alkene Difunctionalization, J. Am. Chem. Soc. 2013, 135, 16070.

(21) Laurence, C.; Graton, J.; Berthelot, M.; El Ghomari, M. J. The Diiodine Basicity Scale: Toward a General Halogen-Bond Basicity Scale, Chem. Eur. J. 2011, 17, 10431. (22) Becker, P.; Duhamel, T.; Stein, C. J.; Reiher, M.; Muñiz, K. Cooperative Light-Activated Iodine and Photoredox Catalysis for the Amination of C −H Bonds, Angew. Chem. Int. Ed. 2017, 56, 8004. (23) Uyanik, M.; Hayashi, H.; Ishihara, K. High-turnover hypoiodite catalysis for asymmetric synthesis of tocopherols, Science 2014, 345, 291. (24) Cavallo, G.; Metrangolo, P.; Milani, R.; Pilati, T.; Priimagi, A.; Resnati, G.; Terraneo, G. The Halogen Bond" Chem. Rev. 2016, 116, 2478. (25) (a) Kniep, F.; Jungbauer, S. H.; Zhang, Q.; Walter, S. M.; Schindler, S.; Schnapperelle, I.; Herdtweck, E.; Huber, S. M. Organocatalysis by Neutral Multidentate Halogen-Bond Donors, Angew. Chem. Int. Ed. 2013, 52, 7028; (b) Jungbauer, S. H.; Walter, S. M.; Schindler, S.; Rout, L.; Kniep, F.; Huber, S. M. Activation of a carbonyl compound by halogen bonding, Chem. Commun. 2014, 50, 6281; (c) Gliese, J.-P.; Jungbauer, S. H.; Huber, S. M. A halogen-bonding-catalyzed Michael addition reaction, Chem. Commun. 2017, 53, 12052; (d) Dreger, A.; Engelage, E.; Mallick, B.; Beer, P. D.; Huber, S. M. The role of charge in 1,2,3-triazol(ium)-based halogen bonding activators, Chem. Commun. 2018, 54, 4013; (e) Schulz, N.; Sokkar, P.; Engelage, E.; Schindler, S.; Erdelyi, M.; Sanchez-Garcia, E.; Huber, S. M. The Interaction Modes of Haloimidazolium Salts in Solution, Chem. Eur. J. 2018, 24, 3464; (f) Stoesser, J.; Rojas, G.; Bulfield, D.; Hidalgo, P. I.; Pasán, J.; Ruiz-Pérez, C.; Jiménez, C. A.; Huber, S. M. Halogen bonding two-point recognition with terphenyl derivatives, New J. Chem. 2018, 42, 10476; (g) von der Heiden, D.; Detmar, E.; Kuchta, R.; Breugst, M. Activation of Michael Acceptors by Halogen-Bond Donors, Synlett 2018, 14, 1307.

(26) (a) Nigretto, J. M.; Jozefowicz, M. Etude electrochimique de l'iode et de ses derives en solution dans la pyridine variation des proprietes d'oxydo-reduction en fonction du pH, Electrochim. Acta 1974, 19, 809; (b) Aronson, S.; Epstein, P.; Aronson, D. B.; Wieder, G. Ionic equilibriums in pyridine-iodine solutions, J. Phys. Chem. 1982, 86, 1035. (27) (a) Cancès, E.; Mennucci, B.; Tomasi, J. A new integral equation formalism for the polarizable continuum model: Theoretical background and applications to isotropic and anisotropic dielectrics, J. Chem. Phys. 1997, 107, 3032; (b) Marenich, A. V.; Cramer, C. J.; Truhlar, D. G. Universal Solvation Model Based on Solute Electron Density and on a Continuum Model of the Solvent Defined by the Bulk Dielectric Constant and Atomic Surface Tensions, J. Phys. Chem. B 2009, 113, 6378; (c) Engelage, E.; Schulz, N.; Heinen, F.; Huber, S. M.; Truhlar, D. G.; Cramer, C. J. Refined SMD Parameters for Bromine and Iodine Accurately Model Halogen-Bonding Interactions in Solution, Chem. Eur. J. 2018, 24, 15983; (d) Klamt, A.; Schüürmann, G. COSMO: a new approach to dielectric screening in solvents with explicit expressions for the screening energy and its gradient, J. Chem. Soc., Perkin Trans. 2 1993, 799; (e) Andzelm, J.; Kölmel, C.; Klamt, A. Incorporation of solvent effects into density functional calculations of molecular energies and geometries, J. Chem. Phys. 1995, 103, 9312; (f) Barone, V.; Cossi, M. Quantum Calculation of Molecular Energies and Energy Gradients in Solution by a Conductor Solvent Model, J. Phys. Chem. A 1998, 102, 1995; (g) Cossi, M.; Rega, N.; Scalmani, G.; Barone, V. Energies, structures, and electronic properties of molecules in solution with the C-PCM solvation model, J. Comput. Chem. 2003, 24, 669.

ACS Paragon Plus Environment

Page 8 of 9

Page 9 of 9 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

Insert Table of Contents artwork here

Carbonyl-Olefin Metathesis (COM) Reaction R2 R1

O

R1

cat. I2 10 mol%

R2

O neat, rt, 24 h 27 examples 21-96% practical on gram-scale

ACS Paragon Plus Environment

9