Download Palladium(II)-Catalyzed Oxidative Cyclization Strategies Andreas K. Å. Persson

Document related concepts

Fischer–Tropsch process wikipedia , lookup

Enantioselective synthesis wikipedia , lookup

Marcus theory wikipedia , lookup

Haloalkane wikipedia , lookup

Kinetic resolution wikipedia , lookup

Woodward–Hoffmann rules wikipedia , lookup

Physical organic chemistry wikipedia , lookup

Asymmetric induction wikipedia , lookup

George S. Hammond wikipedia , lookup

Tiffeneau–Demjanov rearrangement wikipedia , lookup

Diels–Alder reaction wikipedia , lookup

Alkene wikipedia , lookup

Vinylcyclopropane rearrangement wikipedia , lookup

Elias James Corey wikipedia , lookup

Ene reaction wikipedia , lookup

Hofmann–Löffler reaction wikipedia , lookup

Baylis–Hillman reaction wikipedia , lookup

Discodermolide wikipedia , lookup

Hydroformylation wikipedia , lookup

Ring-closing metathesis wikipedia , lookup

Petasis reaction wikipedia , lookup

Wolff–Kishner reduction wikipedia , lookup

Strychnine total synthesis wikipedia , lookup

Stille reaction wikipedia , lookup

Transcript
Palladium(II)-Catalyzed Oxidative
Cyclization Strategies
Selective Formation of New C-C and C-N Bonds
Andreas K. Å. Persson
© Andreas Persson, Stockholm 2012
Cover picture by Johanna Persson
ISBN: 978-91-7447-495-4
Printed in Sweden by US-AB, Stockholm 2012
Distributor: Department of Organic Chemistry, Stockholm University
ii
Where is the rest?
[Unknown]
iii
iv
Abstract
The use of transition metal catalysts has proven to be one of the most diverse
tools for the mild and selective formation of carbon-carbon bonds. In
particular palladium-catalyzed cross-coupling reactions have revolutionized
the field.
The main focus of this thesis has been directed towards preparation and
oxidative carbocyclization of en-, dien- and aza-enallenes.
In the first part of this thesis, a stereoselective oxidative carbocyclization
of dienallenes was realized. By employing cheap and readily available
palladium trifluoroacetate we were able to efficiently cyclize a variety of
dienallenes into hydroxylated carbocycles in high yield and high selectivity.
This oxidative process was compatible with two different reoxidation
protocols: one relying on p-benzoquinone (BQ) as the oxidant and the other
employing molecular oxygen as the oxidant.
In the second part of the thesis the carbocyclization methodology was
extended to include carbocyclization of aza-enallenes. This was achieved in
two distinct steps. First, a copper-catalyzed coupling of allylic sulfonamides
with bromoallenes was developed, giving access to the corresponding azaenallenes. Subjecting these substrates to catalytic amounts of palladium
acetate, along with BQ as the oxidant, rendered N-heterocycles in good
yield. The reactivity of these N-heterocycles towards activated dienophiles
was later exploited in a tandem (aerobic) oxidative carbocyclization/DielsAlder reaction.
The third topic involves efficient oxidative arylative/borylative
carbocyclization of enallenes. These reactions, catalyzed by palladium
acetate, relies on transmetallation of a (σ-alkyl)palladium(II) intermediate
with diboranes or arylboronic acids. With this novel methodology we were
able to obtain an array of arylated or borylated carbocycles, as single
diastereomers, in high yield.
Finally, we developed a palladium(II)-catalyzed cyclization of allylic
carbamates. This mild, operationally, simple and scalable catalytic reaction
opens up access to an array of oxazolidinones in high yield and excellent
diastereoselectivity.
v
vi
List of Publications
This thesis is based on the following publications, referred to in the text by
their Roman numerals I-VI. Reprints were made with the kind permission of
the publisher (Appendix A).
I.
Water As Nucleophile in Palladium-Catalyzed Oxidative
Carbohydroxylation of Allene-Substituted Conjugated Dienes
Piera, J.; Persson, A.; Caldentey, X.; Bäckvall, J. -E.
J. Am. Chem. Soc. 2007, 129, 14120-14121
II.
Copper Catalyzed N-Allenylation of Allylic Sulfonamides
Persson, A. K. Å.; Johnston, E. V.; Bäckvall, J. -E.
Org. Lett. 2009, 11, 3814-3817
III.
Palladium(II)-Catalyzed Oxidative Carbocyclization of AzaEnallenes
Persson, A. K Å.; Bäckvall, J. -E.
Angew. Chem. Int. Ed. 2010, 49, 4624-4627
IV.
Palladium-Catalyzed Oxidative Borylative Carbocyclization of
Enallenes
Persson, A. K. Å.; Jiang, T.; Johnson, M. T.; Bäckvall, J. -E.
Angew. Chem. Int. Ed. 2011, 50, 6155-6159
V.
Palladium-Catalyzed Oxidative Carbocyclization/Arylation of
Enallenes
Jiang, T.; Persson, A. K. Å.; Bäckvall, J. -E.
Org. Lett. 2011, 13, 5838-5841
VI.
Palladium(II)-Catalyzed Oxidative Cyclization of Allylic
Tosylcarbamates – Scope, Derivatization and Mechanistic
Aspects
Joosten, A. Y.; Persson, A. K. Å.; Millet, R.; Johnson, M. T.;
Bäckvall, J. -E. Submitted for publication.
vii
Related papers by the author, but not included as part of this thesis:
Palladium-Catalyzed Oxidative Arylating Carbocyclization of
Allenynes
Deng, Y.; Bartholomeyzik, T.; Persson, A. K. Å.; Sun, J.;
Bäckvall, J. -E.
Angew. Chem. Int. Ed. 2012, 51, 2703-2707
Enantioselective Synthesis of a-Methyl Carboxylic
Acids via Chemoenzymatic Dynamic Kinetic Resolution
Thalén, L. K.; Sumic, A.; Bogár, K.; Norinder, J.; Persson, A. K.
Å.; Bäckvall, J. -E.
J. Org. Chem. 2010, 75, 6842–6847
Palladium-Catalyzed Oxidative Carbocyclization –
Diversity in Carbon-Carbon Bond formation
Deng, Y†.; Persson, A. K. ņ.; Bäckvall, J. -E.
Review Article, Submitted for publication.
†
Authors contributed equally to this work
viii
ix
Table of Contents
Abstract ....................................................................................................................... v
List of Publications .................................................................................................. vii
Table of Contents ........................................................................................................ x
Abbreviations .......................................................................................................... xiii
1. Introduction............................................................................................................. 1
1.1 Catalysis ........................................................................................................... 1
1.2 Transition Metal Catalysis ............................................................................... 1
1.3 Palladium Catalysis - Background ................................................................... 2
1.3.1 Palladium Catalysis – Fundamental Reactions ........................................ 3
1.4 Palladium(II)-Catalyzed Oxidations Not Involving Carbon-Carbon Bond
Formation ............................................................................................................... 3
1.4.1 Reoxidation of Palladium(0) Using Different Oxidation Systems ........... 5
1.4.2 Aerobic Reoxidation of Palladium(0) ...................................................... 5
1.5 Palladium(II)-Catalyzed Oxidations Involving Carbon-Carbon Bond
Formation ............................................................................................................... 7
1.5.1 Palladium(II-IV)-Catalyzed Oxidative Carbocyclizations ....................... 8
1.5.2 Palladium(II)-Catalyzed Oxidative Carbocyclizations Involving Allenes
........................................................................................................................ 10
2. Preparation of Starting Materials .......................................................................... 13
2.1 Representative Materials Prepared for Paper I............................................... 13
2.2 Representative Materials Prepared for Paper II ............................................. 13
2.3 Representative Materials Prepared for Paper IV-V........................................ 14
3. Thesis Aim and Objectives ................................................................................... 15
4.Water As Nucleophile in Palladium(II)-Catalyzed Oxidative Carbocyclizations
(Paper I) .................................................................................................................... 16
4.1 Results and Discussion .................................................................................. 17
4.1.1 Attempts to Use Amide Nucleophiles.................................................... 17
4.1.2 Water as Nucleophile ............................................................................. 18
4.1.3 Reaction Scope ...................................................................................... 18
4.1.4 Mechanistic Discussion ......................................................................... 20
4.1.5 The Hydrolysis Dilemma ....................................................................... 22
x
4.2 Conclusions.................................................................................................... 23
5. Copper(I)-Catalyzed N-Allenylation of Allylic Sulfonamides (Paper II) ............. 24
5.1 Results and Discussion .................................................................................. 26
5.1.1 Identifying a Model Substrate and Initial Optimization ....................... 26
5.1.2 Allenylation of Allylic Sulfonamides .................................................... 27
5.1.3 Structural Limitations ............................................................................ 29
5.1.4 Allenylation of Non-Allylic Sulfonamides ............................................ 29
5.1.5 Stability of the Aza-Enallenes ............................................................... 30
5.1.6 Mechanistic Discussion ......................................................................... 31
5.2 Conclusions.................................................................................................... 32
6. Palladium(II)-Catalyzed Oxidative Carbocyclization of Aza-Enallenes (Paper III)
.................................................................................................................................. 33
6.1. Results and Discussion ................................................................................. 34
6.1.1 Optimization of the Reaction Conditions.............................................. 34
6.1.2 In situ 1H NMR Experiments ................................................................. 35
6.1.3 Scope and Limitations ........................................................................... 36
6.1.4 Aerobic Reoxidation and Tandem Diels-Alder Reaction ...................... 38
6.1.5 Stereochemical Assignment ................................................................... 40
6.1.6 Mechanistic Discussion ......................................................................... 40
6.2 Conclusions.................................................................................................... 42
7. Palladium(II)-catalyzed Oxidative Carboborylation/Carboarylation of Enallenes
(Paper IV-V) ............................................................................................................. 43
7.1 Palladium(II)-Catalyzed Oxidative Carbocyclization/Borylation of Enallenes.
............................................................................................................................. 45
7.1.1
Development
of
a
Catalytic
System
for
Oxidative
Carbocyclization/Borylation of Enallenes ...................................................... 45
7.1.2 Investigation of the Scope and Limitations............................................ 47
7.1.3 Stereochemical Assignment of the Borylated Carbocycles ................... 50
7.1.4 Mechanistic Discussion ......................................................................... 51
7.2 Palladium(II)-Catalyzed Oxidative Carbocyclization/Arylation of Enallenes.
............................................................................................................................. 54
7.2.1 Arylboronic acids as Transmetallation Reagent. ................................... 54
7.2.2 Scope and Limitations ........................................................................... 55
7.2.3 Mechanistic Discussion ......................................................................... 59
7.3 Conclusions on the Oxidative Arylative/Borylative Carbocyclization of
Enallenes. ............................................................................................................. 60
8. Palladium(II)-Catalyzed Formation of Oxazolidinones Through Cyclization of
Allylic Tosylcarbamates – Scope, Mechanistic Aspects and Further Derivatization.
(Paper -VI) ................................................................................................................ 61
8.1 Results and Discussion .................................................................................. 63
8.1.1 Finding Efficient Reaction Conditions .................................................. 63
xi
8.1.2 Scope and Limitations ........................................................................... 65
8.1.3 Biomimetic Reoxidation and Scale Up .................................................. 67
8.1.4 Further Transformations of the Oxazolidinones .................................... 68
8.1.5 Stereochemical Analysis and X-ray Structures ...................................... 70
8.1.6 Deuterium Labeling and Mechanistic Discussion.................................. 71
8.2 Conclusions.................................................................................................... 74
10 Concluding Remarks ............................................................................................ 75
Reprint Permissions (Appendix A) ........................................................................... 77
Acknowledgements ................................................................................................... 78
References................................................................................................................. 81
xii
Abbreviations
Abbreviations and acronyms are used in agreement with the standard of the
subject.1 Only nonstandard and unconventional ones that appear in the thesis
are listed here.
Acac
B2pin2
BQ
Cu(OTf)2
CuTc
DCM
DEEDA
DIBAL-H
DMBQ
DMEDA
dppe
ETM
EtOAc
FePc
HOAc
HQ
NHC
OAc
PIDA
PhBQ
PIFA
TBAF
TBDMSCl
TBDPS
TEMPO
TIPS-EBX
Acetylacetone
Bis(pinacolato)diboron
1,4-Benzoquinone
Copper(II)-trifluoromethylsulfonate
Copper(I)-thiophene 2-carboxylate
Dichloromethane
N,N'-Diethylethylenediamine
Diisobutylaluminium hydride
2,6-Dimethylbenzoquinone
N,N'-Dimethylethylenediamine
1,2-Bis(diphenylphosphino)ethane
Electron transfer mediator
Ethyl acetate
Iron(II)-phthalocyanine
Acetic acid
1,4-Hydroquinone
N-heterocyclic carbene
Acetate
Iodobenzene I,I-bis(diacetate)
2-Phenylbenzo-1,4-quinone
Iodobenzene I,I-bis(trifluoroacetate)
Tetrabutylammonium fluoride
tert-Butyldimethylsilyl chloride
tert-Butyldiphenylsilyl
2,2,6,6-Tetramethyl-1-piperidinyloxy
1-[(Triisopropylsilyl)ethynyl]-1,2-benziodoxol-3(1H)-one
xiii
xiv
1. Introduction
1.1 Catalysis
A catalyst is a substance that accelerates a chemical reaction by lowering
the activation energy without undergoing any net-reaction itself.2 The term
“catalysis” was first introduced by the Swedish chemist J. J Berzelius
already in 1836.3 Under ideal conditions, the catalyst is not consumed during
the reaction and theoretically it can be reused in an infinite number of cycles.
As a result, catalytic processes have the potential to become environmentally
friendly as well as very cost effective. Although there are numerous
industrial catalytic processes currently in operating such as the Monsanto-4
and Cativa-5 acetic acid processes, there is still a demand for more
environmentally friendly chemical processes employing sustainable and
robust catalysts. Owing to this demand, the field of catalysis has received
massive attention across all fields of chemistry.
Catalysis can be roughly divided into several different subgroups
depending on the nature of the actual catalyst. Some of the major individual
areas are biocatalysis, Lewis-acid/base catalysis, organocatalysis and
transition metal catalysis6. All four catalytic areas may be further categorized
depending on if they act as homogeneous or heterogeneous catalysts.7
1.2 Transition Metal Catalysis
The use of transition metals in catalysis has received considerable attention
during the past century. It all started with a discovery made by Humphry
Davy in 1817; he discovered that a preheated platinum wire became red hot
when exposed to an atmosphere of a combustible gas and air. What he had
discovered was the first example of a heterogeneous catalytic oxidation.8
Many of the properties of the transitions metals that make them useful in
catalysis arise from their ability to exist in different oxidation states,
allowing the metal to interact in a specific manner with a substrate
depending on electron configuration.
1
One particular group of transition metals has been of special interest to
chemists, namely the platinum group: ruthenium (Ru), rhodium (Rh),
palladium (Pd), osmium (Os), iridium (Ir) and platinum (Pt). The platinum
group metals have shown outstanding catalytic activities in a broad range of
chemical reactions ranging from oxidations, reductions, carbon-carbon
couplings, carbon-heteroatom couplings, and isomerizations.6,9
1.3 Palladium Catalysis - Background
Although Kolbe10 reported the first palladium mediated catalytic
hydrogenation back in 1871, the major breakthrough in palladium catalysis
is generally considered to be the development of the Wacker-process in 1959
(Scheme 1, Eq 1).11 This industrially important process allows for selective
oxidation of ethylene to acetaldehyde in the presence of catalytic amounts of
PdCl2 and CuCl2 together with molecular oxygen.12 The discovery of this
reaction initiated an era of intense research, resulting in the development of
some very versatile palladium-catalyzed processes, such as the palladiumcatalyzed cross-coupling reactions13 developed by Heck,14,15 Negishi16 and
Suzuki17 for which they obtained the 2010 Nobel Prize (Scheme 1, Eq 2).
Scheme 1. Wacker oxidation (Eq. 1) and the Suzuki-Miyaura cross-coupling (Eq. 2)
The versatility of palladium catalysts is directly related to some of its
unique physical properties. First, palladium interconverts between the 0 and
+II oxidation state with relative ease (d10and d8, respectively). As a rule of
thumb, palladium in oxidation state 0 is considered nucleophilic while the
+II oxidation state is considered electrophilic. Until recently, the majority of
research conducted on palladium catalysis has involved palladium in the 0
and +II oxidation state. Nevertheless, species having oxidation states +I,
+III18 and +IV do exist, and have recently found application in catalysis.19,20
Palladium is moderately electronegative (2.2) and has a rather “soft”
character. Properties like these make the metal center less prone to interact
with electronegative hard groups (such as alcohols). Instead it has a higher
affinity towards “soft” σ- and π-donors. Consequently, alkenes, alkynes and
allenes readily form π-complexes with palladium.21
2
In addition, palladium is prone to form complexes with for example
amines, N-heterocyclic carbenes (NHC`s) and phosphines, to mention a few.
Variation of the ligand on palladium can completely alter the reactivity or be
used to tune the catalyst for each process, regardless of the oxidation state
involved. Consequently, chemists have been able to develop very robust
catalytic systems as well as achieving asymmetric transformations with
chiral ligands.22,23
1.3.1 Palladium Catalysis – Fundamental Reactions
All transition metal-catalyzed processes can be divided into several
fundamental organometallic reactions.21 Out of these, 7 are depicted in
Figure 1 as they are directly or indirectly involved in further chapters of this
thesis.
Figure 1. Fundamental reactions occurring on a palladium metal center.
1.4 Palladium(II)-Catalyzed Oxidations Not Involving
Carbon-Carbon Bond Formation
In palladium-catalyzed oxidation chemistry, a lot of effort has been
addressed towards developing systems that allow for selective oxidation of
alcohols to ketones using molecular oxygen as terminal oxidant.24,25,26 In this
class of transformations a few of the more successful catalytic protocols
employ Pd/pyridine/O2 (Uemura),27 Pd-NHC/O2 (Sigman),28 Pd/DMSO/O2
(Larock)29 and Pd/bathophenanthroline/O2 (Sheldon).30 As mentioned earlier,
3
palladium(II) readily forms π-complexes with a wide range of unsaturated
hydrocarbons such as alkynes, alkenes and allenes. The coordination to
palladium renders these fragments susceptible towards nucleophilic attack
and/or migratory insertion. Many palladium(II)-catalyzed oxidative
processes, such as the Wacker-oxidation,26,31,32 allylic acetoxylations,33,34,35
aminations,26,36,37 functionalizations of 1,3-dienes38 and allenes39 are based
on derivatization of olefins.
One very recent contribution to the field of palladium(II)-catalyzed
aerobic amination has been reported by Stahl.40 They developed an aerobic
palladium(II)/NHC-catalyzed (L1) aza-Wacker type cyclization of olefinsubstituted sulfonamides (Scheme 2). The process renders hydroindoles (2)
in good to high yield using only molecular oxygen as the terminal oxidant.
Scheme 2. Stahl´s aza-wacker cyclization of olefin substituted sulfonamides.
The properties of any incoming nucleophile play a key role in
determining whether trans- or cis-nucleopalladation is predominant, and it is
generally quite troublesome to prove the actual mechanistic pathway. On
many occasions, both mechanistic processes proceed in parallel and subtle
solvent or ligand alterations can affect the reaction mechanism.37a One
example which beautifully illustrates this phenomenon is how an acetate can
be directed to undergo either trans-attack or cis-migratory insertion in the
1,4-diacetoxylation of 1,3-cyclohexadiene (3) (Scheme 3).38
Scheme 3. Palladium(II)-catalyzed cis/trans 1,4-diacetoxylation of 1,3-dienes.
The selectivity observed arises from the fact that chloride coordinates
stronger to palladium than acetate. Consequently, cis-migratory insertion of
4
(π-allyl)palladium(II)-complex 5b is blocked in the presence of chloride, and
the acetate is forced to attack externally (trans) to give cis-6. On the other
hand, the reactivity of (π-allyl)palladium(II) complex 4b is different (no
chlorides) and it undergoes cis-migratory insertion to give trans-6. In both
cases, p-benzoquinone (BQ) was required in at least catalytic amounts to
obtain an efficient catalytic reaction.
1.4.1 Reoxidation of Palladium(0) Using Different Oxidation Systems
The net-reaction of a palladium(II)-catalyzed oxidation produces
palladium(0). Efficient reoxidation is necessary in order to achieve catalytic
turnover as well as to prevent precipitation of palladium black. Early
palladium-catalyzed protocols employed stoichiometric amounts of
palladium but this is for obvious reasons not ideal. To circumvent this
problem a variety of oxidants have been employed to achieve catalytic
selective oxidation. The most commonly used oxidants in this field include
copper-salts, MnO2, quinones,41 TEMPO,42 various peroxides,43 and
PIDA/PIFA20. Of particular importance to this thesis is the use of BQ and its
analogues, and the suggested BQ promoted pathway for oxidation of
palladium(0) to palladium(II) is depicted in Scheme 4 (M1 to M3).
Scheme 4. Schematic mechanism for the BQ-mediated reoxidation of Pd(0).
Unfortunately, many of the above listed oxidants, except hydrogen peroxide,
contribute to decreased atom efficiency and production of large amounts of
waste. Replacement of the stoichiometric organic or metal oxidant with
molecular oxygen is generally not straightforward because stoichiometric
oxidants can be essential as cofactors and/or ligands in many
transformations.
1.4.2 Aerobic Reoxidation of Palladium(0)
To circumvent the use of stoichiometric amounts of organic or metal-based
oxidants, much research has been devoted to find catalytic systems allowing
for air or molecular oxygen to be used as the terminal oxidant.42 In this field,
Stahl,44 Sigman,45,32 Stoltz46 and Goddard47 have all made significant
contributions with respect to both the mechanistic understanding of aerobic
reoxidation as well as developing new synthetic methods. Most of the
5
progress in aerobic oxidation has been done with Wacker/aza-Wacker type
reactions or in the field of alcohol oxidation. There are of course many
exceptions. For example, Sigman recently reported an efficient aerobic
Pd(IPr)(OTs)2/Cu(OTf)2-catalyzed oxidative Heck-reaction between
electronically non-biased olefins (7) and phenylboronic esters (8).48 The
arylation proceeded smoothly at 35 °C giving styrenes (9) with high E/Zselectivity and chemical yield under an atmosphere of oxygen (Scheme 5).
Scheme 5. Sigman´s palladium(II)-catalyzed oxidative Heck-reaction of non-biased
olefins.
Finding a reoxidation system compatible with both the substrate and the
product is one of the major hurdles in oxidation chemistry. Occasionally,
addition of stabilizing ligands (ligand modulation) or co-catalysts (co-factor
modulation) shuts down, or severely inhibits the intended reaction. One
successful and well documented approach that rely on co-factor modulated
oxidation has been pioneered by Bäckvall. This particular approach was
developed as solution to the problem of reoxidizing palladium(0) to
palladium(II) directly using molecular oxygen.42,49 The authors achieved this
oxidation by mimicking the cell´s respiratory chain i.e. relying on several
coupled redox-catalysts as electron transfer mediators (ETM’s). With this
strategy, the relatively high energy barrier for direct oxidation of
palladium(0) by molecular oxygen was broken down into several steps, all
with relatively low energy barriers (Scheme 6).
Scheme 6. Bäckvall´s triple-catalytic (biomimetic) aerobic oxidation.
This approach is based on introduction of an oxygen activating metal
complex (Scheme, 6, M(ox) + M(red)), which readily gets oxidized by
molecular oxygen. The oxidized metal complex M(ox) then oxidizes 1,46
hydroquinone (HQ) to BQ, and BQ in turn oxidizes palladium(0) to
palladium(II). To this date there are many different metal complexes, serving
as ETM’s, capable of activating molecular oxygen. Some of the most
commonly employed complexes are FePc,50 Cobalt-salophen,51 FePc-AMPS
(10)52 and more recently, a hybrid HQ-tethered-cobalt(II)-catalyst (11).53 In
the latter, BQ has been tethered directly to the oxygen-activating metal
complex, thus fusing two of the components of the triple catalytic system
into one (Figure 2).
Figure 2. Structure of the oxygen activating complexes FePcAMPS (10) and Co(II)hybrid (11).
Implementation of such a “biomimetic” triple-catalytic system has been
successfully achieved in a variety of palladium(II)-catalyzed processes, such
as carbocyclization of en- and dienallenes52,54 1,4-oxidation of dienes55 as
well as ruthenium-catalyzed oxidation of alcohols to ketones and
lactones.56,57
1.5 Palladium(II)-Catalyzed Oxidations Involving
Carbon-Carbon Bond Formation
The formation of new carbon-carbon bonds is absolutely central to organic
chemists. Palladium(0)-catalyzed reactions, relying on oxidative addition of
organo halides, has dominated the field and provided chemists with a
plethora of efficient methods for carbon-carbon bond formation such as the
Suzuki-, Heck-, Negishi- and Tsuji-Trost-reactions.58 On the contrary, there
are rather few examples of oxidative palladium(II)-catalyzed carbon-carbon
bond forming reactions. This concept has proven especially difficult when a
carbon nucleophile, such as an enolate, is used to attack a π-ligand.59 Several
problems arise when using carbon nucleophiles. First of all, even stabilized
nucleophiles are relatively easy to oxidize leading to potential radical
formation and/or side reactions. Secondly, palladium(II) can easily be
7
reduced to palladium(0) by the carbon nucleophile itself. Because of this,
reports on palladium(II)-catalyzed oxidative olefin alkylation are scarce.
However, there are still a few very efficient and well documented reactions
involving carbon-carbon bond formation under oxidative conditions, mainly
in the area of oxidative Heck-type reactions,48,60 allylic C-H alkylation61,62
and various carbocyclizations (See Chapter 1.5.1).63
One of the more recent contributions to this field was reported by White
et al.64 They disclosed an intermolecular oxidative palladium(II)-catalyzed
alkylation of allylbenzene derivatives (12) using stabilized carbon
nucleophiles (13). Formation of the linear alkylated product 14
predominated and the overall process proceeds in good yield (Scheme 7).
Scheme 7. White´s approach to palladium(II)-catalyzed oxidative allylic alkylation.
Furthermore, they determined that the reaction proceeds through C-H
activation of the allyl benzene followed by a subsequent nucleophilic attack
by the nitroacetate on the (π-allyl)palladium(II) intermediate.
In addition to the more classical palladium(II)-catalyzed processes,
palladium(IV) catalyzed carbon-carbon bond forming reactions have been
reported recently.20,65
1.5.1 Palladium(II-IV)-Catalyzed Oxidative Carbocyclizations
Carbocyclization involving palladium(0)-catalysts has received considerable
attention. However, further elaboration on this field is beyond the scope of
this thesis.66 Additionally, several non-oxidative palladium(II)-catalyzed,
approaches (cycloisomerizations) have also been reported by Trost67,
Hitamo68 and others.69
Palladium(II)-catalyzed oxidative carbocyclizations have received
considerably less attention, mainly due to the previously mentioned
incompatibility between oxidants and carbon nucleophiles.70,71 Successful
examples include reports by Bäckvall et al. They reported an intramolecular
alkylation (carbocyclization) of various cyclic 1,3-dienes, substituted with
stabilized carbon-nucleophiles, catalyzed by Pd(OAc)2.72 Furthermore,
Widenhoefer70b,73 and others74 reported palladium(II)-catalyzed oxidative
carbocyclization protocols involving olefin-substituted diketones (Scheme
8). In the latter case CuCl2 was used as co-oxidant in combination with
molecular oxygen, and it was suggested that the enol-form (M4) undergoes
8
trans nucleophilic attack on a (π-olefin)palladium(II)-complex, to form the
new carbon-carbon bond (17).
Scheme 8. Widenhoefer´s palladium(II)-catalyzed oxidative carbocyclization of
diketones.
A related, ligand-modulated approach to aerobic carbocyclization has been
developed by Stoltz for the cyclization of olefin substituted indoles (Scheme
9, 18).75 This elegant reaction uses ethyl nicotinate (L2) as ligand in
combination with Pd(OAc)2 and molecular oxygen to provide tricyclic
heterocycles (19) in high yields under relatively mild conditions.
Mechanistically, this reaction proved much different from Widenhoefers
methodology as it most likely involves C-H-activation of the indole followed
by cis-insertion. Since the initial discovery of this reaction several additional
contributions to this particular field have been made.76
Scheme 9. Stoltz´s palladium(II)-catalyzed oxidative carbocyclization of olefinsubstituted indoles.
Carbocyclizations via palladium(IV) intermediates have emerged as
interesting alternatives to palladium(II)-catalysis.19,20 One such example is
the oxidative carbocyclization of 1,6-enynes (20), in which a hypervalent
iodine compound, PhI(OAc)2 (PIDA), was used as oxidant (Scheme 10).77
Scheme 10. Bellers´s Pd(OAc)2/PIDA-mediated cyclopropanation of 1,6-enynes.
9
The main feature of this approach is the one pot formation of two new
carbon-carbon bonds including one cyclopropyl ring in high yield and
stereoselectivity (21).
Finally, there are protocols that employ nontraditional carbon
nucleophiles, more specifically π-nucleophiles. In these instances, the πsystem acts as a nucleophile instead of a “free” electron pair. The most well
established method utilizing this approach is the reaction of allylsilanesubstituted 1,3-dienes, leading to products formed through an overall transcarbopalladation.78
1.5.2 Palladium(II)-Catalyzed Oxidative Carbocyclizations Involving
Allenes
Over the past 10 years, preparation and subsequent palladium(II)catalyzed carbocyclization of en-52,54,79 and dienallenes80 has been
extensively studied by the Bäckvall-group (Scheme 11).81
Scheme 11. Palladium(II)-catalyzed oxidative carbocyclization of en- and
dienallenes.
The main feature of these protocols has proven to be its high
stereoselectivity and formation of new carbon-carbon bonds under oxidative
conditions.82 Oxidative cyclization of allene-substituted olefin (enallene) 22
has been efficiently carried out with stoichiometric amounts of organic
oxidants (Table 1, entry 4),79 as well as with molecular oxygen via electron
transfer mediators (Table 1, entries 1-3).52,54 Some representative data, with
regards to different reoxidation systems, for the conversion of 22 into 23 is
presented in Table 1 for comparison.
10
Table 1. Comparison of different oxidation systems in the palladium(II)-catalyzed
oxidative carbocyclization of enallene 22 to give 23.
Time
Temp
Yield of 23
(h)
(°C)
(%)
1
75
97
Toluene
4
95
99
5% Pd(O2CCF3)2
Toluene
16
95
89
1% Pd(O2CCF3)2
THF
4
75
94
Entry
Co-catalyst
Pd-cat. (mol%)
Solvent
1
Co-Hybrid (11), O2
1% Pd(OAc)2
EtOH
2
FePc, BQ, O2
5% Pd(O2CCF3)2
3
4
FePc-AMPS (10), BQ, O2
1.5 equiv. of BQ
The mechanism for this carbocyclization has been suggested to proceed
through activation of the allene (M5), possibly through an allenylic C-H
activation, forming an initial (σ-vinyl)palladium(II) intermediate (M6). This
complex undergoes exclusive cis-carbopalladation (M7) followed by βhydride elimination to form 23 (Scheme 12)
Scheme 12. Postulated mechanistic cycle for the palladium(II)-catalyzed oxidative
carbocyclization of enallenes.
An alternative mechanism considered for this transformation goes via an
initiation of the reaction through a syn-C-H abstraction of the allylic side
chain, rendering a (π-allyl)palladium(II) intermediate. However, previous
investigators probed and deemed this pathway less likely through deuterium
labeling studies and by subjecting compound 26 to carbocyclization
conditions (Scheme 13).79
11
Scheme 13. Results disfavoring formation of π-allyl intermediates in the
palladium(II)-catalyzed oxidative carbocyclization of enallenes.
For allene-substituted 1,3-dienes the mechanism shares several key features
with that of enallenes (22). The major difference is that the cyclization step
is followed by the formation of a (π-allyl)palladium(II) intermediate,
susceptible to attack by a nucleophile. These transformations have been
studied in great detail both under stoichiometric and catalytic conditions
using HOAc/acetate as the external nucleophile.80 Other nucleophiles such as
benzoic acid or pentafluorophenol also proved compatible. However,
increased steric bulk on the nucleophile resulted in low yields, exemplified
by the comparison between acetic acid (79%) and isobutyric acid (15%).80c
Overall, palladium(II)-catalyzed oxidative carbocyclization of en- and
dienallenes is an effective way to achieve selective carbon-carbon bond
formation under oxidative conditions and provide potentially useful
synthetic intermediates.
Although this thesis is dedicated to the formation of new carbon-carbon
bonds via palladium(II)-catalyzed oxidative cyclization of allenes, many
other transition metal catalyzed allene cyclizations have been described. In
this class, the majority of transformations belong to the non-oxidative
category.83
12
2. Preparation of Starting Materials
Most of the required starting materials for this thesis have been reported
previously (vide infra). The following section broadly outlines the synthesis
of these materials for additional clarity.
2.1 Representative Materials Prepared for Paper I
Allene-substituted 1,3-dienes (dienallenes) and bromoallenes were prepared
in accordance with literature procedures (Scheme 14).38b,80a Notably, cyclic
dienallenes with unsymmetrical allenic alkyl-groups were obtained as 1:1
mixture of diastereomers.
Scheme 14. Overview of the synthesis of allene-substituted 1,3-dienes (dienallenes).
2.2 Representative Materials Prepared for Paper II
Synthesis of N-tosyl protected allylic amines were conducted according to
published procedures (Scheme 15).84 For the preparation of alkyl substituted
allylic sulfonamides, the isomerization of allylic carbamates was preferred
over alkylation of allylic bromides due to matters involving purification.
13
Scheme 15. Synthesis of allylic sulfonamides.
2.3 Representative Materials Prepared for Paper IV-V
Preparation of enallenes was performed as previously described (Scheme
16).79
Scheme 16. Synthesis of allene-substituted olefins (enallenes).
14
3. Thesis Aim and Objectives
The aim of this thesis has been to extend the scope, generality and
knowledge of our previously developed methodologies on palladium(II)catalyzed oxidative carbocyclization of en- and dienallenes. In addition to
investigating the possibilities and limitations of incoming nucleophiles in the
dienallene carbocyclization, our primary focus was to evaluate a
palladium(II)-catalyzed oxidative carbocyclization of aza-enallenes. To
achieve this, we intended to identify and develop novel synthetic methods to
gain access to suitable nitrogen containing precursors (aza-enallenes) and
study their subsequent palladium(II)-catalyzed oxidative carbocylization.
Furthermore, previous mechanistic data and DFT-calculations conducted
on the palladium(II)-catalyzed carbocyclization reaction have suggested that
interception of a palladium(II) intermediate could be possible. We wanted to
illuminate this possibility and investigate if an expansion of the scope of our
carbocyclization methodology was possible through classical trapping
methodologies.
Finally, as a response to the rather limited substrate tolerance observed in
the preparation of aza-enallenes, we intended to launch a project where
oxazolidinones are efficiently synthesized via palladium catalysis. Ideally,
these substrates would serve as precursors for aza-enallenes.
15
4.Water As Nucleophile in Palladium(II)Catalyzed Oxidative Carbocyclizations (Paper
I)
Previous members of our group have reported that dienallenes undergo
carbocyclization when subjected to catalytic amounts of palladium(II) in
combination with certain oxygen nucleophiles and an oxidant (Scheme 17).80
The highlights of these protocols are: formation of a new carbon-carbon
bond, as well as a carbon-oxygen bond, under oxidative palladium(II)catalysis. The overall reaction across the 1,3-diene can be denoted “1,4carboacetoxylation” and proceeds with high stereoselectivity (>99% trans1,4-addition). For some time, acetate was the only efficient nucleophile,
simply because acetic acid was necessary as solvent. Later developments
allowed for other carboxylic acids as well as some alcohols to be used as
nucleophiles.80c Related oxidative palladium(II)-catalyzed carbocyclizations
of 1,3-diene substituted allylsilanes78 and dienynes85 have also been reported.
Scheme 17. Palladium(II)-catalyzed 1,4-carboacetoxylation of dienallenes.
In order to further broaden the scope and knowledge regarding this
transformation, we decided to investigate whether amides were accepted as
nucleophiles. Also, because biomimetic reoxidations have successfully been
integrated in several related palladium(II)-catalyzed processes42, we argued
that a similar approach could be applicable in this case.
16
4.1 Results and Discussion
4.1.1 Attempts to Use Amide Nucleophiles
As we initially set out trying to find new classes of nucleophiles for the
carbocyclization reaction, a number of amides (Figure 3, 29-33) with
varying pKa values were selected for this study.
Figure 3. Amide nucleophiles evaluated in the palladium(II)-catalyzed oxidative
carbocyclization.
Reactions were conducted using 1 equiv. of 24, 10 mol% of Pd(OAc)2 or 10
mol% Pd(O2CCF3)2, 2 equiv. of BQ and 3 equiv. of the respective amide
(29-33). All reactions were conducted for 24 hours at 45 ºC in three different
solvents: acetone, THF and EtOAc. These specific solvents were selected as
they have previously been highly efficient in related oxidative reactions.50a
Unfortunately, the desired amidated carbocycle was not observed in any of
the reactions described above. Instead, a small amount (~10%) of an
unexpected carbocyclization product was detected. Isolation and
characterization of this species (34) revealed that, instead of attack by the
added amide, nucleophilic attack by 1,4-hydroquinone (HQ) had occurred
(Scheme 18). The presence of the latter side product implies that some form
of background oxidation is occurring forming small amounts of HQ.
Addition of a weak external base (1-5 equiv. LiOAc or NaOAc) or addition
of free amines (e.g. Et3N) did not promote amide attack.
Scheme 18. Detected product from attempted use of amide nucleophiles in the
oxidative carbocyclization of dienallene 24.
One additional product was observed when the reaction was performed in
acetone (not anhydrous), and after some experimentation it was determined
to be the hydroxylated carbocycle 35 (Scheme 19). Excited by this result
excess water was added to a reaction in THF, and the same product from
17
water attack could be observed in improved yield. Although our initial goal
of using amides as nucleophiles was unsuccessful we believed that the water
attack was worth looking into in more detail.
4.1.2 Water as Nucleophile
Treatment of 24 with 10 mol% of Pd(O2CCF3)2, 2 equiv. of BQ in a
water/THF mixture (1:10) for 20 hours at room temperature produced
hydroxylated product 35 in 66% yield. Analysis of product 35, through
derivatization with Ac2O and comparison with the NMR-data of the known
acetate,80a revealed that the product had been formed via an overall trans1,4-carbohydroxylation (>99% trans), with formation of a cis-fused ringjunction. Encouraged by these results the H2O/THF ratio together with the
palladium(II)-loading was studied. We disclosed that 1 mol% of
Pd(O2CCF3)2 together with 2 equiv. of BQ in a H2O/THF mixture (4:1)
rendered the desired product 35 in 85% isolated yield after a 9 hour reaction
time (Scheme 19).
Scheme 19.
dienallenes.
Palladium(II)-catalyzed
oxidative
1,4-carbohydroxylation
of
Comparable results were obtained with Pd(OAc)2, but introduction of halides
to the palladium catalyst (Li2PdCl4 or PdCl2(CH3CN)2) completely inhibited
this transformation. Attempts were also made to conduct this reaction in a
purely aqueous medium; unfortunately this only provided 16% of cyclized
product, even after use of more harsh conditions.
4.1.3 Reaction Scope
To investigate the substrate scope of this carbohydroxylation, a series of
dienallenes were synthesized as previously described.80 Gratifyingly, all
investigated dienallenes reacted smoothly with 1-5 mol% Pd(O2CCF3)2 and 2
equiv. of BQ to give carbocycles 35-41 in 74-92% isolated yield (Table 2,
Method A). With 1 mol% Pd(O2CCF3)2 the desired carbohydroxylated
products were obtained within 12-24 hours. With 5 mol% of catalyst, higher
yields were achieved with much shorter reaction times (5 hours), hence a
more practical procedure.
18
Substrates with non-symmetrical allenic side chains (Table 2, entries 3-5,
37-39) provided hydroxylated carbocycles with the more substituted double
bond as the major isomer (thermodynamic product).
Table 2. Scope of the 1,4-carbohydroxylation of dienallenes (E = CO2Me).
Entry
Product
Yield (%)
Yield (%)
Method A (a,c)
Method B (b,c)
85
88
E E
HO
1
90(d)
35
2
3
4
5
6
7
83
89
81
(d)
80
82(e)
37/37’
37/37’
2:1
2:1
84(d)
82(e)
38/38’
38/38’
8:1
7:1
85(d)
84(e)
39/39’
39/39’
7:1
6:1
90
91
92(d)
94(e)
74(d)
71(e)
(a) Method A (Stoichiometric BQ): Dienallene (1.0 equiv.), Pd(O2CCF3)2 (1 mol%) and BQ (2 equiv.),
H2O/THF (4:1), r. t., 9-12 h (b) Method B (Aerobic): Dienallene (1.0 equiv.), Pd(O2CCF3)2 (1 mol%),
BQ (5 mol%), and FePc (1 mol%) under O2 atmosphere (balloon), H2O/THF (4:1), r. t. 24-48 h. (c)
Isolated yields. (d) 5 mol% of Pd(O2CCF3)2, 4-6 h. (e) With 5 mol% of Pd(O2CCF3)2, 20 mol% of BQ,
and 5 mol% of FePc in H2O/THF (2:1) at r.t. for 24 h.
The dienallene bearing an 8-membered cyclic olefin proved least
efficient, producing the desired hydroxylated carbocycle 41 in 74% isolated
yield (Table 2, entry 7). For this specific carbocycle, as well as for 35, the
configuration of the ring-junction was determined to be cis. On the contrary,
carbocyclization of the 7-membered dienallene gave the trans-fused ring
19
product 40 but still through trans-1,4-carbohydroxylation, all in accordance
with previous studies (Table 2, entry 6).80
The results discussed so far have involved the use of stoichiometric
amounts of BQ as the terminal oxidant. To make this process catalytic in
BQ, a triple catalytic aerobic oxidation system based on
iron(II)phthalocyanine (FePc, 42) and catalytic amounts of BQ was
introduced (Scheme 20).
Scheme 20. Triple catalytic
carbohydroxylation of dienallenes.
(biomimetic)
palladium(II)-catalyzed
1,4-
Subjecting 24 to 1 mol% of Pd(O2CCF3)2, 5 mol% BQ and 1 mol% FePc
under an atmospheric pressure of molecular oxygen (balloon), resulted in an
effective oxidation protocol (Table 2, entry 1, Method B). At room
temperature carbocycle 35 was isolated in 88% yield after 36 hours. A few
cyclizations were slower; carbocyles 37 and 37’ (Table 2, entry 3) required
48 hours to fully consume the starting dienallene. To reduce the reaction
time, the catalytic protocol was modified to 5 mol% of Pd(O2CCF3)2, 20
mol% of BQ and 5 mol% of FePc in a 2:1 H2O/THF-mixture. This
adjustment allowed for full conversion within 24 hours. With these data in
hand we applied this catalytic system to the remaining substrates (Table 2,
Method B). Comparison between the triple catalytic system and the
stoichiometric BQ-mediated version showed that they were comparable with
respect to chemical yield.
4.1.4 Mechanistic Discussion
Mechanistic studies have previously been performed on oxidative 1,4carboacetoxylation of dienallenes,80a and several aspects of the previous
studies has been taken into consideration in our suggested mechanism for the
1,4-carbohydroxylation. To begin with, this process requires overall trans1,4-carbohydroxylation to occur across the 1,3-diene. We suggest that the
20
reaction is initiated by activation of the pendant allene (M12) to form (σvinyl)palladium(II)-complex M13 by loss of CF3CO2H (TFA) or HOAc,
depending on the palladium source. This initial process has been in part
supported by monitoring the reaction by 1H NMR spectroscopy. When
conducting this reaction with stoichiometric amounts of Pd(OAc)2 in THFd8, without any BQ, we only observed rapid disappearance of the allenic
methyl groups. No cyclized product was observed without the addition of at
least catalytic amounts of BQ. Based on this we believe that the reaction
probably starts with something resembling an allenylic C-H activation
(Scheme 21, M9-M11).
Scheme 21. Possible pathway for the “allene attack on palladium” in the oxidative
carbocyclization of en- and dienallenes.
After “allene attack on palladium”, cis-migratory insertion of the 1,3-diene
into the vinyl palladium bond in M13 would form the new carbon-carbon
bond together with the (π-allyl)palladium(II)-complex M14 (Scheme 22).
The final step raises the question why water can act as a nucleophile in this
specific case, as this observation is relatively unique in palladium-diene
chemistry (M14-M15).86 Our explanation is based on that the adjacent olefin
on the prenyl side chain coordinates to the palladium(II)-center. This will
increase the bond length of the palladium-carbon bond furthest away from
the prenyl side chain in M14. The effect of this “chelate” increases the
electrophilicity of M14, thus making it more susceptible towards
nucleophilic attack (by water). This proposal has been supported by a recent
DFT-calculation (Scheme 22, M14DFT).87 Finally, decoordination of M15
followed by BQ-mediated reoxidation closes the catalytic cycle (Scheme
22).
Another mechanism a priori possible and can explain the reaction
outcome of the catalytic reaction. In this mechanism, (η4-π-1,3diene)palladium(II) coordination from the same face as the already existing
substituent, followed by a trans-Wacker addition of water, could give rise to
a hydroxylated (π-allyl)palladium(II) intermediate (not shown). Insertion of
the pending allene, and subsequent β-hydride elimination would produce the
carbocycle with the correct stereochemistry. Although this is a possible
pathway it was considered less likely based on the results obtained from the
stoichiometric Pd(OAc)2 reaction discussed earlier.
21
Scheme 22. Proposed mechanism
carbohydroxylation of dienallenes.
for
the
palladium(II)-catalyzed
1,4-
4.1.5 The Hydrolysis Dilemma
In principle, the carbocyclization reaction could proceed through a simple
trifluoroacetoxylation/hydrolysis pathway, only having the formal
appearance of a hydroxylation. To evaluate this possible pathway
Pd(O2CCF3)2 was replaced with Pd(OAc)2. With the latter catalyst, the
expected 1,4-carbohydroxylated product 35 was obtained in 80% isolated
yield from 24, without any further modification to the catalytic protocol.
Next, a mixture of dienallene 24 and acetoxylated carbocycle 25 was
subjected to the carbocyclization protocol using Pd(OAc)2 as catalyst. In this
experiment we observed that 25 remained intact after the reaction and that
the desired hydroxylated carbocycle formed as expected in 76% isolated
yield (Scheme 23).
Scheme 23. Probing the possibility of an acetoxylation/hydrolysis pathway in the
palladium(II)-catalyzed oxidative carbocyclization of dienallene 24.
22
4.2 Conclusions
Allene-substituted 1,3-dienes (dienallenes) have been shown to undergo a
palladium(II)-catalyzed
stereoselective
carbocyclization (1,4-carbohydroxylation) at room temperature upon treatment with 1-5 mol% of
Pd(O2CCF3)2 and 2 equiv. of BQ in H2O/THF solvent mixtures. This
catalytic system allows for the selective formation of new carbon-carbon
bonds under oxidative conditions in an aqueous medium. The transformation
relies on attack by water on a (π-allyl)palladium(II)-complex to produce
trans-1,4-carbohydroxylated carbocycles as the only observed isomer in high
yield. The present methodology was further extended by the implementation
of a FePc-based biomimetic reoxidation system, using molecular oxygen as
the terminal oxidant. Comparison between the two reoxidation methods
showed that they were equally efficient with respect to yield.
A mechanism for this transformation was also suggested. It is likely that
the external attack by water is made possible due to formation of a (πallyl)palladium(II)-olefin complex (chelate). This “chelate” would make the
carbon atom furthest away from the palladium(II)-center more electrophilic
and susceptible to nucleophilic attack.
23
5. Copper(I)-Catalyzed N-Allenylation
Allylic Sulfonamides (Paper II)
of
As we are interested in investigating the possibility of a palladium(II)catalyzed oxidative carbocyclization of aza-enallenes we sought a method
for the preparation of this substrate class. We envisioned that a transition
metal catalyzed coupling reaction could potentially be the key step in their
preparation.
Transition metal-catalyzed cross-couplings have emerged as one of the
most diverse procedures for selective construction of new carbon-carbon
bonds. Notably, these protocols are not limited to formation of carboncarbon bonds. Vast numbers of selective carbon heteroatom (N, O, S) bond
formation, catalyzed by transition metals have been disclosed over the
years.88 One very effective, and currently dominating reaction for carbonnitrogen bond formation, was originally developed by Buchwald and
Hartwig.89 A common feature of these protocols is that they employ
palladium(0) as the catalyst, in combination with phosphine- and, more
recently, NHC-ligands.90 This particular catalytic approach has been studied
in great detail, and even coupling reactions between amines/amides and
deactivated aryl chlorides have recently been realized through exhaustive
ligand modification.91 Apart from palladium(0)-catalyzed couplings, the use
of copper(I) has been reported as a cheap and almost as efficient
alternative.92
Copper(I)-based systems have their origin in the carbon-carbon bond
forming Ullmann-protocol, reported by Fritz Ullmann in 1901.93 Originally,
the coupling was used to form of new aromatic carbon-carbon bonds, but
later developments have made it equally useful in carbon-heteroatom bond
formation.92 The advantage of the Ullmann protocol is mainly that the price
of copper is much lower than that of palladium, hence, low catalytic loadings
are less crucial. In general, copper(I)-based protocols require harsher
conditions to facilitate coupling of amines and amides with organohalides.
24
Scheme 24. Examples of carbon-nitrogen bond formation using the Ullmann- and
Buchwald-Harwig-reactions.
Two representative carbon-nitrogen bond-forming reactions based on the
Buchwald-Hartwig94- and Ullmann-protocols95 are given in Scheme 24.
The Ullmann-type reaction has proven effective in a range of carbonnitrogen bond forming reactions. Despite this, we were only able to find two
representative protocols demonstrating coupling of haloallenes with nitrogen
nucleophiles. These two protocols, developed by Trost96 and Hsung97,
respectively, both described the copper(I)-catalyzed coupling of carbamates
(43) and ureas with bromoallenes and iodoallenes (44a and 44b) to give Nallenylated products (45) in high yields (Scheme 25).
A few other interesting alternatives to prepare N-allenylated compounds
have been reported previously. Most of these reactions involve basecatalyzed isomerization of propargylic amines and amides, opening up
access to monosubstituted N-allenyl amides/carbamates with ease.98
Unfortunately, synthesis of disubstituted N-allenyl amides has proven very
difficult using this approach.99
Based on the results reported by Trost, we set out to investigate if a
copper(I)-catalyzed coupling between allylic sulfonamides and haloallenes
was a viable method to produce aza-enallenes.
Scheme 25. Trost´s copper(I)-catalyzed N-allenylation of oxazolidinones.
25
5.1 Results and Discussion
5.1.1 Identifying a Model Substrate and Initial Optimization
The choice of model substrate finally fell on cyclic tosyl-protected allylic
amide 47, and initial couplings were attempted using iodoallene 44b.
Coupling using 1 equiv. of 47 with 2.5 equiv. of iodoallene 44b (or
bromoallene 44a), together with 10 mol% of CuI, 20 mol% of N,N dimethylethylenediamine (DMEDA, Figure 4, L4), and 2 equiv. of Cs2CO3
in refluxing toluene did not give rise to any detectable amount of the coupled
product 48. After several attempts, varying copper sources (Cu2O, CuI,
CuCN, CuBr, CuTC, CuOAc), bases, solvents, temperature and various
ligands (mainly phosphines and amines), it was decided to abandon 47 as
model substrate (Scheme 26). Instead we focused on reevaluating the
reaction using a less sterically demanding allylic sulfonamide.
Scheme 26. Unsuccessful Cu(I)-catalyzed coupling of 47 with haloallenes.
Our attention turned towards the simplest tosyl-protected allylic
sulfonamide 49 (Scheme 27). Treatment of 1.0 equiv. 49 with 2.5 equiv. of
bromallene (44a), 10 mol% of CuI, 20 mol% of DMEDA (Figure 4, L4) and
2.0 equiv. of Cs2CO3 in toluene (90 °C) afforded the N-allenylated product
50 in a modest isolated yield of 49% after 16 hours. Encouraged by these
results, several copper sources were evaluated again. Copper(I)-thiophene-2carboxylate (CuTC) distinguished itself as the superior catalyst (64%
isolated yield) followed by CuI, resulting in approximately 15% lower
conversion. Throughout this study, DMEDA (L4) was used as ligand. A
Comparison between a few other nitrogen-based ligands revealed that only
structurally related DEEDA (L5) provided coupled product in moderate
yield under the conditions described above in Figure 4.
26
Figure 4. Ligands evaluated for the copper(I)-catalyzed coupling of allylic
sulfonamide 49 with bromoallene 44a.
The influence of various bases in addition to Cs2CO3 was also briefly
investigated. Powdered K3PO4 did in fact give the desired product in 38%
isolated yield but K2CO3 and Na2CO3 were completely ineffective. Other
bases such as; KOtBu or amines (DBU, DBN and Et3N) failed to improve
the reaction
An interesting phenomenon caught our attention when the temperature
was increased. In refluxing toluene N-allenylation did proceed, but was
accompanied by large amounts of a side product derived from an
intramolecular Alder-ene reaction of the product (50AE).100 Already at 90
°C, small amounts of this product were detected. Hence, it was decided to
conduct the coupling at 80 °C to avoid this problem (Scheme 27).
Scheme 27. Optimized conditions for the Cu(I)-catalyzed N-allenylation of allylic
sulfonamides.
At an early stage we noticed that the aquired aza-enallenes were very
sensitive towards acidic conditions. Purification over silica gel proved
impossible, even after deactivation with Et3N. Isolation was instead
performed using basic aluminum oxide (Al2O3, basic, Brockmann I). After
overcoming these initial problems we set out to investigate the limitations of
the catalytic reaction.
5.1.2 Allenylation of Allylic Sulfonamides
Coupling of various alkyl-substituted allylic sulfonamides is presented in
Table 3. Monosubstituted olefins reacted smoothly with bromoallene 44a,
affording aza-enallene 50 in high yield (Table 3, entry 1). An additional (E)methyl substituent at the allylic moiety did not affect the reaction to any
considerable extent (Table 3, entry 2). When the size of this substituent was
27
further increased from methyl to butyl, a slight drop in isolated yield (Table
3, entry 7) to 72% was observed. N-Allenylation of disubstituted allylic
sulfonamides proceeded with comparable yields regardless of the
substituents on the bromoallene (Table 3, entries 3, 5 and 6).
Table 3. Scope of the copper(I)-catalyzed N-allenylation of allylic sulfonamides.(a)
Entry
Sulfonamide
Aza-enallene
Yield (%)a, b
1
89
2
82
3
74
4
35
5
65
6
67
7
72
8
53c
9
20c
10
52c
(a) Reaction conditions: Allylic sulfonamide (1.0 equiv.), 15 mol% CuTC (46), 30 mol% DMEDA (L4),
Cs2CO3 (2.0 equiv.), bromoallene (2.5 equiv.), toluene, 80 °C, 24 h. (b) Isolated yield. (c) Conducted at 40
°C.
Surprisingly, N-allenylation of 53 using a monosubstituted bromoallene
displayed a dramatic decrease in yield (Table 3, entry 4). We speculate that
the initially formed aza-enallene undergo a base-catalyzed isomerization to
form the corresponding propargylic isomer (not shown).
Somewhat surprising results were also obtained with (E)-aryl substituted
allylic substrates 60, 62 and 64 (Table 3, entries 8-10). In all these instances,
the starting material was completely consumed, but decomposed into an
28
unknown mixture of products in toluene at 80 °C. Lowering the temperature
to 40 °C generated the desired aza-enallenes, 61 and 65 in approximately
50% yield, accompanied by around 25% of a byproduct originating from an
intramolecular [2+2]-cycloaddition.100,101 Unfortunately, N-allenylation of 62
did not yield more than 20% of the desired N-allenylated product (63) even
after several repeated attempts (Table 3, entry 9).
5.1.3 Structural Limitations
A logical extension of this coupling protocol would be to simply exchange
the amine protecting group from p-toluenesulfonyl (Ts) to for instance
acetyl. For this purpose, a few analogues of 49, where the tosyl-group had
been replaced with Cbz, Boc, benzyl, acetyl and trifluoroacetyl were
prepared. Unfortunately all of the amides, except trifluoroacetamide, failed
to give the desired product. In the case of trifluoroacetamide, around 20%
conversion was observed and the product proved equally unstable on silica
gel as the corresponding tosyl-protected derivatives. A likely explanation for
the failed coupling of these substrates is that a sufficiently acidic amide N-H
is crucial for reactivity. In our series of amides, the tosyl and
trifluoroacetylamides have the lowest pKa.
A few other potentially useful allylic sulfonamides were prepared and
evaluated in the coupling with bromoallenes. Introduction of a methyl group
on the internal carbon of the allylic bond inhibited the reaction (Figure 5,
66). Replacement of the allylic side chain with a simple ethyl group (Figure
5, 67) resulted in only 10% conversion, and the same observation was made
for homoallylic sulfonamides (not shown). Furthermore, the catalytic system
does not accept substrates where a substituent has been introduced in the αposition with respect to the amide nitrogen (Figure 5, 68-70)
Figure 5. Allylic sulfonamides incompatible with the developed copper(I)-catalyzed
N-allenylation.
5.1.4 Allenylation of Non-Allylic Sulfonamides
Although the developed catalytic reaction had some limitations, some
additional compound classes proved applicable in the N-allenylation reaction
without further optimization (Table 4). For example, Ts-protected anilides
(71 and 73) could be allenylated in moderate yields, even with a substituent
29
in the ortho-position (Table 4, entry 2). Ditosylated 1,2-anilides formed an
interesting cyclized product when subjected to the coupling protocol (Table
4, entry 3).
Table 4. Copper(I)-catalyzed cross-coupling between sulfonyl amides and
bromoallene 44a.(a)
Entry
Substrate
Yield (%)a, b
Product
•
1
NTs
56
72
•
2
NTs
74
44
O
3
63c
4
41
(a) Reaction conditions: Sulfonamide (1.0 equiv.), 15 mol% CuTC (46), 30 mol% DMEDA (L4), Cs2CO3
(2.0 equiv.), 44a (2.5 equiv.), dioxane (reflux), 24 h. (b) Isolated yields. (c) Refluxing THF.
Presumably, after coupling, a subsequent 5-exo-trig-cyclization step
(catalyzed or non-catalyzed) occurs, rendering the cyclic protected “aminal”
76 in moderate yield. Finally, tosyl-protected amino acid 77 also underwent
N-allenylation in a modest isolated yield of 41%.
5.1.5 Stability of the Aza-Enallenes
The majority of the prepared aza-enallenes had to be handled with great care.
Even when stored under argon at -20 °C their lifetime was generally limited
to weeks. After approximately 2 weeks, the isolated products could contain
up to 10% of various byproducts, of which 1,3-dienes were the major ones.
Stirring the aza-enallene in THF together with 20 mol% of AcOH did not
induce immediate decomposition. However, the same mixture underwent
partial hydrolysis of the N-allenyl bond when heated (50 °C), to give N-allyl4-methylbenzenesulfonamide (83) and 3-methyl-2-butenal (84) (Scheme 31,
Chapter 6.1.1).
30
5.1.6 Mechanistic Discussion
The actual mechanism in operation in these Ullmann-type couplings is still
subject to debate and multiple catalytic pathways have been put forward.102
In this study, no direct investigation of the mechanism was conducted.
Instead, experimental data have been the basis of the schematic mechanism
suggested in Scheme 28. Our proposal is based on a copper(I)-copper(III)
cycle. The cycle is initiated by an overall oxidative addition of a
DMEDA/CuTC-complex to the bromoallene, forming the d8-square-planar
(σ-allenyl)copper(III) complex M16 (Scheme 28). It is not clear whether the
sulfonamide is replacing the thiophene-2-carboxylate(TC)-ligand on copper
prior to, or after oxidative addition (Scheme 28, central pathway). Next, we
suggest that the pending olefin can act as a ligand giving rise to complex
M17-M18 (in equilibrium). Reductive elimination of M18 would then give
the aza-enallene as well as regenerating the active catalyst. This hypothesis
is based on the fact that removal of the allylic olefin, or addition of a
substituent in the α-position, resulted in an inactive catalytic reaction. Also,
bulkier substituents on the olefin provided lower yields of coupling product.
Ts
N
Br
NH
Cu (I) TC
HN
•
•
35
decoordination
Cs
NTs
-CsTC
oxidative addition
+
Ts
N
TC
HN
NH
Cu(I) NTs
HN
•
Cu (I)
HN
NH
Cu(III)
HN
Br
M19
M16
oxidative
addition
Br
reductive elimination
Ts
N
(III)
N
H
•
Cu
N
H
M18
•
•
ligand exchange
Ts
N
NH
(III)
Cu
HN
•
M17
-CsBr
Cs
NTs
Scheme 28. Possible pathway for the copper(I)-catalyzed formation of azaenallenes.
31
5.2 Conclusions
As the first stage towards the carbocyclization of aza-enallenes we have
developed a copper(I)-catalyzed N-allenylation of allylic sulfonamides. Our
coupling reaction between allylic sulfonamides and bromoallenes renders
aza-enallenes in moderate yields using 15 mol% of CuTC as catalyst and 30
mol% of DMEDA as the privileged ligand. The substrate scope is currently
limited to allylic tosyl-protected sulfonamides together with a few nonallylic tosyl-amides/anilides.
This study has provided us with valuable information on the stability,
purification, and storage of this new substrate category. This information
will be truly helpful when studying future catalytic reactions. We were
intrigued to see that the developed protocol could supply us with substantial
quantities of aza-enallenes, suitable for palladium(II)-catalyzed oxidative
carbocyclizations.
32
6.
Palladium(II)-Catalyzed
Oxidative
Carbocyclization of Aza-Enallenes (Paper III)
Many biologically interesting molecules have at least one heterocyclic ring
in its basic backbone. Consequently, synthetic methods that provide access
to structurally diverse nitrogen containing heterocycles have always been
greatly appreciated by the synthetic community.103 A few representative
examples of naturally occurring compounds with N-heterocylic cores are
presented in Figure 6 (79-81).104
Figure 6. Examples of naturally occurring substances having a N-heterocyclic
backbone.
Our research group has had a long lasting interest in palladium-catalyzed
cyclizations of enallenes105 especially oxidative variants (Scheme 29).54,79
This category of catalytic transformations has facilitated selective formation
of carbon-carbon bonds, but the substrate scope has so far been limited to
fully carbocyclic structures.
Scheme 29. Features and limitations of the palladium(II)-catalyzed oxidative
carbocyclization of enallenes.
Introduction of a heteroatom into the core of the enallene would significantly
increase the substrate scope. The first step to achieve this goal was presented
in chapter 5, outlining a copper(I)-catalyzed cross-coupling protocol between
allylic sulfonamides and bromoallenes that open up access to aza-enallenes
33
in moderate yields. With these substrates in hand, we aimed to investigate
whether an oxidative carbocyclization was a viable method for also the
preparation of pyrrole/pyrrolidine-type N-heterocycles (Scheme 30).
Evaluation of this novel reaction is not just interesting from an academic
point of view as the formed N-heterocycles are potential precursors to
natural products and pharmaceuticals.104
Scheme 30 Outline towards N-heterocycles using palladium(II)-catalyzed oxidative
carbocyclization of aza-enallenes.
6.1. Results and Discussion
6.1.1 Optimization of the Reaction Conditions
When aza-enallene 52 was allowed to react with 5 mol% of Pd(OAc)2, 1.05
equiv. of BQ in THF at 50 °C for 5 hours, the desired product 82 was formed
in approximately 80% yield (Table 5, entry 1).
Table 5. Investigation of different palladium(II)-sources in the oxidative
carbocyclization of aza-enallene 52.(a)
Entry
Pd(II)-source
Conv.
(%)(b)
Prod (82:83:84:85)
Ratio (%)(b)
Yield of 82
(%)(c)
1
Pd(OAc)2
100
90:0:0:10
80
2
PdCl2(CH3CN)2
100
5:45:45:5
<5
3
Pd(O2CCF3)2
95
1:49:49:0
<1
4
Pd(acac)2
0
s.m
0
5
Pd(OAc)2 + Phen(d)
0
s.m
0
6
none
0
s.m
0
(a) Reaction conditions: Aza-Enallene (1.0 equiv.), Pd(II)-source (5 mol%), BQ (1.05 equiv.), THF, 50
°C, 5 h. (b) Product distribution and conversion determined using 1H NMR analysis. (c) Yield based on
internal standard (anisole). (d) Phen = 1,10-phenantroline (L9)
34
In addition to the expected product (82) we also detected around 10% of a
by-product (85), arising from a subsequent Diels-Alder reaction between the
cyclized product 82 and BQ.
Switching palladium(II)-source from Pd(OAc)2 to PdCl2(CH3CN)2 had a
profound negative effect on the outcome of the reaction. Instead of selective
cyclization, only trace amounts of the product, accompanied by an
approximate 1:1 mixture of N-allyl-4-methylbenzenesulfonamide 83 and 3methyl-2-butenal 84 (Table 5, entry 2) was observed. We believe that this
was a result of an acid-catalyzed hydrolysis of the N-allenyl bond (Scheme
31, M20-M22). The same characteristics was observed during the
preparation of aza-enallenes.
Scheme 31. Acid-catalyzed hydrolysis of aza-enallenes (P. S = proton shift).
Pd(acac)2 or Pd(OAc)2 + 1,10-phenantroline (1:1) did not catalyze the
reaction and the majority of the starting material was recovered. Background
reactions (without any palladium(II)-catalyst) rendered no reaction a part
from some slight background decomposition of the starting aza-enallene.
It has previously been demonstrated that oxidative carbocyclizations are
most effective in THF or ethanol.54,79 When catalysis was performed in
ethanol, the reaction proceeded smoothly, albeit with formation of an
unidentified byproduct (~5%), which proved impossible to separate from the
carbocycle using chromatographic methods. Owing to this, the best solvent
was finally decided to be reagent grade THF (~100 ppm H2O). Some further
experimentation using THF as solvent revealed that the transformation was
not sensitive towards moisture or air, and that a few different quinones
(napthoquinone, 2,6-dimethylbenzoquinone (DMBQ)) were suitable as
oxidants.
6.1.2 In situ 1H NMR Experiments
Formation of the Diels-Alder adduct has an eroding effect on the overall
yield, consuming both the aza-enallene and the oxidant. To obtain further
insight into the formation of 82 and 85, monitoring of the reaction using 1H
NMR spectroscopy in THF-d8 (48 °C) using 2 mol% of Pd(OAc)2, 1.05
equiv. of BQ and 1.00 equiv. of anisole (internal standard) was undertaken
(Figure 7). From this experiment it became apparent that the reaction was
complete within approximately 2.5 hours, furnishing 82 in 82% yield based
on the internal standard.
35
Conversion (%)
100
90
80
70
60
50
40
30
20
10
0
0
50
100
150
200
Time (min)
♦= Starting material (52), ■= Product (82)▲= Diels-Alder adduct (85)
Figure 7. NMR study on the formation of 82 and Diels-Alder adduct 85.
The Diels-Alder-adduct (85) started to become measurable at around 40%
conversion, and the total amount was around 10% after full conversion.
Some further interesting observations were made when studying the data
from Figure 7. Formation of 82 displayed a sigmoidal curve. The probable
explanation to this phenomenon is that commercially available Pd(OAc)2
(being trimeric) requires some form of activation or dissociation to produce a
catalytic active monomeric species.106 Another plausible explanation could
be that the unwanted Diels-Alder adduct (85) participates as a ligand in the
reaction. A similar observation was recently described in the stereoselective
1,4-diacetoxylation of 1,3-dienes.107
6.1.3 Scope and Limitations
The scope and limitations of this transformation were evaluated with 2 mol%
of Pd(OAc)2 and 1.05 equiv. of BQ in THF at 50 °C for 4 hours. The results
are summarized in Table 6. In general, aza-enallenes were oxidatively
cyclized to give N-heterocycles in good to high yields. Methyl substituted
substrate 52 afforded pyrroline 82 in 74% isolated yield after 4 hours.
Aza-enallene 50 (Table 6, entry 2) somewhat surprisingly afforded the Ntosyl protected pyrrole 86 in 64% yield after extending the reaction time to
24 hours. This aromatic product was probably formed through syn-β-hydride
elimination (after bond rotation) to form the N-heterocycle with the
exocyclic double (Table 6, entry 2, 86b). Studying this reaction using 1HNMR spectroscopy revealed that this was indeed the case. 86b formed first
but was almost completely isomerized during the course of the catalytic
reaction to the corresponding aromatic derivative 86. (Scheme 36, M30).
36
An interesting observation was made when a phenyl group was introduced
on the allylic side chain (Table 6, entry 3). Aza-enallene 61 was smoothly
cyclized to give (Z)-87 in 95% yield without any detectable amount of
[2+2]-cycloaddition, aromatized, or Diels-Alder by-products.
Table 6. Substrate scope of the palladium(II)-catalyzed carbocyclization of azaenallenes.(a)
Entry
Aza-enallene
Product
Yield (%)(b)
1
74
2
64(c)
Ts
N
3
95 (Z)
•
61
4
71
5
84
Ts
N
6
57
70
90/90’
1:3
•
Ts
N
7
8
91
No reaction.
81
0d
(a) Reaction conditions: Aza-Enallene (1.0 equiv.), Pd(OAc)2 (2 mol%), BQ (1.05 equiv.), THF, 50 °C, 4
h. (b) Isolated yields. (c) 24 h reaction time. (d) S.m. was recovered.
The absence of the latter two products was explained by the extended
conjugation added by the phenyl group, making the structure less prone to
undergo Diels-Alder reactions and/or isomerization.
37
Disubstituted olefins 55-57 provided good yields of carbocyclization
products regardless of the substituent on the allenic side chain (Table 6,
entries 4-7). As anticipated, aza-enallenes with an internal substituent on the
double bond failed to give any carbocyclized product, even at elevated
temperatures (Table 6, entry 8). This reactivity can be explained by the
absence of β-hydrogens in the presumed palladium(II) intermediate (Scheme
36, M29).
The Diels-Alder adduct could be formed selectively by subjecting 52 to
slightly modified reaction conditions (2.1 equiv. BQ), This modification
resulted in almost quantitative formation of 85 with excellent selectivity
(Scheme 32, 99% endo, trans/cis 97:3).
Scheme 32. Selective formation of tricyclic Diels-Alder adduct 85.
6.1.4 Aerobic Reoxidation and Tandem Diels-Alder Reaction
Inspired by the success of the stoichiometric BQ-mediated oxidation
procedure, we decided to further broaden the synthetic utility of the
transformation by investigating a biomimetic triple catalytic oxidation
system. Our choice fell on a catalytic system previously developed in our
laboratories.53 This specific system utilizes an oxygen-activating catalyst
tethered to BQ (11), thus making it possible to perform the biomimetic
oxidation without addition of catalytic amounts of BQ (Scheme 33, M23).
Removal of the BQ will have one other positive effect. Perhaps it will be
possible to perform the particular carbocyclization in the presence of other
dienophiles in a tandem carbocyclization/Diels-Alder sequence. In addition
to a more practical and operationally simple procedure, tethering of BQ to
the metal macrocycle has also been shown to accelerate the reoxidation
system.54
38
Scheme 33. Biomimetic oxidation of palladium(0) using Co-hybrid-catalyst 11.
With 11 as the only co-catalyst, a higher yield of cyclized product was
expected because secondary Diels-Alder reactions between BQ and 82 are
minimized. After some initial experimentation employing different oxygenactivating catalysts and solvents we found good conditions for the
biomimetic reoxidation. Subjecting 52 or 56 to 5 mol% of Pd(OAc)2, 5
mol% of 11 in THF at 50 °C under 1 atm of O2 (balloon) provided the
desired dihydropyrroles, 82 and 91, in 94% and 86% isolated yield (Scheme
34). As postulated, formation of Diels-Alder adducts was not detected in the
crude mixtures.
Ts
N
•
52
Ts
N
5 mol% Pd(OAc)2
5 mol% Co-cat (11)
Ts
N
•
THF, 50 °C
O2, 6 h
94% isol. yield
5 mol% Pd(OAc)2
5 mol% Co-cat (11)
82
56
Ts
N
THF, 50 °C
O2, 6 h
86% isol. yield
91
Scheme 34. Biomimetic palladium(II)-catalyzed oxidative carbocyclization of azaenallenes 52 and 56.
We considered it quite surprising that the reactions proceeded without
substantial decomposition of the starting aza-enallene or formation of overoxidized side-products. The absences of large amount of side- and overoxidized products once again confirm that the biomimetic reoxidation
approach is a mild and very selective reoxidation system.
Apart from providing a more environmentally friendly reoxidation,
biomimetic reoxidation systems using Co-hybrid catalyst 11 opened up the
possibility to use a variety of dienophiles in a tandem
39
carbocyclization/Diels-Alder reaction (Scheme 35). For example, the
oxidative cyclization of 52, conducted with the addition of 1 equiv. of
maleimide, resulted in the formation of tricyclic N-heterocycle 92 in 93%
isolated yield. In this case, both the endo- and cis/trans-selectivity was lower
than the analogous reaction with BQ. Less activated dienophiles, such as
acrylates, displayed disappointing results and by-products started to form
before acceptable conversions could be obtained. So far, the
carbocyclization/Diels-Alder sequence is limited to activated dienophiles
such as maleic acid and maleimide.
Scheme 35. Tandem palladium(II)-catalyzed oxidative carbocyclization/Diels-Alder
sequence using maleimide as dienophile.
6.1.5 Stereochemical Assignment
The stereochemical relation in 85 and 92 were determined using 2DNOESY experiments (Figure 8) The crucial NOE-interactions are indicated
on an energy-minimized 3D-model structure of 85. NOE-interactions
between the internal, olefinic-C-Ha and the TsN-C-Hb indicated that they
were indeed on the same side of the 5,6-fused ring system. In addition, TsNC-Hb displayed a clear NOE with the two adjacent α-carbonyl protons Hc
and Hd (Hd not shown) consistent only with the endo-isomer.
Figure 8. Selected 2D-NOESY interactions observed in tricyclic compound 85. Most
hydrogens are omitted for clarity.
6.1.6 Mechanistic Discussion
We believe that the mechanism of the oxidative carbocyclization of azaenallenes has much in common with the carbocyclization of enallenes
40
(Chapter 1.5.2). Attack of the allene on the palladium(II)-source would
initially form M28, which can undergo a cis-carbopalladation of the olefin to
give M29. β-Hydride elimination of this species (after bond rotation) forms
the exocyclic terminal olefin M30 (R = H). Under our reaction conditions
this species isomerizes (86b), possibly via hydro/dehydropalladation, to the
N-tosyl-protected pyrrole 86. In cases where the intermediate (σalkyl)palladium(II)-complex (M29) can form different β-hydride elimination
products, elimination occurs at the most easily accessible hydride oriented
syn to the palladium(II)-center (R = Me).
Scheme 36. Proposed mechanistic pathway for the palladium(II)-catalyzed
oxidative carbocyclization of aza-enallenes.
The above suggested mechanism has been based on cyclization of azaenallene 50 as this substrate lacked a substituent on the allylic moiety. The
absence of this substituent excludes formation of an initial (πallyl)palladium(II) intermediate, through syn C-H abstraction. Hence, the
latter mechanistic pathway can be considered highly unlikely.79
41
6.2 Conclusions
We have been able to considerably extend the palladium(II)-catalyzed
carbocyclization methodology to include a new substrate class. Azaenallenes underwent efficient oxidative carbocyclization to provide 5membered N-heterocycles when treated with catalytic amounts of Pd(OAc)2
and stoichiometric amounts of BQ. In most cases the aza-enallenes were
cyclized in high yield under mild conditions.
Low concentrations of BQ were found to be essential due to the presence
of an active 1,3-diene in the formed N-heterocycles. When the oxidation was
conducted with stoichiometric quantities of BQ, small amounts of the
undesired Diels-Alder product was observed in essentially all cases. If
excess amounts of BQ were used, a tricyclic Diels-Alder adduct could be
obtained in essentially quantitative yield.
Implementation of a biomimetic reoxidation system, based on BQtethered Co-catalyst (11) was also realized. This strategy has the advantage
of being a more environmentally friendly reaction because the use of “free”
BQ could be completely avoided. Furthermore, this catalytic system
minimizes the formation of the undesired Diels-Alder adduct and maximizes
the yield of N-heterocycles. Additionally, using the biomimetic reoxidation,
the carbocyclization of aza-enallenes could be performed with the addition
of activated dienophiles to render a tandem carbocyclization/Diels-Alder
reaction. This approach rendered tricyclic heterocycles in high yield and
good selectivity. Unfortunately, due to substrate-limitations, the
diastereoselectivty of the reaction could not been studied. We are currently
evaluating the possibility to use the pyrrole/hydropyrrole-type products in
the total synthesis of kainic acid.108
42
7.
Palladium(II)-catalyzed
Oxidative
Carboborylation/Carboarylation of Enallenes
(Paper IV-V)
Our group has previously been involved in palladium(II)-catalyzed oxidative
carbocyclizations of en-79 and dienallenes,80 both using stoichiometric
amounts of BQ or aerobic reoxidation protocols.42 We have shown earlier
that fully carbocyclic structures as well as N-heterocycles can be constructed
using this powerful methodology (Chapter 5 and 6). The main feature of this
methodology has been the selective formation of carbon-carbon bonds under
oxidative conditions, a process commented on in earlier parts of this thesis
(Chapter 1.5). Carbocycles obtained from the oxidative carbocyclization,
especially the 5,6- and 5,7-fused ring systems, are common motifs in
numerous interesting naturally occurring substances. A few of the more
representative examples are cyanthiwigin B109 and aphanamol101 (Figure 9).
cyanthiwigin B
O
O
H
H
aphanamol
enallene carbocyclization core
R
MeO
R
H
R
OH
Figure 9. Structures of cyanthiwigin B and aphanamol along with the core skeleton
obtained after palladium(II)-catalyzed oxidative carbocyclization of enallenes.
Our previous contributions to the field of oxidative carbocyclizations
have all involved catalytic cycles terminated with β-hydride elimination to
give the unsaturated carbo/heterocycle (Scheme 37, B) This fact should not
be considered a limitation, as the newly formed olefin can act as a diverse
chemical handle, opening up for further selective transformations.
Functionalization of the proposed (σ-alkyl)palladium(II) intermediate,
(Scheme 37, A) could potentially be achieved by promoting transmetallation
with a suitable transmetallation reagent. To the best of our knowledge, such
an oxidative approach has previously not been reported. However, a closely
related reaction has recently been disclosed by Cárdenas. In summary,
Cárdenas reported that enynes (93) and enallenes reacted under palladium43
catalysis together with bis-(pinacolato)diboron (B2pin2) and methanol to give
borylated carbocycles (95) in good yields with high chemoselectivity
(Scheme 38).110
Scheme 37. Rationalization of the palladium(II)-catalyzed
carbocyclization-transmetallation of enallenes.
oxidative
Another source of inspiration to this trapping approach originates from a
previous DFT-calculation conducted by our group regarding the
palladium(II)-catalyzed 1,4-carbohydroxylation of dienallenes (Chapter 4).87
In this study, the DFT-calculations suggested that the formed (πallyl)palladium(II) intermediate was coordinating to the pending (exocyclic)
alkene, making it more electron deficient and susceptible to nucleophilic
attack (Scheme 37, D).
Scheme 38. Cárdenas Palladium-catalyzed borylative carbocyclization of enynes.
Based on this interaction we reasoned that a similar palladium(II)-olefin
interaction could be present when the substrate is changed from dienallene to
enallene. Stabilization of the (σ-alkyl)palladium(II) intermediate could
potentially divert the reaction to undergo transmetallation in favor of syn-βhydride elimination (Scheme 37, C). If indeed possible, one could quickly
gain access to diverse borylated or arylated carbocycles, depending on the
choice of reagent, in one pot under oxidative conditions.
44
7.1 Palladium(II)-Catalyzed Oxidative
Carbocyclization/Borylation of Enallenes.
7.1.1 Development of a Catalytic
Carbocyclization/Borylation of Enallenes
System
for
Oxidative
We first settled for enallene 22 as the model substrate as it was relatively
easy to synthesize. In our first cyclization attempt, we found that treatment
of 22 with 5 mol% of Pd(OAc)2, 1 equiv. of B2pin2 and 1.5 equiv. of BQ in
THF at 40 °C for 4 hours gave rise to 40% of the desired overall cisborylated carbocycle 96 along with 8% of the β-hydride eliminated product
23 (Scheme 39). Remarkably, the borylated carbocyle was obtained as one
single diastereomer.
Scheme 39. Initial attempt of a palladium(II)-catalyzed oxidative borylative
carbocyclization of enallene 22.
Encouraged by these results, optimization of the overall reaction parameters
was undertaken, starting with the role of the solvent (Table 7). Clearly, the
choice of solvent played a crucial role for the selectivity between 96 and 23,
and toluene proved to be by far the most effective solvent (Table 7, entry 7).
In toluene, at 40 °C, 70% of 96 and only 5% of 23 was observed by1H NMR
analysis of the crude reaction mixture. Other solvents such as acetone and
acetonitrile showed overall poor selectivity between the two species (Table
7, entries 3 and 6). As anticipated, a range of solvents did provide relatively
good conversions, but with fluctuating selectivity between 96 and 23. Once
toluene had been established as the most effective solvent, the influence of
the palladium source was investigated. Previous studies on the
carbocyclization of en- and dienallenes have shown that Pd(OAc)2 and
Pd(O2CCF3)2 were the most suitable catalysts. This proved to be true in this
case aswell, as Pd(OAc)2 displayed the most promising conversion and
selectivity. Most other common Pd(II)-salts such as: PdCl2, PdCl2(CH3CN)2,
PdCl2(PPh3)2, Pd[(nacnac)(OAc)], Pd(acac)2 performed poorly.
45
Table 7. Solvent effects in the palladium(II)-catalyzed oxidative carbocyclization of
enallene 22.(a)
Solvent
Yield of 96 (%)(b)
1
THF
40
8
2
DCM
45
34
3
CH3CN
20
46
4
EtOH
56
13
5
EtOAc
55
12
6
Acetone
60
30
Entry
Yield of 23 (%) (b)
7
Toluene
70
5
(a) Reaction conditions (unless otherwise noted): Enallene 22 (1.0 equiv.), Pd(OAc)2 (5 mol%), B2pin2
(1.0 equiv.) and BQ (1.5 equiv.) in the indicated solvent at 40 °C for 4 h. (b) Yields were determined by
1
H NMR analysis of crude mixtures using anisole as the internal standard.
Next, our focus turned towards minimizing the catalyst loading.
Fortunately, the amount of catalyst could be lowered down to 1 mol% (at 40
°C) simply by extending the reaction time from 4 to 10 hours. Attempts were
also made to suppress the formation of the β-hydride elimination product 23
by modification of the protocol. First, different temperatures were evaluated
(Table 8), using 1 mol% Pd(OAc)2, and the ratio between
borylated/eliminated product (96/23) was determined from the crude mixture
by 1H NMR analysis. Elevated temperatures resulted in significantly higher
amounts of the undesired product 23 (Table 8, entries 3-4).
Table 8. Product distribution in the palladium(II)-catalyzed
carbocyclization of enallene 22 at various temperatures.(a)
Entry
T (°C)
Yield of 96 (%)
Yield of 23 (%) (b)
1
r.t
54
6(c)
2
40
79
7
3
60
62
14
oxidative
4
80
43
50
(a) Reaction conditions (unless otherwise noted): Enallene 22 (1.0 equiv.), Pd(OAc)2 (1 mol%), B2pin2
(1.0 equiv.) and BQ (1.2 equiv.) in the indicated solvent, 10 h (b) Yields were determined by 1H NMR
analysis of crude mixtures using anisole as the internal standard. (c) Full conversion could not be obtained
after 24 h.
After some further optimization we found that excess BQ (1.5 equiv.)
and/or excess B2pin2 (1.5 equiv.) did not improve the yield nor the
selectivity.
A range of commonly used oxidants are potentially capable of reoxidizing
Pd(0) back to Pd(II) in these reactions, but it has been observed on previous
occasions that quinones are the only class of oxidants that are truly effective
for the oxidative carbocyclization methodology. Owing to this, further
investigation of alternative oxidants was not attempted.111
46
Some common additives such as NaOAc, HOAc and H2O were
introduced to determine if the acidity of the medium, or the amount of water
present affected the reaction outcome. Of these additives, a slight increase in
rate was observed when water was introduced; however, the effect was
rather small and not investigated further. Nevertheless, it proved that the
reaction was not very sensitive to water, which in turn makes this
transformation robust and easy to perform (Scheme 40).
Scheme 40. Optimized conditions for the palladium(II)-catalyzed oxidative
borylative carbocyclization of enallenes.
7.1.2 Investigation of the Scope and Limitations
After having established an efficient catalytic protocol for the model
substrate 22 an investigation regarding the substrate scope was initiated
(Table 9). Overall this transformation displayed rather consistent results,
providing borylated carbocycles in good yields (Table 9).
Table 9. Scope of the palladium(II)-catalyzed oxidative borylative carbocyclization
of enallenes.(a)
Entry
Enallene
1
Carbocycle
Yield (%) (b)
77
E
2
E
•
78
97
3
68
47
Table 9 contd. Scope of the palladium(II)-catalyzed oxidative borylative
carbocyclization of enallenes.(a)
Entry
Enallene
Carbocycle
Yield (%) (b)
4
60
102/102’
(3:1)
5
58
E
6
E
•
5c, d
105
7
86
8
64e
9
89
10
63f
11
62f
(a) Reaction conditions (unless otherwise noted): Enallene (1.0 equiv.), Pd(OAc)2 (1 mol%), B2pin2 (1.0
equiv.), BQ (1.2 equiv.), toluene, 40 °C, 10 hours. (b) Isolated yield after column chromatography (c) Full
conversion could not be achieved within 36 hours. (d) Determined from crude reaction mixture. (e) 109
consisted of an inseparable mixture of allene/alkyne (1:1.6) where the alkyne does not react during
catalysis. (f) 20 h reaction time.
However, there are a few observations that require some further comments.
In most cases products originating from β-hydride elimination were detected.
The quantity of this product was always below 10%, and easily separated
from the desired product using chromatographic methods. It is also worth
mentioning that we could not detect any other side products arising from
palladium(0)-catalysis. Related palladium(0)-catalyzed reactions have been
reported previously by both Morken112 and Cárdenas110.
Terminally disubstituted olefin 105 displayed very poor compatibility
with our catalytic system, and only 25% conversion of the starting material
was observed after 36 hours (Table 9, entry 6). The major product turned out
48
to be the β-hydride elimination product and only around 5% of the desired
borylated compound was detected in the crude reaction mixture. This result
discouraged us to further pursue terminally disubstituted olefins in this
study.
Cyclic enallenes (Table 9, entries 1 and 9-11) performed much better and
the products were obtained with an overall stereospecific cis-addition of the
elements of carbon and boron to the double bond. Notably, the 5-, 6- and 8membered cyclic enallenes (111, 22 and 115) formed borylated carbocycles
with overall cis-stereochemistry (Table 9, entries 1, 9 and 11) while the 7membered enallenes (113) formed the borylated carbocycle with a transfused ring junction (Table 9, entry 10). Another clearly detectable trend with
regard to ring size, was a drop in yield for 7- and 8-membered rings (63%
and 62% respectively), whereas the 5- and 6-membered rings allowed for
formation of borylated carbocycles in much higher yield (89% and 77%
respectively). We do believe this behavior can be traced back to the
stabilizing effect of the exocyclic olefin on the (σ-alkyl)palladium(II)
intermediate (Scheme 42, M34). Through modeling we reasoned that the
5,5-fused ring-system should have the strongest chelating properties, and the
larger rings the weakest, all in accordance with our observed chemical
yields. This trend was also supported by the amounts of β-hydride
elimination products observed in each case. Enallenes 113 and 115 produced
significantly more side products than 22 and 111.
Acyclic, monosubstituted olefins cyclized as expected to give primary
alkylboronates in moderate yields (Table 9, entries 2-4). Introduction of
substituents on the internal carbon efficiently rendered the borylated
carbocycle with one new quaternary stereocenter (Table 9, entry 7). Clearly,
this substrate class was one of the most efficient, giving the desired product
108 in 86% isolated yield. We rationalize this result by the fact that the (σalkyl)palladium(II) intermediate lacks the presence of a β-hydrogen, i.e.
transmetallation can proceed smoothly without this eroding pathway.
Enallenes substituted on the olefin (E) with alkyl (103) or aryl groups
(109) rendered carbocycles 104 and 110 in moderate yield (Table 9, entries 5
and 8). These specific borylated carbocycles were obtained as sticky oils,
which made X-ray analysis impossible. The stereochemical relationship was
instead determined with the aid of 1H/13C NMR techniques (COSY, HSQC,
HMBC, 2D-NOESY and NOE-diff) together with molecular modeling.
Finally, we decided to investigate whether the previously developed azaenallenes (Chapters 5 and 6) could potentially undergo this novel
transformation. Surprisingly, treatment of 117 under the conditions
developed for the carbocyclic reaction (1 mol% of Pd(OAc)2, 1.0 equiv. of
B2pin2, 1.2 equiv. of BQ, toluene, 40 °C) rendered the borylated
hydropyrrole 119 in acceptable yield, albeit accompanied by large amounts
of the Diels-Alder adduct 119b (Scheme 41). Minor adjustments to the
catalytic protocol, mainly reduction of the amount of BQ to 1.05 equivalents,
49
allowed access to borylated dihydropyrrole in 71% isolated yield. In the
same manner aza-enallene 118 afforded 120 as one single isomer in 60%
isolated yield.
Scheme 41. Palladium(II)-catalyzed oxidative borylative carbocyclization of azaenallenes 117 and 118.
7.1.3 Stereochemical Assignment of the Borylated Carbocycles
The stereochemical relation of the carbocyclized products 96, 112 and 116
was determined by careful 1H NMR analysis. 2D-NOESY along with NOEdiff (3.5-10% NOE) experiments clearly revealed that the two bridgehead
protons, along with the pinB-C-Ha were situated on the same side (cis)
(Figure 10, 96, Hb and Hc).
Ha
Hb E
Hc
Bpin
96
E
Ha Hc
Bpin
E = CO2 Me
Selected J -couplings for 112
Hb
E
E
Ha : δ 1.30-1.24 ppm poorly resolved multiplet
Hb: δ 3.03 ppm (ddd), J = 11.2, 6.5, 5.3 Hz
Hc: δ 3.54 ppm (ddd), J= ~6.4 (*2), 2.5 Hz
112
Figure 10. Schematic representation of the bridgehead protons in the borylated
carbocycles 96 and 112.
Eventually, we managed to obtain crystals of compound 96 suitable for
X-ray analysis, which confirmed our stereochemical assignment (Figure 11,
left). In contrast to the 5,6-fused carbocycle, the 5,7-fused product 114 did
not at all show the same NOE-characteristics. Instead, interactions indicative
of a trans-ring junction and overall cis-carboborylation was observed. The
configuration of 114 was later confirmed by an X-ray structure obtained
after the original publication (Figure 11, right).113
The stereochemistry of the monocyclic borylated compounds 104, 110
and 120 were established by matching 2D-NOESY interactions with energy
minimized 3D-models along with J-couplings of the requisite compound, as
exemplified by the study of compound 110 (Figure 12).
50
Figure 11. X-ray structures for the borylated carbocycles 96 (left) and 114 (right).
Figure 12. Observed NOE-interactions for borylated carbocycle 110, indicated on
an energy minimized model (some hydrogens are omitted for clarity).
7.1.4 Mechanistic Discussion
Our proposed mechanism is derived from previous studies on the palladiumcatalyzed carbocyclizations of enallenes79 together with data disclosed in the
current study. We suggest that after π-coordination of Pd(OAc)2 cis to the
already present stereocenter (Scheme 42, M32), the metal undergoes an
allenylic C-H abstraction, denoted “allene attack” on palladium. This step
would produce the relatively stable (σ-vinyl)palladium(II) intermediate M33.
From M33, the olefin exclusively undergoes cis-insertion, most likely
mediated by coordination of one molecule of BQ. After insertion, a (σalkyl)palladium(II) intermediate M34 is formed. Normally, this intermediate
undergoes facile syn-ß-hydride elimination. However, our results strongly
indicate that the pending (exocyclic) olefin stabilizes this intermediate in a
way that allows for transmetallation with B2pin2. Notably, the cyclopentyl
enallene 111 provided the highest yield of the corresponding carbocycle. We
believe that this outcome can be explained by the relative stability of the
51
palladium intermediate (M34), where the 5,5-fused carbocycle have the most
stabilization and the 5,8-fused the least favored one.
Scheme 42. Suggested mechanism for the palladium(II)-catalyzed oxidative
borylative carbocyclization of enallenes.
Regardless of chelation, M34 undergoes transmetallation and subsequent
reductive elimination (with retention of configuration) to produce the
carboborylated products. Expelled palladium(0) is finally reoxidized by BQ
in the presence of catalytic amounts of HOAc to complete the catalytic
cycle.
One matter of controversy regarding this mechanistic pathway has been
whether transmetallation was occurring much earlier, for example already in
M32. Transmetallation at this stage would completely avoid having to pass
through transmetallation of a highly sensitive (σ-alkyl)palladium(II)
intermediate (M34). Although this pathway has been considered, our
hypothesis is that a more electrophilic catalyst, Pd(OAc)2 as opposed to
(pin)BPdOAc, is essential for the initial allene attack on palladium.
A second possible, but less likely pathway involves initial
transmetallation of Pd(OAc)2 to give (pin)BPdOAc (Scheme 43, M37).
Insertion of the cyclic olefin could occur, providing intermediate M38.
Another insertion, this time of the allene, (M39) followed by ß-hydride
elimination would then render the product. In our opinion the experimental
data does not support this mechanistic pathway, as we would expect (some)
borylation of the allene moiety, without formation of carbocycles. Also, this
52
particular mechanism does not account for the formation of the observed ßeliminated products.
In addition to the above discussed mechanisms, we have considered two
other pathways, one involving a Pd(IV) intermediate and one starting from
oxidative addition on B2pin2 by a palladium(0)-source. The high oxidation
pathway was judged highly unlikely using BQ as the oxidant. The second
alternative was ruled out as the expected borylated product was not observed
under palladium(0)-catalysis.
Scheme 43. Alternative pathway explaining the palladium(II)-catalyzed borylative
carbobocyclization of enallenes.
We should point out that the detailed mode of transmetallation has not
been investigated for this novel transformation. However, our belief is that
transmetallation is mediated by acetate, as previously described in the
Miyaura borylation.114
53
7.2 Palladium(II)-Catalyzed Oxidative
Carbocyclization/Arylation of Enallenes.
7.2.1 Arylboronic acids as Transmetallation Reagent.
After having realized the catalytic oxidative borylative carbocylization of
enallenes we turned our focus towards an arylative carbocyclization
sequence (Scheme 44). This attractive process could substantially widen the
substrate scope, and supply us with further knowledge regarding the
carbocyclization. Starting from the borylation protocol described in Chapter
7.1, simple re-optimization using arylboronic acids (ArB(OH)2) was
anticipated to give arylation instead of borylation. Our choice of model
substrate fell on enallene 97 as we expected that this substrate was least
prone to undergo ß-hydride elimination.
In preliminary experiments, using the same conditions as in the borylation
case, treatment of 97 with 5 mol% of Pd(OAc)2, 1.2 equiv. of BQ and 1
equiv. of PhB(OH)2 in toluene at 40 oC afforded the desired phenylated
carbocycle 121a in 53% yield. In this instance 121a was accompanied by
some unreacted starting material together with trace amount of the expected
by-product 122 (Scheme 44).
Scheme 44. First attempt on the palladium(II)-catalyzed arylative carbocylization of
enallene 97.
A look at different solvents quickly revealed that the transformation
proceeded in most solvents, except strongly coordinating acetonitrile.
Notably, very little ß-hydride eliminated product (122) was formed in all
cases. The conversion differed substantially for each case (Table 10). The
best solvent proved to be THF, rendering the desired product in 92% yield
after 16 hours together with 4% of 122 (Table 10, entry 6). During this
optimization, products arising from homo-coupling of PhB(OH)2 (biphenyls)
were indeed observed, albeit in a low yield of around 2-5% in each separate
catalytic reaction. Additional attempts to improve the performance of this
transformation were based on interchanging the boronic acid derivative.
Unfortunately, almost all other boronic acid derivatives such as boronic
esters and potassium trifluoroborates were all completely inactive for this
54
transformation, and only trace amounts of the desired arylated carbocycles
were detected.
Table 10. Solvent effects in the palladium(II)-catalyzed oxidative arylative
carbocyclization of enallene 97.(a)
Entry
Solvent
Yield of 121a (%)(b)
Yield of 122 (%)(b)
1
Acetone
90
7
2
Acetonitrile
0
0
3
DCM
67
4
4
DCE
64
5
5
1,4-Dioxane
86
2
6
THF
92
4
7
Toluene
53
1
(a) Reaction conditions (unless otherwise noted): Pd(OAc)2 (5 mol%), PhB(OH)2 (1.0 equiv.) and BQ
(1.2 equiv.) in the indicated solvent at 40 oC for 16 h (b) Yields were determined by 1H NMR analysis of
crude mixtures using anisole as the internal standard.
Most of the commercially available arylboronic acids are delivered as
mixtures of the “free” arylboronic acid (124) and the boroxine (123).
Fortunately no significant difference was observed when the reaction was
conducted with one or the other (Scheme 45).
Scheme 45. Optimized conditons for the palladium(II)-catalyzed oxidative arylative
carbocylization of enallenes.
7.2.2 Scope and Limitations
The reaction conditions in Scheme 45 were used to explore the scope of
the palladium(II)-catalyzed oxidative arylative carbocyclization. Both
electron-rich and electron-deficient arylboronic acids were evaluated and the
results are summarized in Table 11. Overall, monosubstituted enallene 97
was cyclized efficiently independent of the electronic properties of the
arylboronic acid (61-95% isolated yield).
Electron-rich boronic acids displayed marginally lower yields compared
to the electron-deficient variants. The least effective arylboronic acid, phydroxyphenylboronic acid, also being the most electron rich, did not reach
full conversion after 24 hours and the isolated yield was only 68%. In
55
addition to the effect of electron-rich substrates, a possible (small) metaeffect was noticed. Comparison of 2-Me- (121ab), 3-Me- (121ac), and 4Me-C6H4B(OH)2 (121ad) showed that the 2-Me boronic acid performed
worst (79%) followed by the para-substituted (84%) and finally by the metasubstituted (88%). 2-OMe-substituted boronic acid (Table 11, entry 5)
delivered the overall least efficient reaction in our series. Whether this is due
to the electron rich character and/or because of steric demands (meta-effect)
is not clear.
Table 11. Scope of arylboronic acids in the palladium(II)-catalyzed oxidative
arylative carbocylization of enallene 97.(a)
Entry
ArB(OH)2
Carbocycle
Time (h)
Yieldb (%)
1
C6H5
121aa
2
83
2
2-Me-C6H4
121ab
4
79
3
3-Me-C6H4
121ac
4
88
4
4-Me-C6H4
121ad
4
84
5
2-OMe-C6H4
121ae
4
61
6
3-OMe-C6H4
121af
4
87
7
4-OMe-C6H4
121ag
4
75
t
8
4- Bu-C6H4
121ah
5
82
9
4-TMS-C6H4
121ai
5
88
10
4-vinyl-C6H4
121aj
5
90
11
4-F-C6H4
121ak
5
90
12
4-Cl-C6H4
121al
5
82
13
4-Br-C6H4
121am
5
92
14
4-CF3-C6H4
121an
5
73
15
4-acetyl-C6H4
121ao
5
86
16
4-formyl-C6H4
112ap
5
95
17
3-NO2-C6H4
121aq
5
86
18
4-OH-C6H4
121ar
21
68c
(a) Reaction conditions: 97 (1.0 equiv.), Pd(OAc)2 (1 mol% ) BQ (1.1 equiv.), ArB(OH)2 (1.0 equiv.)
THF, 60 oC. (b) Isolated yield after column chromatography. (c) 1.3 equiv. ArB(OH)2 was used.
Electron-deficient boronic acids (Table 11, entries 15-17) did not
influence the reaction to any great extent apart from the slightly longer
reaction times required to reach full conversion. We were also pleased to
find that 4-formyl- (121ap) and 4-acetyl-substituted boronic acids (121ao)
56
performed well, as these functional groups might serve as useful chemical
handles in subsequent reactions (Table 11, entries 15-16).
Carbocyclizations employing halogenated boronic acids 121ak-121am
gave the desired carbocyles without any further optimization (Table 5,
entries 11-13,). Unfortunately, 4-I-C6H4B(OH)2, completely failed to
produce the desired product, and the starting material could be recovered
(not shown). Two other synthetically useful aryl boronic acids; 4-TMS
(Table 11, entry 9) and 4-vinyl (Table 11, entry 10) also provided the desired
arylated carbocycle in high yield without formation of side products arising
from secondary catalytic reactions of these functional groups.
After realizing that most arylboronic acids seemed compatible with our
protocol we wanted to investigate the limitations of the starting enallene. As
depicted in Table 12, chemical yields were noticeably lower when steric
bulk on the allene (Table 12, entry 6) or additional substituents on the olefin
were introduced. Internal olefin 107 produced the desired carbocycle, having
one new quaternary carbon, in good yield after extending the reaction time to
7 hours. A much longer reaction time (24 hours) was necessary to transform
the E-phenyl substituted enallene 109, and even after an extended reaction
time the corresponding carbocycle could be isolated in only 55% yield.
In previous studies regarding the oxidative borylation, we were searching
for products being borylated on the allene-moiety (Chapter 7.1.2). However,
such products were never observed in our crude reaction mixtures. In the
arylative catalytic variant such arylated side-products were observed for the
first time when enallene 109 was subjected to the catalytic reaction (Scheme
46, 126b). This type of by-product is of special interest when considering
various mechanistic pathways.
Cyclic enallenes (22 and 111) underwent the carbocyclization/arylation
sequence to give products with stereospecific cis-addition of the central
carbon of the allene and the phenyl (from phenylboronic acid) to the olefin
(Table 12, entries 4 and 5). For these instances a slight excess (1.3 equiv.) of
phenylboronic acid clearly improved the transformation.
Increasing the steric demand on the allenic side chain completely shut
down any catalytic activity (Table 12, entry 6). This was in sharp contrast to
our previous studies on the borylative carbocyclization (Chapter 7.1.2) as
well as the carbocyclization of aza-enallenes (Chapter 6.1.3), where this lack
of activity was not as profound. The only reasonable explanation we have for
this behavior is that in the arylation case, transmetallation might occur prior
to coordination of the enallene. Formation of a Ar-PdOAc complex will
probably suppress the “allene attack on palladium” (M40-M41), thus
derailing the catalytic reaction.
57
Table 12. Scope of different enallenes in the palladium(II)-catalyzed oxidative
arylative carbocyclization.(a)
Time (h)
Yieldb (%)
1
2
83
2
7
76
3
24
55
4
2
61
5
4
63
6
24
N.R.c
Entry
Enallene
Carbocycle
(a) Reaction conditions: Enallene (1.0 equiv.), Pd(OAc)2,(1 mol%), BQ (1.1 equiv.), PhB(OH)2 (1.0
equiv., entries 1 and 2), (1.3 equiv., entries 3 to 6), THF, 60 oC. (b) Isolated yield after column
chromatography. (c) No reaction based on 1H NMR analysis of the crude mixture, N. R = no reaction.
58
7.2.3 Mechanistic Discussion
We believe that the mechanism in operation for the oxidative arylative
carbocyclization (Scheme 46) is essentially the same as that suggested for
the carbocyclization/borylation case. There are however some observations
that require some comments. After initial coordination (M40) and
subsequent allene attack on Pd(OAc)2 a (σ-vinyl)palladium(II) intermediate
is obtained (M41). Probably, with certain substrates, this intermediate
undergoes transmetallation (slow) followed by reductive elimination to give
a new arylated-1,3-diene (126b).
E
E
E
E
Pd(OAc)2
•
red. elim.
- Pd(0)
Ph
E
Ph
Pd
Pd(0)
reoxidation
(BQ + HOAc)
E
coordination
E
Pd(II)/Pd(0)
catalytic cycle
additional ligands omitted for clarity
E = CO2Me
(II)
M43
E
•
AcO
Pd(II)
M40
AcOB(OH)2
transmetallation
E
E
E
E
allene attack
-HOAc
ArB(OH)2
-H elimination
AcO
E
Pd(II)
M42
BQ-coordination
cis insertion
Pd(II)
AcO
1) transmetallation
2) red. elim.
M41
E E
E
Observed
Observed
Ph
126b
Scheme 46. Possible mechanism for the palladium(II)-catalyzed oxidative arylative
carbocyclization of enallenes.
Despite this side product, the usual reaction pathway is still in operation.
After insertion of the olefinic side-chain (M41) the crucial π-olefin-chelated
(σ-alkyl)palladium(II) intermediate M42 is formed. This intermediate
undergoes facile transmetallation, eliminating AcOB(OH)2 followed by
reductive elimination to give the arylated carbocyle. Products from βhydride elimination were also observed during this catalytic reaction and we
believe they are being formed in the same manner as in the
carbocyclization/borylation reaction, hence through ß-hydride elimination of
M42.
59
7.3 Conclusions on the Oxidative Arylative/Borylative
Carbocyclization of Enallenes.
In conclusion, we have developed two new powerful protocols evolved from
our previous studies on palladium(II)-catalyzed oxidative carbocyclization of
en- and dienallenes. Initially, we developed an oxidative borylative
carbocyclization of enallenes. The overall catalytic protocol employs
minimal amounts of the catalyst, Pd(OAc)2, as well as minimal amounts of
both B2pin2 and BQ. Secondly, with minor adjustments of the reaction
parameters, mainly replacing B2pin2 with ArB(OH)2, the reaction could be
directed to produce arylated carbocycles. Gratifyingly, a range of
structurally and electronically diverse arylboronic acids proved effective for
this reaction.
Both procedures accepted a range of cyclic and acyclic enallenes.
Generally, carbocycles were formed in moderate to high yield, including
products with new quaternary carbon centers. In all cases the carbocycles are
formed as one single diastereomer. We have determined that both the
borylative and arylative carbocyclization proceeds exclusively through ciscarboborylation (or cis-carboarylation) of the olefin. This was done by
examining the stereochemistry of the products using NMR- and X-ray
techniques. Furthermore, a plausible mechanism for both the borylation- and
arylation-processes, based on the observed product distribution was
proposed.
60
8. Palladium(II)-Catalyzed Formation of
Oxazolidinones Through Cyclization of
Allylic Tosylcarbamates – Scope, Mechanistic
Aspects and Further Derivatization. (Paper VI)
Oxazolidinones are common precursors to 1,2-aminoalcohols, a functionality
present in many pharmaceuticals and natural products, as well as in the
fundamental backbone of many asymmetric ligands.115 Owing to the
abundance of this functionality, there are many efficient catalytic methods
available leading these structural motifs.115,116 One very popular approach
towards the oxazolidinone backbone involves the use of a palladium catalyst
in combination with allylic tosylcarbamates (130). This particular
methodology has been demonstrated to be highly modular, and catalytic
reactions involving Pd(0), Pd(II) and Pd(IV) are efficient for this cyclization.
A few recent representative examples of these processes are: aminocarbonylation117, amino-acetoxylation118 and amino-alkynylation119 (Scheme
47).
Scheme 47. Palladium-catalyzed formation of oxazolidinones from allylic
tosylcarbamates.
In the majority of the above-mentioned catalytic protocols, cheap and readily
available allylic alcohols are used as starting materials. However, this
methodology is not strictly limited to this class of alcohols. Several
61
examples involving palladium-catalyzed cyclization of carbamates bearing
alkynes120 and allenes121 are represented in the literature.122
A closer look at these previous protocols revealed that one fundamental
type of palladium(II)-catalyzed oxidative cyclization of allylic carbamates
was not yet described, namely formation of vinyl substituted oxazolidinones.
White37,123 recently published a range of cyclizations leading to the
formation of oxazolidinones, albeit using homoallylic carbamates (134) as
starting materials. For these examples, the cyclization proceeded via a (πallyl)palladium(II) intermediate (M44), formed by a palladium(II)-bissulfoxide (White catalyst, 15) mediated allylic C-H activation (Scheme
48).123a
Scheme 48. White's approach to vinyl-oxazolidinones through palladium(II)catalyzed C-H activation of homoallylic carbamates.
Our initial interest in the oxazolidinone skeleton emerged from previous
projects on the preparation and oxidative carbocyclization of aza-enallenes
(Chapters 5 and 6). More specifically, Trost96 and Hsung97 reported that
bromoallenes could be coupled to oxazolidinones in high yield using
copper(I)-catalysis., Unfortunately, neither Hsung´s nor Trost´s approach
included vinyl-oxazolidinones. With this in mind we initiated a project,
based on palladium(II)-catalyzed oxidation chemistry, which could
potentially give us access to a variety of vinyl-substituted oxazolidinones.
This would represent the first step towards a possible new substrate-class
suitable for palladium(II)-catalyzed carbocyclizations (Scheme 49).
Scheme 49. Palladium(II)-catalyzed formation of oxazolidinones and anticipated
further derivatizations.
Although this was the primary objective, we were also puzzled over the fact
that the oxidative approach, relying on β-hydride elimination to furnish the
oxazolidinone, appeared to be overlooked in the literature. We argue that the
62
development of a reaction terminated by ß-elimination would be a useful
addition to the already existing synthetic methodologies.
8.1 Results and Discussion
8.1.1 Finding Efficient Reaction Conditions
Crotyl carbamate 136 (E/Z = 12:1) was selected as the model substrate due
to its availability and low cost. Treatment of crotyl alcohol (1.0 equiv.) with
p-toluenesulfonyl isocyanate (TsNCO, 1.0 equiv.) in THF at room
temperature produced the target allylic tosylcarbamate as a white solid in
nearly quantitative yield. After considerable experimentation, reaction
conditions rendering measurable amounts of vinyl-oxazolidinone 137 were
found. The yields were at first very low, but solvent mixtures of THF and
DMSO along with addition of a weak base, in the form of NaOAc, improved
the efficiency.
At an early stage, a more detailed analysis of the formed by-products was
undertaken. This revealed that in addition to the desired 5-exo-trig
cyclization delivering the oxazolidinone, an allylic sulfonamide (138) was
also formed in almost equimolar amounts (Table 13).
Table 13. Finding working conditions for the Pd(OAc)2-catalyzed oxidative
cyclization of allylic carbamate 136.(a)
Entry
HOAc
NaOAc
Conversion
Distribution
(equiv.)
(equiv.)
(%)(b)
(137:138:139)(b)
1
0.5
none
100%
45:5:50
2
none
0.5
100%
40:20:40
3
2
1
100%
75:20:5
4
1
2
65%
55:35:10
5
7
1
100%
80:10:10
6
7
0.5
100%
85:10:5
7
3.5
0.5
100%
85:10:5
(a) Reaction conditions (unless otherwise noted): Allylic tosylcarbamate 136 (1.0 equiv.), 1 mol%
Pd(OAc)2, 1.5 equiv. BQ, HOAc/NaOAc, THF/DMSO (9:1), 50 °C, 12 h.(b) Determined by NMR
spectroscopy
63
Selective formation of this rearranged product had been previously reported
under Pd(OAc)2/LiBr catalysis by Lei84b, and seems to occur through a
decarboxylative (6-endo-trig, Aza-Claisen) Overman-type rearrangement.124
It was also observed that the starting tosylcarbamate appeared to undergo
hydrolysis under certain conditions to give the starting allylic alcohol 139. In
order to minimize formation of the Overman- and hydrolysis-products an
investigation of various reaction parameters was initiated. A short summary
is presented in Table 13. The results could be concluded as follows: If
catalytic amounts of NaOAc or HOAc were used separately (Table 13,
entries 1-2) high conversions and low amounts of the Overman-product 138
were observed together with large amounts of the starting allylic alcohol
139. The most positive effect was observed when mixtures of HOAc and
NaOAc were used (Table 13, entries 3-7). For example, a 2:1 mixture of
HOAc/NaOAc (equivalents, with respect to allylic tosylcarbamate) gave full
conversion of the carbamate with good selectivity for the oxazolidinone. The
optimal ratio between HOAc and NaOAc was eventually found to be around
3.5-7:0.5 (equiv.). Regarding the solvent, mixtures of THF/DMSO (9:1)
proved optimal. When less DMSO was used, lower conversions together
with formation of palladium black were observed. On the contrary, addition
of more DMSO had no further beneficial effect. Optimal reaction conditions
were found to be 1 mol% of Pd(OAc)2, 1.5 equiv. of BQ, HOAc/NaOAc (7
equiv./0.5 equiv.) in THF/DMSO (9:1) at 50 °C. Using this set of
parameters, full conversion was achieved within 12 hours, and further
extending the reaction time did not seem to erode the yield to any
measurable extent (Scheme 50).
64
Scheme 50. Optimal conditions for the palladium(II)-catalyzed cyclization of allylic
tosylcarbamate 136.
After having found efficient reaction conditions for the cyclization of
crotyl carbamate 136 (E/Z = 12:1) to give the oxazolidinone 137, the
influence of the stereochemistry of the double bond was investigated
(Scheme 51).
Scheme 51. Behaviour of E- and Z-140 in the palladium(II)-catalyzed cyclization of
allylic tosylcarbamates.
Z-140 was smoothly cyclized to give the oxazolidinone 141 in 95% yield
together with only trace amounts of the Overman-product (142). This was in
sharp contrast to the reactivity displayed by E-140 where 59% of the
oxazolidinone 141 was obtained, accompanied by 20% of the Overmanproduct 142. This behavior was explained through a 6-membered transition
state (Scheme 51), in which the Z-olefin suffers from 1,3-pseudo-diaxial
strain, hence, disfavoring the 6-endo-trig type Overman-isomerization. This
observation led us to use Z-substituted allylic alcohols for the subsequent
investigation regarding the scope and limitations. Finally, at this stage of the
investigation the initial 2-step procedure was replaced by a one-pot protocol
by avoiding isolation of the allyl tosylcarbamate.
8.1.2 Scope and Limitations
Confident that an efficient catalytic protocol had been developed, a
customary investigation of the scope and limitations was conducted (Table
14). Simple Z-substituted allylic alcohols 144 and (Z)-146 were smoothly
transformed into the target oxazolidinones in excellent yield and high E/Zselectivity (>20:1) using the one pot, two step procedure (Table 14, entry 2
and 4). Prenyl alcohol 147 was however considerably less reactive, and full
conversion could not be achieved even after prolonged reaction times or
65
increased catalytic loading (10 mol%) (Table 14, entry 5). Despite these
much more forcing conditions the expected oxazolidinone 148 was isolated
in only 40% yield.
Next the stereoselectivity of this transformation was probed using
different secondary (Z)-allylic alcohols (Table 14, entries 6-11). In all cases
tested, the reaction proceeded with excellent diastereoselectivity (>20:1)
rendering trans-oxazolidinones in 62-95% isolated yield. The high degree of
diastereoselectivity is a clear advantage of this approach, compared to the
less diastereoselective protocol described by White.123 In the array of allylic
carbamates evaluated, phenyl-substituted allylic alcohol 149 performed the
worst (62%). In this case we were not able to separate the phenyloxazolidinone derivative from the Overman-product.
Table 14. Investigation of the scope and limitations for the palladium(II)-catalyzed
cyclization of allylic tosylcarbamates.(a)
Entry
1
Alcohol
Oxazolidinone d.r.
-
Yield
(%)
83
Entry
Alcohol
Oxazolidinone
d.r.
Yield
(%)
20:1 79(d)
7
O
2
-
81
8
O
TBDMSO
NTs
20:1 78(d)
154
3
-
59
9
20:1 83(d)
4
-
95
10
20:1 81(b)
5
-
41(b)
11
46(b,c)
6
20:1 61(c,d)
(a) Carbamate formation: Allylic alcohol (1.0 equiv), TsNCO (1.05 equiv.), THF, r.t. 1 h. (b)
Cyclization conditions: Allylic tosylcarbamate (1.0 equiv.), 1 mol% of Pd(OAc)2, 1.5 equiv. BQ, 7 equiv.
AcOH, 0.5 equiv. NaOAc, THF/DMSO (9:1), 50 ºC, 12 h. (b) 10 mol% Pd(OAc)2, 48 h. (c) Hydrolysis
of the crude oxazolidinone was undertaken to facilitate purification of the product. (d) Same conditions as
(a) but 5 mol% Pd(OAc)2 and 48 h reaction time.
66
Instead, the crude oxazolidinone was immediately hydrolyzed to give the
1,2-amidoalcohol 150 as a crystalline solid.
Cyclic allylic alcohol 157 proved slightly less efficient (Table 14, entry
10). Increasing the catalytic loading from 1 mol% to 5 mol% of Pd(OAc)2
was enough to regain acceptable conversions. Although an extensive study
on the effect of peripheral functional groups were not undertaken, TMSprotected alcohol 153 was indeed compatible with the reaction conditions
(Tabe 14, entry 8), producing TMS-functionalized oxazolidinone 154 in high
yield and diastereoselectivity.
One attempt was made to cyclize homoallylic alcohols (159). Under our
catalytic conditions, without any further modification, the desired 6membered product could be obtained in moderate yield after prolonged
reaction times and increased catalytic loading (Table 14, entry 11).
Unfortunately, isolation of the cyclic carbamate proved impossible because
of partial hydrolysis during both extraction and purification on silica gel.
Instead, the analogous 1,3-amidoalcohol 160 was isolated after one
hydrolysis step in 46% yield. Nonetheless, this indicated that the catalytic
protocol was capable of forming 1,3-amido-alcohols, albeit in a rather
modest yield.
8.1.3 Biomimetic Reoxidation and Scale Up
To probe the possibility of an aerobic, i.e. green(er) catalytic process,
reaction conditions using molecular oxygen as the terminal oxidant was
briefly evaluated. Without any co-catalysts large amounts of hydrolyzed,
over-oxidized and rearranged products were observed in pure DMSO under
an atmosphere of oxygen.
Instead our focus turned to a biomimetic approach using a cobalt-based
metal macrocycle as electron transfer mediator. Replacing stoichiometric
amounts of BQ with catalytic amounts (10 mol%), together with 5 mol% of
Co(II)-salophen (161) under an atmospheric pressure of molecular oxygen
rendered the desired oxazolidinones in acceptable yields (Scheme 52).
Without essentially any further modification to the catalytic system, allylic
alcohols 139 and (Z)-146 provided oxazolidinone 137 and 141 in 60% and
75% yield, respectively, after 72 hours. The yields for the aerobic version
were slightly lower than that with stoichiometric amounts of BQ.
67
Scheme 52. Palladium(II)-catalyzed cyclization of allylic tosylcarbamates using an
aerobic (biomimetic) reoxidation system.
As we were aiming to use the substrates for further functionalization
within various other projects, we decided to scale up this process. Linear
scale up of our small scale reaction proved reasonably effective, with one
minor modification: Formation of the starting allylic tosylcarbamate was
conducted at 0 °C instead of room temperature. Two different substrates
were used for this scale up 139 (27.7 mmol) and 144 (34.8 mmol). The
corresponding oxazolidinones were isolated in 75% and 78% yield,
respectively (Scheme 53). The desired products were collected by simple
precipitation from ethyl acetate (after extraction), hence no further
chromatographic purification was necessary.
Scheme 53. Scale up of the palladium(II)-catalyzed cyclization of allylic
tosylcarbamates 139 and 144.
8.1.4 Further Transformations of the Oxazolidinones
To further extend the synthetic utility of the developed oxidative cyclization,
a range of subsequent transformations was developed. First an enzymatic
kinetic resolution (KR) protocol was investigated for substrate 137 (Scheme
54). Our approach can be outlined as follows: Hydrolysis of oxazolidinone
137 under standard conditions rendered the 1,2-amidoalcohol (rac-162),
which was then subjected to Lipase AK (Amano 20) along with an
acyldonor (164) in toluene at room temperature for 20 hours. At 51%
conversion the amidoacetate (R)-163 was isolated in 42% yield with 89% ee.
The remaining unreacted amidoalcohol (S)-162 was isolated in 48% yield
68
and 93% ee. The E-value for this kinetic resolution was determined to be
around 60.
Scheme 54. Hydrolysis and kinetic resolution of oxazolidinone 137 using Lipase AK
(Amano 20).
A few oxazolidinones were also subjected to a few other synthetically
interesting transformations. First a Wacker-type oxidation (Scheme 55, A)
was developed using conditions previously reported by Feringa.125 Indeed
this gave rise to the expected methyl ketone 165 in 75% isolated yield.
Selectivity for the ketone (vs. aldehyde) was high (95:5) and the reaction
was complete within 16 hours.
Scheme 55. Further transformations of tosyl-protected oxazolidinones.
Oxazolidinone 137 was also subjected to a cross-metathesis (Scheme 55,
B) reaction with methyl acrylate. Addition of 5 mol% Hoveyda-Grubbs 2nd
generation catalyst in two portions (2.5 mol% each time), together with 137
(1 equiv.) and methylacrylate (3 equiv.) in DCM produced the desired α-ßunsaturated compound 166 in an acceptable yield of 69% with a 10:1
selectivity for the E-isomer. This particular starting material is currently
69
being evaluated further in the total synthesis of (+/-)-Kainic acid (Scheme
55, 79).126
Inspired by recent achievements in oxidative Heck-chemistry, 137 was
subjected to an oxidative Heck-type reaction (Scheme 55, C) using BQ as
oxidant and phenylboronic acids as the arylsource.127,48,60 This transformation
was originally intended to be conducted as a one-pot reaction coupled to the
cyclization, however we were unable to obtain acceptable yields using this
approach. Instead a two-step sequence was developed. Treatment of 137
with 5 mol% of Pd(OAc)2, 10 mol% of 1,10-phenanthroline (L9), 1.3 equiv.
of ArB(OH)2 and 1.5 equiv. of BQ in a THF/DMSO mixture (1:1) at room
temperature gave rise to (E)-arylated oxazolidinones (167a-e) in high yields
(87-96%) with E/Z-selectivites around 15:1. An array of diverse arylboronic
acids, including TMS-groups and chlorides, proved compatible with this
transformation.
Finally, and most importantly, an easy and reliable procedure for the
removal of the p-toluenesulfonyl (Ts) group was investigated. The classical
deprotection strategy using sodium/napthalene at low temperatures is
generally rather time consuming and impractical.128 Ragnarsson et. al.
reported that magnesium in MeOH under ultrasonic conditions was capable
of cleaving various tosyl-carbamates.129 When oxazolidinone 145 was
subjected to magnesium powder in MeOH at room temperature, cleavage
along with large amounts of the tosyl protected 1,2-amidoalcohol (169) was
observed (Scheme 56). Gratifyingly, the undesired hydrolysis product could
be almost completely suppressed when the reaction was conducted at 0 °C in
dry MeOH. Using these conditions the deprotected oxazolidinone 168 was
obtained in 82% isolated yield as a colorless solid.
Scheme 56. Deprotection of tosyl-oxazolidinone 145 under ultrasonic conditions.
8.1.5 Stereochemical Analysis and X-ray Structures
Most of the oxazolidinones described in this study are known in the
literature and good spectrometric data is available for either identical
compounds or structurally related derivatives. Although compound 158 is
known and has appeared in a series of publications, the NMR-shifts reported
were not of satisfactory quality to be compared with our results. Instead of
relying on previous data, crystals were grown and subjected to X-ray
70
analysis. Indeed this confirmed the cis stereochemistry of 158, as already
indicated by NMR spectroscopy (Figure 13, left).130 The trans-relationship
of the remaining oxazolidinones was assigned based on the X-ray structure
of 156131 together with known 1H NMR coupling constants (Figure 13,
right).123b
Figure 13. X-ray structure of 158 (left) and 156 (right).
8.1.6 Deuterium Labeling and Mechanistic Discussion
Palladium is known to undergo both cis- and trans-amino/amidopalladation
depending on the characteristics of the amide and reaction parameters. We
set out to investigate which mechanistic pathway was in operation for this
novel cyclization. To illuminate this aspect, trans-deuterated cyclic allylic
alcohol trans-175-d1 was synthesized as outlined in Scheme 57.132
Scheme 57. Synthesis of trans-175-d1 form 1,3-cyclohexadiene.
The synthesis was achieved through cis-1,4-acetoxychlorination38c of 1,3cyclohexadiene to give 171 followed by DIBAL-H mediated cleavage of the
acetyl-group to give cis-1,4-chloroalcohol cis-172. Protection of the alcohol
with TBDMSCl and imidazole then gave cis-173. Finally, SN2 type
nucleophilic substitution of the chloride with a deuterium and subsequent
deprotection of the TBDMS-group was achieved using LiAlD4 and TBAF to
give trans-175-d1.
71
The trans-deuterated allylic alcohol (trans- 175-d1) was then subjected to
the standard cyclization conditions described in Scheme 58. Analysis of 158
after the reaction revealed that the deuterium atom had been completely
removed.
Scheme 58. Investigation of the mechanism for palladium(II)-catalyzed cyclization
of trans-176-d1.
It has been shown that the stereochemistry of true aminopalladations are
always trans.133 On the contrary, in the case of amidopalladations such as
sulfonamidopalladation and sulfoacetamidopalladation the addition to the
double bond can occur with either cis37a,134 or trans135,118 stereochemistry.
Stahl and coworkers conducted an elegant and systematic study (similar to
the work by Stoltz for oxypalladation46d) on the aza-Wacker addition of
different sulfonamides to double bonds. Their work concluded that many
different parameters guide the stereochemistry of the aminopalladation
step.136
Having performed the cyclization of trans-176-d1 and observed complete
removal of the deuterium we can only argue that the mechanism has to occur
through either trans-amidopalladation (Scheme 59, upper pathway) or via
some form of π-allyl intermediate (Scheme 59, middle). This conclusion rely
on the simple fact that the final step, ß-elimination, occurs with a
hydride/deuteride aligned syn to the palladium(II)-center. As indicated in the
cis-amidopalladation pathway (Scheme 59, M49 + M50), the palladium
metal center cannot align syn to the deuterium, hence they are anti, and ßelimination will occur with the hydride.
For trans-amidopalladation, coordination of palladium(II) to the olefin
occurs prior to external attack (trans) by the tosyl-carbamate. In this case the
deuterium and the metal center will end up on the same side (syn) and the
deuterium is expected to be removed through ß-deuteride elimination
(Scheme 59, M45-M46). The intermediate π-allyl pathway (Scheme 59,
middle) could possibly give rise to 158 from trans-176-d1, however control
experiments using the homoallylic carbamate 177, instead of the allylic
derivative, lead us to the conclusion that this pathway was probably not in
operation (Scheme 60).
72
Scheme 59. Possible mechanistic pathways for the palladium(II)-catalyzed
cyclization of allylic tosylcarbamate trans-176-d1.
Some additional evidence strengthening the trans-amidopalladation
reaction pathway was indirectly obtained. Clearly, the developed process
displayed high diastereoselectivity (>20:1), which is in sharp contrast to the
related protocol developed by White. In the latter case, diastereoselectivity
was generally around 6:1. This might indicate a different mechanistic
pathway between the two different reactions. White has shown that their
catalytic reaction is operating through C-H activation, forming a (πallyl)palladium(II) intermediate.
Scheme 60. Attempted
tosylcarbamate 177.
palladium(II)-catalyzed
cyclization
of
homoallylic
73
Taking all of our experimental data into account we conclude that the
reported palladium(II)-catalyzed aza-Wacker cyclization proceeds through
trans-amidopalladation followed by β-H elimination.
8.2 Conclusions
Starting from readily available allylic alcohols and tosyl isocyanate
(TsNCO) a highly efficient cyclization protocol employing as little as 1
mol% Pd(OAc)2 has been developed. Overall, the protocol leads
oxazolidinones in isolated yields ranging from 40-95% with excellent
diastereoselectivity (>20:1). This particular catalytic process was easily
scaled up to 35 mmol (of the starting allylic alcohol) without any substantial
loss in efficiency.
In addition to the study of the cyclization of allylic tosyl carbamates, a
variety of subsequent transformations of the acquired oxazolidinones have
been investigated. For example, a palladium(II)-catalyzed ketone-selective
Wacker-oxidation and a palladium(II)-catalyzed oxidative Heck-type
reaction were successfully realized. A mild and operationally simple tosyl
deprotection strategy using only magnesium in MeOH under ultrasonic
irradiation was also developed.
In order to determine the actual sterical mode of amidopalladation (trans
vs. cis.), the mechanism was studied through deuterium labeling
experiments. The requisite deuterated compound was synthesized and
characterized. Results from the deuterium-labeling study strongly indicate
that the reaction occurs via an overall trans-amidopalladation.
We are currently testing whether the deprotected oxazolidinones will
undergo coupling with bromoallenes to give well-defined aza-enallenes,
suitable for subsequent palladium(II)-catalyzed oxidative carbocyclizations.
74
10 Concluding Remarks
To begin with, we have developed a stereoselective palladium(II)catalyzed oxidative 1,4-carbohydroxylation of allene-substituted 1,3-dienes
(dienallenes). This novel catalytic carbocyclization rendered hydroxylated
carbocycles in high yield. Two different reoxidation-protocols were
evaluated, one relying on stoichiometric amounts of BQ, and the other using
molecular oxygen as terminal oxidant. Remarkably, both reoxidationsystems proved equally efficient with respect to conversion and chemical
yield. Most importantly, the developed protocol showcases that water can act
as a nucleophile on (π-allyl)palladium-complexes, a process which is rarely
observed.
The second and third chapter involved development of new synthetic
methodology towards a palladium(II)-catalyzed carbocyclization of azaenallenes. In order to obtain the starting aza-enallenes, we first successfully
developed a copper(I)-catalyzed coupling of allylic sulfonamides with
bromoallenes. In addition to the study of the scope and limitations of this
process, the stability of the acquired aza-enallenes was determined. The
requisite aza-enallenes were then used in a palladium(II)-catalyzed oxidative
carbocyclization. After considerable experimentation, catalytic conditions
rendering hydropyrroles in good yield were found. An array of aza-enallenes
proved compatible with the developed protocol. In a similar fashion as
discussed above, reoxidation was achieved by the use of either
stoichiometric quantities of BQ or a biomimetic reoxidation system utilizing
O2. After realizing an efficient biomimetic reoxidation, employing a BQtethered metal macrocycle, we also developed a tandem oxidative
carbocyclization/Diels-Alder reaction using activated dienophiles.
By combining information from all our previous enallene
carbocyclizations, and conclusions drawn from a DFT-study, an oxidative
arylative/borylative carbocyclization of enallenes was successfully
developed. This transformation was achieved by introducing
transmetallation reagents such as B2pin2 or arylboronic acids to the oxidative
carbocyclization reaction. The reaction proved highly stereoselective and
allowed for formation of new carbon-carbon and carbon-boron bonds. The
mechanism behind this novel reactivity has been suggested to involve a (σalkyl)palladium(II) intermediate being stabilized by a pending olefin. This
interaction probably opens up a path to achieve transmetallation.
75
In the final chapter a new cyclization methodology was developed for the
formation of vinyl substituted oxazolidinones. This Pd(OAc)2-catalyzed
process proved efficient and readily scalable. It was demonstrated that the
reaction proceeds through trans-amidopalladation.
76
Reprint Permissions (Appendix A)
Paper I - Reprinted with permission from:
Water
as
Nucleophile
in
Palladium-Catalyzed
Oxidative
Carbohydroxylation of Allene-Substituted Conjugated Dienes
Julio Piera, Andreas Persson, Xisco Caldentey, and Jan-E. Bäckvall
J. Am. Chem. Soc., 2007, 129 (46), pp 14120–14121
Copyright (2007) American Chemical Society.
Paper II - Reprinted with permission from:
Copper-Catalyzed N-Allenylation of Allylic Sulfonamides
Andreas K. Å. Persson, Eric V. Johnston and Jan-E. Bäckvall
Org. Lett., 2009, 11 (17), pp 3814–3817
Copyright (2009) American Chemical Society.
Paper III - Reprinted with permission from:
Palladium(II)-Catalyzed Oxidative Carbocyclization of Aza-Enallenes
Andreas K. Å. Persson and Jan-E. Bäckvall
Angew. Chem. Int. Ed. 2010, 49, 4624–4627
Copyright (2010) John Wiley and Sons.
Paper IV - Reprinted with permission from:
Palladium-Catalyzed Oxidative Carbocyclization/Arylation of Enallenes
Tuo Jiang, Andreas K. Å. Persson, and Jan-E. Bäckvall
Org. Lett., 2011, 13 (21), pp 5838–5841
Copyright (2011) American Chemical Society.
Paper V - Reprinted with permission from:
Palladium-Catalyzed Oxidative Borylative Carbocyclization of
Enallenes
Andreas K. Å. Persson, Tuo Jiang, Magnus T. Johnson and Jan-E. Bäckvall
Angew. Chem. Int. Ed. 2011, 50, 6155 –6159
Copyright (2011) John Wiley and Sons.
77
Acknowledgements
I would like to express my sincerest gratitude to the following people:
My supervisor, Prof. Jan-Erling Bäckvall for accepting me as a Ph. D.
candidate in your group. Thanks for all the inspiration, motivation, support,
and for letting me work on challenging and interesting projects.
My collaborators and friends
Jan Deska, what would I have done without you? Thanks for the massive
support you have given me during these years. I would be much less of a
chemist without your wise words. Also, thanks for taking such good care of
me during my 3 month stay at Uni. Cologne, unforgettable! Wish you the
best in the future and I am looking forward to future discussions. “allenes on
top of the world!”
Julio Piera Balaguer, for being the number one diploma-work supervisor
and friend. My time at Stockholm University would not have been the same
without you and your epic quote (one out of many): “relax your long body”.
It has been a true pleasure to have met and worked with you.
Tuo Jiang, for the collaboration on the allene-projects, and for being an
exceptionally good diploma student! I was very lonely with my allenes
before you came and cheered me up…
Annika Träff, for all the nice discussions, lunches, “spex” etc (long list!).
Most importantly, thanks for providing me with a tiny bit of your “monster”
organization skills, they have been much needed! I am pretty sure we will
see more of each other in the future…
Reanud Millet, for being a good friend both inside and outside the lab.
Chemistry was good, fondue was great, and the rest was excellent!
Magnus Johnson, for supplying X-ray structures with such enthusiasm!
Our conference trips have been great and I hope our paths will cross again
both on a personal and professional level.
78
My co-workers for good collaborations on various projects: Tuo Jiang,
Teresa Bartholomeyzik, Antoine Joosten, Julio Piera Balaguer, Xisco
Caldentey, Lisa K. Thalén, Eric Johnston, Magnus Johnson and Youqian
Deng.
My friends and colleagues who have improved and commented on this
thesis: Jan Deska, Teresa Bartholomeyzik(!), Nanna Ahlsten, Tuo Jiang,
Johanna Persson and Renaud Millet. Your help has been priceless!
My diploma workers and Erasmus-students Tuo Jiang, Lidya Aho Soysal,
Paula Martirez, and last but definitely not least: Natascha “spatz” Bruckner.
It has been a pleasure working with each and every one of you.
All the people at the Department (past and present), for making my time
here at Stockholm University very enjoyable!
All the people who made my stay at Uni. Cologne such a great time: Jan,
Chicco, Axel, Philip, David, Thomas, Sabrina, Lars, Daniel, Benedikt and
Matthias. Best of luck to all of you!
Financial support from K & A Wallenberg´s travel scholarship, C. F.
Liljevalch travel scholarship, The Swedish Chemical Society travel
scholarship, Ångpanneföreningens (ÅF) scholarship, Astra Zeneca travel
scholarship in memory of Nils Löfgren and finally the Berzelii Center
EXSELENT for traveling grants.
All past and present members of the JEB-group: Jan, Madde, Krisztián,
Patrik, Katrin , Christine, Dirk, Julio, Lisa, Linnéa, Jenny, Karin L., Annika,
Karin E., Tuo, Teresa, Sasha, Maria, Jonas, Anuja, Tony, Anders, Deng,
Min, Baneesh, Ibrar, Jean-Baptiste, Renaud, Antione, Alex, Ylva, Richard,
Micke, Byron, Hoa, Bao-Lin, Markus, Oscar, Eric, Xisco, Erik, Yoshi,
Mozaffar, Xu, Thanks for making my time at the department great! (I hope I
didn´t miss anyone!)
The TA-staff: Martin, Britt, Louise, Christina, Kristina, Joakim, Pia and
Olle. A special thanks goes out to: Martin Roxengren for making the
aluminum heating blocks, and to Britt Eriksson for always keeping track of
the administrative things that I don’t understand.
My lecturers at MdH: Simon Dunne and Sarah Angus-Dunne for giving
me the best education in organic chemistry.
79
All the other MdH-people who made my time there so enjoyable: Jocke
a.k.a. “Bosse”, Jonas a.k.a. “Bosse”, Linda, Martina, Kristofer, Henrik and
Therese.
The “French dinner mafia” Richard Lihammar, Renaud Millet, Antoine
Joosten.
The monthly (pseudo-scientific?) Akkurat beer gang: Alex Kasrayan,
Patrik Krumlinde, Gaston Lavén and Julio Piera Balaguer. We continue,
right?
All the members of ÅGC: Kalle, Johan, Tälje, Hage, Jens, Macke and
Svenne for being the best friends in the world, period!
All the Östlund family (and affiliates) Perra, Annika, Magda, Staffan,
Marita and Fredrik. for nice dinners and gatherings:
My mother (Carina Mårtensson), father (Åke Persson) and sister (Lina
Persson) for your unconditional love, mental support and believing in me
from day one. It is always a pleasure travelling “back south” to let go of
some steam and chat about life in general.
Finally…
To my wife: Johanna Thank you for all the support, love, tears and
laughs. My 8 years with you have been the best of my life and I’m sure it
will only become better…Älskar dig!
80
References
(1)
(2)
(3)
(4)
(5)
(6)
(7)
(8)
(9)
(10)
(11)
(12)
(13)
(14)
(15)
(16)
(17)
Following ACS Standard Abbreviation/Acronyms, 2012 Guidelines for
Authors, Organic Letters. Organic Letters Homepage.
http://pubs.acs.org/page/orlef7/submission/authors.html. (accessed Jan 2012)
Atkins, P.; De Paula, J. Atkins Physical Chemistry, 7th ed.; Oxford
University Press: New York, 2000, pp 830-861.
Berzelius, J. J. Ann. Chem. et. Phys. 1836, 61, 146.
Rothe, J. F.; Caddock, J. H.; Hershman, A.; Paulik, F. E. Chem. Tech. 1971,
1, 600.
Jones, J. H. Platinum Metals Rev. 2000, 44, 94.
Crabtree, R. H.; The Organometallic Chemistry of the Transitions Metals, 4th
ed.; John Wiley and Sons, New Jersey, 2005 pp 1-53.
Rothenberg, G.; Catalysis: Concepts and Green applications, Wiley VCH
Verlag, Weinheim, 2008, pp 39-177.
Thomas, J. M. Angew. Chem. Int. Ed. 1994, 33, 913.
Hartley, F. R.; Chemistry Of the Platinum Group Metals: Recent
Developments; Elsevier, Amsterdam, New York 1991, 11.
(a) Kolbe, H. J. Prakt. Chem. 1871, 4, 418. (b) Saytzeff, M. J. Prakt. Chem.
1873, 6, 128.
(a) Smidt, J.; Hafner, W.; Jira, R.; Sedlmeier, J.; Sieber, R.; Rutinger, R.;
Kojer, H. Angew. Chem. 1959, 71, 176. (b) Smidt, J.; Hafner, W.; Jira, R.;
Sieber, R.; Sedlmeier, J.; Sabel, A. Angew. Chem. Int. Ed. 1962, 1, 80.
Negishi, E. -i.; In Handbook of Organopalladium Chemistry for Organic
Synthesis, Negishi, E. -i., Eds; John Wiley & Sons: New York, 2002, 1, pp 315.
de Meijere, A.; Diederich, F.; Metal-Catalyzed Cross-Coupling Reactions;
Wiley-VCH: Weinheim, 2004; 1, Chapters 1-6.
(a) Heck, R. F; Nolley, J. P. J. Org. Chem. 1972, 37, 2320. (b) Hermann, W.
A.; In Applied Homogenous Catalysis with Organometallic Compounds, 2nd
ed; Cornils, B.; Hermann, W. A. Eds; 2002, 1, pp 775-793.
Mizoroki, T.; Mori, K.; Ozaki, A. Bull. Chem. Soc. Jpn. 1971, 44, 581.
(a) Negishi, E. -i.; King, A. O.; Okukado, N. J. Org. Chem. 1977, 42, 1821.
(b) Negishi, E. -i. J. Organomet. Chem. 2003, 687, 229.
(a) Miyaura, N.; Yamada, K.; Suzuki, A. Tetrahedron Lett. 1979, 20, 3437.
(b) Hermann, W. A.; In Applied Homogenous Catalysis with Organometallic
Compounds, 2nd ed; Cornils, B.; Hermann, W. A. Eds; 2002, 1, 591-598.
81
(18)
(19)
(20)
(21)
(22)
(23)
(24)
(25)
(26)
(27)
(28)
(29)
(30)
(31)
(32)
(33)
(34)
(35)
82
(a) Campbell, M. G.; Powers, D. C.; Raynaud, J.; Graham, M. J.; Xie, P.;
Lee, E.; Ritter, T. Nature Chem. 2011, 3, 949. (b) Powers, D. C.; Ritter, T.
Top. Organomet. Chem. 2011, 35, 129.
Canty, A.; In Handbook of Organopalladium Chemistry for Organic
Synthesis, Negishi, E. -i.; Eds; John Wiley & Sons: New York, 2002, 1, pp
189-211.
Muñiz, K. Angew. Chem. Int. Ed. 2009, 48, 9412.
Negishi, E. -i.; In Handbook of Organopalladium Chemistry For Organic
Synthesis, Negishi, E. -i.; Eds; John Wiley & Sons: New York, 2002, 1, pp
17-46.
(a) MacManus, H. A.; Guiry, P. J. Chem. Rev. 2004, 104, 4151. (b) Trost, B.
M.; Van Vranken, D. L. Chem. Rev. 1996, 96, 395.
Tietze, L. F.; Ila, H.; Bell, H. P. Chem. Rev. 2004, 104, 3453.
Arends, I. W. C. E.; Sheldon, R. A. In Modern Oxidation Methods, 2nd ed,
Bäckvall, J. E.; Eds; Wiley-VCH, Weinheim, 2010, pp 147-180.
(a) Sigman, M. S.; Jensen, D. R. Acc. Chem. Res. 2006, 39, 221. (b) Schultz,
M. J.; Sigman, M. S. Tetrahedron 2006, 62, 8227.
Stahl, S. S. Angew. Chem. Int. Ed. 2004, 43, 3400.
(a) Nishimura, T.; Onoue, T.; Ohe, K.; Uemura, S. Tetrahedron Lett. 1998,
39, 6011.
Jensen, D. R.; Schultz, M. J.; Mueller, J. A.; Sigman, M. S. Angew. Chem.
Int. Ed. 2003, 42. 3810.
Peterson, K. P.; Larock, R. C. J. Org. Chem. 1998, 63, 3185.
Brink, G. J. T.; Arends, I. W. C. E.; Sheldon, R. Science, 2000, 287, 1636.
(a) Takacs, J. M.; Jiang, X. T. Curr. Org. Chem. 2003, 7, 369. (b) Tusji, J.
Synthesis, 1984, 369.
(a) Gligorich, K. M.; Sigman, M. S. Chem. Commun. 2009, 3854. (b) Sigman
M. S.; Werner, E. W. Acc. Chem. Res. DOI: 10.1021/ar200236v. (c) Cornell,
C. N.; Sigman, M. S. Inorg. Chem. 2007, 46, 1903.
(a) Grennberg, H.; Bäckvall, J. E. Chem. Eur. J. 1998, 4, 1083. (b) Byström,
S.; Larsson, M.; Åkermark, B. J. Org. Chem. 1990, 55, 5674. (c) Bäckvall, J.
E.; Hopkins, R. B.; Grennberg, H.; Mader, M.; Awasthi, A. K. J. Am. Chem.
Soc. 1990, 112, 5160.
(a) Chen, M. S.; White, M. C. J. Am. Chem. Soc. 2004, 126, 1346. (b)
Vermeulen, N. A.; Delcamp, J. H.; White, M. C. J. Am. Chem. Soc. 2010,
132, 11323. (c) Stang, E. M.; White, M. C. Nature Chem. 2009, 1, 547. (d)
Delcamp, J. H.; White, M. C. J. Am. Chem. Soc. 2006, 128, 15076. (e) Chen,
M. S.; Prabagaran, N.; Labenz, N. A.; White, M. C. J. Am. Chem. Soc. 2005,
127, 6970,
(a) Pilarski, L. T.; Janson, P. G.; Szabó, K. J. J. Org. Chem., 2011, 76, 1503.
(b) Pilarski, L. T.; Selander, N.; Böse, D.; Szabó, K. J. Org. Lett., 2009, 11,
5518.
(36)
(37)
(38)
(39)
(40)
(41)
(42)
(43)
(44)
(45)
(46)
(a) Hegedus, L. S.; Allen, G. F.; Bozell, J. J.; Waterman, E. L. J. Am. Chem.
Soc. 1978, 100, 5800. (b) Larock, R. C.; Hightower, T. R.; Hasvold, L. A.;
Peterson, K. P. J. Org. Chem. 1996, 61, 3584. (c) van Benthem, R. A. T. M.;
Hiemstra, H.; Longarela, G. R.; Speckamp, W. N. Tetrahedron Lett. 1994,
35, 9281.
(a) McDonald, R. I.; Liu, G.; Stahl, S. S. Chem. Rev. 2011, 111, 2981. (b)
White, P. B.; Stahl, S. S. J. Am. Chem. Soc. 2011, 133, 18594. (c) Reed, S.
A.; White, M. C. J. Am. Chem. Soc. 2008, 130, 3316. (d) Rice, G. T.; White,
M. C. J. Am. Chem. Soc. 2009, 131, 11707. (e) Reed, S. A.; Mazzotti, A. R.;
White, M. C. J. Am. Chem. Soc. 2009, 131, 11701. (f) Liu, G.; Yin, G.; Wu,
L. Angew. Chem., Int. Ed. 2008, 47, 4733
(a) Bäckvall, J. E. In Metal-Catalyzed Cross-Coupling Reactions, 2nd ed, deMeijere, A.; Diederich, F.; Eds; 2004, 2, pp 479-529.
(b) Bäckvall, J. E.; Nordberg, R. E. J. Am. Chem. Soc. 1981, 103, 4959. (c)
Bäckvall, J. E.; Byström, S. E.; Nordberg, R. E. J. Org. Chem. 1984, 49,
4619. (d) Bäckvall, J. E.; Nyström, J. E.; Nordberg R. E. J. Am. Chem. Soc.
1985, 107, 3676.
(a) Horvath, A.; Bäckvall, J. E. In Modern Allene Chemistry; Krause, N.;
Hashmi, A. S. K.; Eds; Wiley-VCH: Weinheim, 2004, 2, pp 973-994. (b)
Zimmer, R.; Dinesh, C. U.; Nandanan, E.; Klhan, F. A. Chem. Rev. 2000,
100, 3067.
Rogers, M. M.; Wendlandt, J. E.; Guzei, I. A.; Stahl, S. S. Org. Lett., 2006, 8,
2257.
(a) Popp, B. V.; Stahl, S. S. In Organometallic Oxidation Catalysis, Meyer,
F.; Limberg, C.; Eds; Springer: New York, 2007, 22, pp 149-189. (b)
Decharin, N.; Stahl, S. S. J. Am. Chem. Soc. 2011, 133, 5732.
Piera, J.; Bäckvall, J. E. Angew. Chem. Int. Ed. 2008, 47, 3506.
Michel, B.W.; Steffens, L. D.; Sigman, M. S. J. Am. Chem. Soc., 2011, 133,
8317.
(a) Campbell, A. N.; Stahl, S. S. Acc. Chem. Res. 2012, DOI:
10.1021/ar2002045. (b) Steinhoff, B. A.; Fix, S. R.; Stahl, S. S. J. Am. Chem.
Soc. 2002, 124, 766. (c) Decharin, N.; Popp, B. V.; Stahl, S. S. J. Am. Chem.
Soc. 2011, 133, 13268. (d) Konnick, M. M.; Decharin, N.; Popp, B. V.; Stahl,
S. S. Chem. Sci. 2011, 2, 326. (e) Popp, B. V.; Morales, C. M.; Landis, C. R.;
Stahl, S. S. Inorg. Chem. 2010, 49, 8200. (f) Popp, B. V.; Stahl, S. S. Chem.
Eur. J. 2009, 15, 2915. (g) Konnick, M. M.; Stahl, S. S. J. Am. Chem. Soc.,
2008, 130, 5753. (h) Landis, C. R.; Morales, C. M.; Stahl, S. S. J. Am. Chem.
Soc. 2004, 126, 16302.
(a) Anderson, B. J.; Keith, J. A.; Sigman, M. S. J. Am. Chem. Soc. 2010, 132,
11872. (b) Gligorich, K. M.; Sigman, M. S. Angew. Chem., Int. Ed. 2006, 45,
6612.
(a) Ebner, D. C.; Bagdanoff, J. T.; Ferreira, E. M.; McFadden, R. M.; Caspi,
D. D.; Trend, R. M.; Stoltz, B. M. Chem. Eur. J. 2009, 15, 12978. (b) Ebner,
83
(47)
(48)
(49)
(50)
(51)
(52)
(53)
(54)
(55)
(56)
(57)
(58)
(59)
(60)
(61)
84
D. C.; Trend, R. M.; Genet, C.; McGrath, M. J.; O'Brien, P.; Stoltz, B. M.
Angew. Chem. Int. Ed. 2008, 47, 6367. (c) Trend, R. M.; Ramtohul, Y. K.;
Stoltz, B. M. J. Am. Chem. Soc. 2005, 127, 17778.
(a) Keith, J. M.; Muller, R. P.; Kemp, R. A.; Goldberg, K. I.; Goddard, W.
A.; Oxgaard, J. Inorg. Chem. 2006, 45, 9631. (b) Keith, J. M.; Nielsen, R. J.;
Oxgaard, J.; Goddard, W. A. J. Am. Chem. Soc. 2005, 127, 13172. (c) Keith,
J. M.; Goddard, W. A. J. Am. Chem. Soc. 2009, 131, 1416.
Werner, E.; Sigman, M. S. J. Am. Chem. Soc. 2010, 132, 13981.
Bäckvall, J. E.; Awasthi, A. K.; Renko, Z. D. J. Am. Chem. Soc. 1987, 109,
4750.
(a) Grennberg, H.; Bäckvall, J. E. Acta Chem. Scand. 1993, 47, 506. (b)
Zsigmond, Á.; Nothieisz, F.; Bartok, M.; Bäckvall, J. E. Stud. Surf. Sci.
Catal. 1993, 78, 417.
(a) Wöltinger, J.; Bäckvall, J. E.; Zsigmond, A. Chem. Eur. J. 1999, 5, 1460.
(b) Johansson, M.; Purse, B. W.; Terasaki, O.; Bäckvall, J. E. Adv. Synth.
Catal. 2008, 350, 1807.
Piera, J.; Närhi, K.; Bäckvall, J. E. Angew. Chem. Int. Ed. 2006, 45, 6914.
(a) Purse, B.W.; Tran, L. H.; Piera, J.; Åkermark, B.; Bäckvall, J. E. Chem.
Eur. J. 2008, 14, 7500. (b) Johnston, E. V.; Karlsson, E. A.; Tran, H. L.;
Åkermark, B.; Bäckvall, J. E. Eur. J. Org. Chem. 2009, 3973.
Johnston, E. V.; Karlsson, E. A.; Lindberg, S. A.; Åkermark, B.; Bäckvall, J.
E. Chem. Eur. J. 2009, 15, 6799.
Grennberg, H.; Faizon, S.; Bäckvall, J. E. Angew. Chem. Int. Ed. 1993, 32,
263.
(a) Zsigmond, Á.; Notheisz, F.; Csjernik, G.; Bäckvall, J. E. Topics in
Catalysis, 2002, 19, 119.
Endo, Y.; Bäckvall, J. E. Chem. Eur. J. 2011, 17, 12596.
Trost, B. M. Chem. Pharm. Bull. 2002, 50, 1.
Balame, G.; Bouyssi, D.; Monteiro, N. A. In Handbook of Organopalladium
Chemistry for Organic Synthesis, Negishi, E. –i.; Eds; John Wiley & Sons:
New York, 2002, 2, pp 2245-2265.
(a) Cho, C. S.; Uemura, S. J. Organomet. Chem. 1994, 465, 85.
(b) Du, X.; Suguro, M.; Hirabayashi, K.; Mori, A. Org. Lett. 2001, 3, 3313.
(c) Enquist, P. A.; Lindh, J.; Nilsson, P.; Larhed, M. Green Chem. 2006, 8,
338. (d) Yoo, K. S.; Park, C. P.; Yoon, C. H.; Sakaguchi, S.; O’Neill, J.;
Jung, K. W. Org. Lett. 2007, 9, 3933 (e) Ruan, J.; Li, X.; Saidi, O.; Xiao, J.
J. Am. Chem. Soc. 2008, 130, 2424. (f) Lindh, J.; Enquist, P. A.; Pilotti, Å.;
Nilsson, P.; Larhed, M. J. Org. Chem. 2007, 72, 7957. (g) Delcamp, J. H.;
Brucks, A. P.; White, M. C. J. Am. Chem. Soc. 2008, 130, 11270.
(a) Liu, G.; Wu, Y. Top. Curr. Chem. 2010, 292, 195. (b) Jensen, T.; Fristrup,
P. Chem. Eur. J. 2009, 15, 9632.
(62)
(63)
(64)
(65)
(66)
(67)
(68)
(69)
(70)
(71)
(72)
(73)
(74)
(75)
(76)
(77)
(78)
(79)
(80)
(81)
(82)
(a) Lin, S.; Song, C. X.; Cai, G. X.; Wang, W. H.; Shi, Z. J. J. Am. Chem.
Soc. 2008, 130, 12901. (b) Young, A. J.; White, M. C. Angew. Chem. Int. Ed.
2011, 50, 6824.
Deng, Y.; Persson, A. K. Å.; Bäckvall, J. E. Review submitted for
publication.
Young, A. J.; White, C. M. J. Am. Chem. Soc. 2008, 130, 14090.
Kalyani, D.; Sanford, M. S. J. Am. Chem. Soc. 2008, 130, 2150.
(a) Oppolzer, W. Pure Appl. Chem. 1990, 62, 1941. (b) Beletskaya, I. P.;
Cheprakov, A. V. Chem. Rev., 2000, 100, 3009.
(a) Trost, B. M.; Lautens, M. J. Am. Chem. Soc. 1985, 107, 1781. (b) Trost,
B. M.; Krische, M. J. Synlett 1998, 1.
Mikami, K.; Hatano, M. PNAS, 2004, 101, 5767.
(a) Widenhoefer, R. A. Acc. Chem. Res. 2002, 35, 905. (b) Lloyd-Jones, G.
C. Org. Biomol. Chem. 2003, 1, 215.
(a) Ito, Y.; Aoyama, H.; Hirao, T.; Mochizuki, A.; Saegusa, T. J. Am. Chem.
Soc. 1979, 101, 494. (b) Pei, T.; Wang, X.; Widenhoefer, R. A. J. Am. Chem.
Soc. 2003, 125, 648.
Toyota, M.; Rudyanto, M.; Masataka, I. J. Org. Chem. 2002, 67, 3374.
Rönn, M.; Andersson, P. G.; Bäckvall, J. E. Tetrahedron Lett. 1997, 38,
3603.
(a) Wang, X.; Widenhoefer, R. A. Chem. Commun. 2004, 660. (b) Han, X.;
Wang, X.; Pei, T.; Widenhoefer, R. A. Chem. Eur. J. 2004, 10, 6343.
(a) Yip, K. T.; Li, J. H.; Lee, O. Y.; Yang, D. Org. Lett. 2005, 7, 5717. (b)
Kung, L. R.; Tu, C. H.; Shia, K. S.; Liu, H. J. Chem. Commun. 2003, 2490.
(c) Hibi, A.; Toyota, M. Tetrahedron Lett. 2009, 50, 4888.
(a) Ferreira, E. M.; Stoltz, B. M. J. Am. Chem. Soc. 2003, 125, 9578.
(a) Schiffner, J. A.; Machotta, A. B.; Oestreich, M. Synlett 2008, 2271. (b)
Beccalli, E. M.; Broggini, G. Tetrahedron Lett. 2003, 44, 1919. (c) Beck, M.;
Hatley, R.; Gaunt, M. J. Angew. Chem. Int. Ed. 2008, 47, 3004.
Tong, X.; Beller, M.; Tse, M. K. J. Am. Chem. Soc. 2007, 129, 4906.
(a) Castano, A. M.; Persson, B. A.; Bäckvall, J. E. Chem. Eur. J. 1997, 3,
482. (b) Castano, A. M.; Bäckvall, J. E. J. Am. Chem. Soc. 1995, 117, 560.
Franzén, J.; Bäckvall, J. E. J. Am. Chem. Soc. 2003, 125, 6056.
(a) Löfstedt, J.; Franzén, J.; Bäckvall, J. E. J. Org. Chem. 2001, 66, 8015. (b)
Dorange, I.; Löfstedt, J.; Närhi, K.; Franzén, J.; Bäckvall, J. E. Chem. Eur. J.
2003, 9, 3445. (c) Löfstedt, J.; Närhi, K.; Dorange, I.; Bäckvall, J. E. J. Org.
Chem. 2003, 68, 7243.
Horváth, A.; Bäckvall, J. E. In Modern Allene Chemistry 2nd ed.; Krause, N.;
Hashmi, A. S. K.; Eds; Wiley VCH-Verlag: Weinheim,, 2004, 2, pp 973994.
E. M. Becalli, G. Broggini, M. Martinelli, S. Sotocarnola, Chem. Rev. 2007,
107, 5318.
85
(83)
(84)
(85)
(86)
(87)
(88)
(89)
(90)
(91)
(92)
(93)
(94)
(95)
(96)
(97)
(98)
(99)
(100)
(101)
(102)
(103)
(104)
(105)
86
Hashmi, A. S. K. In Modern Allene Chemistry 2nd ed.; Krause, N.; Hashmi,
A. S. K.; Eds; Wiley VCH-Verlag: Weinheim, 2004, 2, pp 877-923
(a) Busacca, C. A.; Dong, Y. Tetrahedron Lett. 1996, 37, 3947, (b) Lei, A.;
Lu, X. Org. Lett. 2000, 2, 2357.
Bäckvall, J. E.; Nilsson, Y. L. M.; Andersson, P. G.; Gatti, R. G. P.; Wu, J.
Tetrahedron Lett. 1994, 35, 5713.
Nilsson, Y. I. M.; Aranyos, A.; Andersson, P. G.; Bäckvall, J. E.; Parrain, J.
L.; Ploteau, C.; Quintard, J. P. J. Org. Chem. 1996, 61, 1825.
Karlsson, E. A.; Bäckvall, J. E. Chem. Eur. J. 2009, 14, 9175.
Hartwig, J. F.; Nature 2008, 455, 314.
(a) Wagaw, S.; Yang, B.; Buchwald, S. L. J. Am. Chem. Soc. 1999, 121,
10251. (b) Paul, F.; Patt, J.; Hartwig, J. F. J. Am. Chem. Soc. 1994, 116,
5969, (c) Guram, A. S.; Buchwald, S. L. J. Am. Chem. Soc. 1994, 116, 7901.
Kantchev, E. A. B.; O’Brien, C. J.; Organ, M. G. Angew. Chem. Int. Ed.
2007, 46, 2768.
Ikawa, T.; Barder, T. E.; Biscoe, M. R.; Buchwald, S. L. J. Am. Chem. Soc.
2007, 129 43, 13001.
(a) Ley, S. V.; Thomas, A. W. Angew. Chem. Int. Ed. 2003, 42, 5400. (b)
Hassan, J.; Sevignon, M.; Gozzi, C.; Schulz, E.; Lemaire, M. Chem. Rev.
2002, 102, 1359. (c) Ma, D.; Cai, Q. Acc. Chem. Res. 2008, 41, 1450.
Ullmann, F.; Bielecki, J. Chem. Ber. 1901, 34, 2174.
Zim, D.; Buchwald, S. L. Org. Lett., 2003, 5, 2413.
Lu, H.; Li, C. Org. Lett., 2006, 8, 5365.
Stiles, D. T.; Trost, B. M. Org. Lett., 2005, 7, 2117.
Shen, L.; Hsung, R. P.; Zhang, Y.; Antoline, J. E.; Zhang, X. Org. Lett. 2005,
7, 3081.
(a) Armstrong, A.; Emmerson, D. P. G. Org. Lett. 2009, 11, 1547. (b) Wei,
L.; Xiong, H.; Hsung, R. P. Acc. Chem. Res. 2003, 36, 773.
Hashmi A. S. K. In Modern Allene Chemistry, 2nd ed.; Krause, N.; Hashmi,
A. S. K.; Eds; Wiley VCH-Verlag: Weinheim, 2004, vol. 1, pp 21-36.
Närhi, K.; Franzén, J.; Bäckvall, J. E. J. Org. Chem. 2006, 71, 2914.
(a) Alcaide, B.; Almendros, P.; Aragoncillo, C. Chem. Soc. Rev. 2010, 39,
783. (b)Wender, P. A.; Zhang, L. Org. Lett. 2000, 2, 2323.
(a) Strieter, E. R.; Bhayana, B.; Buchwald, S. L. J. Am. Chem. Soc. 2009,
131, 78. (b) Sperotto, A.; van Klink, G. P. M.; van Koten, G. V.; de Vries, J
.G. Dalton Trans., 2010, 39, 10338.
Joule, J. A.; Mills, K. Heterocyclic Chemistry, 4th ed. Blackwell Science ltd.
Malden, MA, 2000, pp 237-272.
Felpin, F. X.; Lebreton, J. Eur. J. Org. Chem, 2003, 3693.
Franzén, J.; Löfstedt, J.; Dorange, I.; Bäckvall, J. E. J. Am. Chem. Soc. 2002,
124, 11246.
(106) (a) Stoyanov, E. S. J. Struct. Chem. 2000, 41, 440. (b) Bakhmutov, V. I.;
Berry; J. F.; Cotton; F. A.; Ibragimov; S.; Murillo, C. A. Dalton Trans. 2005,
1989.
(107) Eastgate, M. D.; Buono, F. G. Angew. Chem. Int. Ed. 2009, 48, 5958.
(108) Persson, A. K. Å.; Bäckvall, J. E. Unpublsihed results.
(109) Enquist, J. A.; Stoltz, B. M. Nature 2008, 453, 1228
(110) (a) Marco-Martinez, J.; Buñuel, E.; López-Durán, R.; Cárdenas, D. J. Chem.
Eur. J. 2011, 17, 2734. (b) Pardo-Rodríguez, V.; Marco-Martínez, J.; Buñuel,
E.; Cárdenas, D. J. Org. Lett. 2009, 11, 4548. (c) Marco-Martínez, J.; Buñuel,
E.; Muñoz-Rodríguez, R.; Cárdenas, D. J. Org. Lett. 2008, 10, 3619. (d)
Marco-Martínez, J.; López-Carrillo, V.; Buñuel, E.; Simancas, R.; Cárdenas,
D. J. J. Am. Chem. Soc. 2007, 129, 1874.
(111) Preliminary results using PhI(OAc)2 (PIDA) as oxidant gives rise to
unidentified acetoxylated carbocycles, as determined by HRMS-analysis.
(112) Burks, H. E.; Liu, S.; Morken, J. P. J. Am. Chem. Soc., 2007, 129, 8766.
(113) Crystal data: C23H30BO6, M = 413.28, monoclinic, a = 13.7804(17), b =
10.0315(10), c =17.5590(19) Å, β = 100.84(1), V = 2384.0(3) Å3, space
group P21/c, Z = 4, 14193 reflections measured, 4164 unique, (Rint = 0.0386)
which were used in all calculations. The final wR(F2) was 0.2118 (all data)
and the R(F) was 0.0640 (I > 2σ(I)) using 271 parameters.
(114) Ishiyama, T.; Murata, M.; Miyaura, N. J. Org. Chem. 1995, 60, 7508.
(115) Bergmeier, S. C. Tetrahedron 2000, 56, 2561.
(116) Zappia, G.; Gacs-Baitz, E.; Delle Monache, G.; Misiti, D.; Nevola, L.; Botta,
B. Curr. Org. Synth, 2007, 4, 81.
(117) Harayama, H.; Abe, A.; Sakado, T.; Kimura, M.; Fugami, K.; Tanaka, S.;
Tamaru, Y. J. Org. Chem., 1997, 62, 2113.
(118) Alexanian, E. J.; Lee C.; Sorensen, E. J. J. Am. Chem. Soc. 2005, 127, 7690.
(119) Nicolai, S.; Piemontesi, C.; Waser, J. Angew. Chem., Int. Ed. 2011, 50, 468.
(120) Lei, A.; Lu, X. Org. Lett., 2000, 2, 2699.
(121) (a) Liu, G.; Lu, X. Org. Lett., 2001, 3, 3879. (b) Jonasson, C.; Karstens, W.
F. J.; Hiemstra, H.; Bäckvall, J. E. Tetrahedron Lett. 2000, 41, 1619.
(122) Tamuru, Y.; Kimura, M. Synlett 1997, 7, 749.
(123) (a) Fraunhoffer, K. J.; White, M. C. J. Am. Chem. Soc. 2007, 129, 7274. (b)
Qi, X.; Rice, G. T.; Lall, M. S.; Plummer, M. S.; White, M. C. Tetrahedron
2010, 66, 4816.
(124) (a) Anderson, C. E.; Overman, L. E. J. Am. Chem. Soc., 2003, 125, 12412 (b)
Overman, L. E. J. Am. Chem. Soc. 1976, 98, 2901. (c): Overman, L. E. Acc.
Chem. Res. 1980, 13, 218.
(125) Weiner, B.; Baeza, A.; Jerphagnon, T.; Feringa, B. L. J. Am. Chem. Soc.
2009, 131, 9473.
(126) Moloney, M. G. Nat. Prod. Rep. 2002, 19, 597.
(127) Jiang, C.; Covell, D. J; Stepan, A. F.; Plummer, M. S.; White, C. M. Org.
Lett. 2012, 14, 1386.
87
(128) Green, T. W.; Wuts, P. G. M. Protective Groups in Organic Synthesis 3rd ed,
Wiley-Interscience; John Wiley Sons, New York, 1999, pp 603-607.
(129) Nyasse, B.; Grehn, L.; Ragnarsson, U. Chem. Commun., 1997, 1017.
(130) Crystal data: C14H15NO4S, M = 293.33, monoclinic, a = 9.7314(7), b =
12.3012(8), c =11.5157(8) Å, β = 97.096(6), V = 1367.96(17) Å3, space
group P21/c, Z = 4, 13234 reflections measured, 4823 unique, (Rint = 0.0327)
which were used in all calculations. The final wR(F2) was 0.1600 (all data)
and the R(F) was 0.0512 (I > 2σ(I)) using 181 parameters.
(131) Crystal data: C17H23NO4S, M = 337.43, monoclinic, a = 12.0570(17), b =
7.1943(12), c =20.571(3) Å, β = 101.221(13), V = 1750.3(4) Å3, space group
P21/c, Z = 4, 9282 reflections measured, 4609 unique, (Rint = 0.0227) which
were used in all calculations. The final wR(F2) was 0.1163 (all data) and the
R(F) was 0.0418 (I > 2σ(I)) using 208 parameters.
(132) Andersson, P. G. J. Org. Chem. 1996, 61, 4154.
(133) (a) Åkermark, B.; Bäckvall, J. E.; Siirala-Hansén, K.; Sjöberg, K. J.;
Zetterberg, K. Tetrahedron Lett. 1974, 15, 1363. (b) Åkermark, B.; Bäckvall,
J.-E.; Hegedus, L. S.; Zetterberg, K.; Siirala-Hansén, K.; Sjöberg, K. J.
Organomet. Chem. 1974, 72, 127. (c) Hegedus, L. S.; Allen, G. F.;
Waterman, E. L. J. Am. Chem. Soc. 1976, 98, 2674. (d) Bäckvall, J. E. Acc.
Chem. Res. 1983, 16, 335. (e) Bäckvall, J. E.; Björkman, E. E. Acta Chem.
Scand. B 1984, 38, 91.
(134) (a) Liu, G.; Stahl, S. S. J. Am. Chem. Soc. 2006, 128, 7179. (b) McDonald, R.
I.; White, P. B.; Weinstein, A. B.; Tam, C. P.; Stahl, S. S. Org. Lett. 2011,
13, 2830
(135) Sibbald, P. A.; Rosewall, C. F.; Swartz, R. D.; and Michaels F. E. J. Am.
Chem. Soc. 2009, 131, 15945.
(136) Liu, G.; Stahl, S. S. J. Am. Chem. Soc. 2007, 129, 6328.
88