07.06.2013 Views

Historical Perspective of the Heck Reaction

Historical Perspective of the Heck Reaction

Historical Perspective of the Heck Reaction

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

Dimitra Kovala-Demertzi<br />

Section <strong>of</strong> Inorganic and Analytical Chemistry, Department <strong>of</strong> Chemistry, University <strong>of</strong><br />

Ioannina, 45110 Ioannina, e-mail:dkovala@cc.uoi.gr<br />

<strong>Historical</strong> <strong>Perspective</strong> <strong>of</strong> <strong>the</strong> <strong>Heck</strong> <strong>Reaction</strong><br />

The first intermolecular <strong>Heck</strong> reaction was reported by <strong>Heck</strong> in 1972<br />

Development <strong>of</strong> <strong>the</strong> general intermolecular reaction suffered due to poor<br />

regiocontrol <strong>of</strong> <strong>the</strong> addition and elimination steps for electronically neutral<br />

unsymmetrical olefins<br />

Nolley, J.P.; <strong>Heck</strong>, R.F.; Tetrahedron 1972, 37, 2320<br />

The first intramolecular <strong>Heck</strong> reaction was reported by Mori and Ban in<br />

1977<br />

Mori, M.; Ban, K.; Tetrahedron 1977, 12, 1037<br />

1


General Catalysis <strong>of</strong> <strong>the</strong> <strong>Heck</strong> <strong>Reaction</strong><br />

Cycle is catalytic in palladium with <strong>the</strong> addition <strong>of</strong> stoichiometric base to<br />

scavenge HX<br />

Palladium catalysed carbon–carbon and carbon–heteroatom bond forming reactions<br />

are widely used and powerful tools in organic syn<strong>the</strong>sis. Such processes are typified<br />

by <strong>the</strong> <strong>Heck</strong> reaction (Schem e 1) and cross-coupling and related reactions where an<br />

aryl halide is coupled with a nucleophilic partner. The <strong>Heck</strong> reaction (also called <strong>the</strong><br />

M izoroki-H eck reaction) is <strong>the</strong> chem ical reaction <strong>of</strong> an unsaturated halide (or<br />

triflate) w ith an alkene and a strong base and palladium catalyst to form a substituted<br />

alkene. It is nam ed after <strong>the</strong> Am erican chem ist Richard F. <strong>Heck</strong> [1-4]<br />

The halide or triflate is an aryl, benzyl, or vinyl compound and <strong>the</strong> alkene contains at<br />

least one proton and is <strong>of</strong>ten electron-deficient such as acrylate ester or<br />

anacrylonitrile.<br />

Schem e 1. The <strong>Heck</strong> coupling reaction<br />

2


Aryl and vinyl halides- Aryl and vinyl chlorides are most reluctant to undergo Pdcatalyzed<br />

activa-tion. <strong>Heck</strong> reactivity - as expected from <strong>the</strong> C-X bond dissociation<br />

energies (Figure 1) - increases in <strong>the</strong> order CI « Br < I, with fluorides being<br />

com-pletely unreactive with any <strong>of</strong> <strong>the</strong> known catalysts.<br />

The ideal substrates for coupling reactions are aryl chlorides since <strong>the</strong>y tend to be<br />

cheaper and more widely available than <strong>the</strong>ir bromide or iodide counterparts.<br />

Unfortunately <strong>the</strong> high C-Cl bond strength compared with C-Br and C-I bonds<br />

disfavours oxidative addition, <strong>the</strong> first step in catalytic coupling reactions, making <strong>the</strong><br />

coupling <strong>of</strong> such substrates far more challenging. Therefore, <strong>the</strong>re is currently much<br />

interest in <strong>the</strong> syn<strong>the</strong>sis <strong>of</strong> catalysts that are able to activate aryl chloride substrates at<br />

ever lower catalyst loading<br />

The activation <strong>of</strong> chlorohydrocarbons is <strong>of</strong> major industrial interest. Progress was<br />

made by introducing a bimetallic Ni/Pd catalyst which converts <strong>the</strong> aryl-CI in a first<br />

step into <strong>the</strong> more reactive aryl-Br bond (NiBr2)' followed by Pd-catalyzed C-Ccoupling<br />

(Pd(dbab)2 dba = dibenzylidene-acetone, (C6H5CH=CH)2C=O).<br />

).<br />

Figure 1. C-X bond dissociation energies (X = F, CI, Hr, I) <strong>of</strong> aryl halides. 1 kcal<br />

mol -1 = 4,184 kJ mol -1 .<br />

3


Catalysts-The reaction is performed in <strong>the</strong> presence <strong>of</strong> an organopalladium catalyst.<br />

Palladium is one <strong>of</strong> <strong>the</strong> most versatile and efficient catalyst metals in organic syn<strong>the</strong>sis,<br />

be it in elemental forms (palladium black and palladium colloids in heterogeneous<br />

hydrogenation) or as palladium salts and complexes. Both <strong>the</strong> renaissance <strong>of</strong><br />

organometallic chemistry in <strong>the</strong> 1960s and <strong>the</strong> subsequent breakthrough <strong>of</strong><br />

homogeneous organometallic catalysis in laboratory-scale and industrial syn<strong>the</strong>ses have<br />

received a major stimulus from palladium coordination chemistry.<br />

Typical catalysts are Pd°-phosphine complexes. The catalyst can be<br />

tetrakis(triphenylphosphine)palladium(0) {Pd[P(C6H5)3]4}, palladium chloride or , or in-situ<br />

catalysts such as palladium(II) acetate Pd(OAc)2 / n P(C6H5)3, with n = 2...4 (OAc =<br />

acetate). The most frequently used catalyst is an in-situ combination <strong>of</strong> Pd(OAc)2 and<br />

P(C6H5)3. Palla-dium is practically <strong>the</strong> only catalyst metal used, in <strong>the</strong> form <strong>of</strong> certain<br />

Pd° and salts or complexes; normally 1-5 mol% <strong>of</strong> catalyst is administered.<br />

Unfortunately, whilst useful <strong>the</strong>se ‘classical’ catalyst systems suffer from two major<br />

limitations. Generally, <strong>the</strong>y need to be used in high loadings — typically a few mol%<br />

Pd and <strong>the</strong>y show little or no activity with aryl chloride substrates. For a reaction to be<br />

attractive for application in <strong>the</strong> industrial sector, such as in <strong>the</strong> fine chemical or<br />

pharmaceutical industries, <strong>the</strong>n palladium contamination <strong>of</strong> <strong>the</strong> product must be in <strong>the</strong><br />

low ppm region, <strong>of</strong>ten necessitating expensive product clean-up. This, coupled with <strong>the</strong><br />

high price <strong>of</strong> not only <strong>the</strong> palladium but <strong>of</strong>ten <strong>the</strong> ligands, can make <strong>the</strong> whole process<br />

prohibitively expensive<br />

Ligands-The ligand is triphenylphosphine or BINAP. <strong>Heck</strong> and Spencer noticed<br />

that phosphines are necessary to somehow "stabilize" <strong>the</strong> catalysts. Phosphine<br />

ligands are expensive, toxic, and unrecoverable. In large-scale applications on<br />

industrial and semi-industrial scale, <strong>the</strong> phosphines might be a more serious<br />

economical burden than even palladium itself, which can be recovered at any<br />

stage <strong>of</strong> production or from wastes. The chemical reason is lower reactivity <strong>of</strong><br />

fully ligated complexes <strong>of</strong> palladium, <strong>the</strong> main result <strong>of</strong> which is <strong>the</strong> need for<br />

higher loads <strong>of</strong> catalyst to achieve appropriate rates <strong>of</strong> reaction and <strong>the</strong>refore<br />

fur<strong>the</strong>r aggravation <strong>of</strong> procedure cost.<br />

Bases-The base is triethylamine (e.g., N(C2H5)3), potassium carbonate<br />

K2CO3,or sodium acetate NaOAc<br />

4


Solvents-<strong>Heck</strong> reactions are conducted in polar aprotic, σ-donor-type<br />

solvents such as acetonitrile, dimethyl sulfoxide, or dimethylacetamide.<br />

<strong>Reaction</strong> temperatures and times largely depend on <strong>the</strong> nature <strong>of</strong> <strong>the</strong><br />

organic halide to be activated and on <strong>the</strong> catalyst's stability limit. Iodo<br />

derivatives are much more reactive (


<strong>Reaction</strong> mechanism<br />

The catalytic cycle for <strong>the</strong> <strong>Heck</strong> reaction involves a series <strong>of</strong> transformations around <strong>the</strong><br />

palladium catalyst. The palladium(0) compound required in this cycle is generally<br />

prepared in situ from a palladium(II) precursor [3]. For instance, palladium(II) acetate is<br />

reduced by triphenylphosphine to di(triphenylphosphine)palladium(0) and<br />

triphenylphosphine is oxidized to triphenylphosphine oxide in step 1.<br />

Step 2 is an oxidative addition in which palladium inserts itself in <strong>the</strong> aryl to bromide<br />

bond.<br />

In step 3, palladium forms a π complex with <strong>the</strong> alkene and in step 4 <strong>the</strong> alkene inserts<br />

itself in <strong>the</strong> palladium - carbon bond in a syn addition step.<br />

Step 5 is a torsional strain relieving rotation and step 6 is a Beta-hydride elimination<br />

step with <strong>the</strong> formation <strong>of</strong> a new palladium - alkene π complex.<br />

This complex is destroyed in step 7. T<br />

he palladium(0) compound is regenerated by reductive elimination <strong>of</strong> <strong>the</strong> palladium(II)<br />

compound by potassium carbonate in <strong>the</strong> final step 8.<br />

In <strong>the</strong> course <strong>of</strong> <strong>the</strong> reaction <strong>the</strong> carbonate is stoichiometrically consumed and<br />

palladium is truly a catalyst and used in catalytic amounts.<br />

6


Oxidative addition is <strong>of</strong>ten <strong>the</strong> rate-determining step in a catalytic cycle.<br />

The relative reactivity decreases in <strong>the</strong> order <strong>of</strong> I > OTf > Br >> C1. Aryl<br />

and 1-alkenyl halides activated by <strong>the</strong> proximity <strong>of</strong> electron-withdrawing<br />

groups are more reactive to <strong>the</strong> oxidative addition than those with donating<br />

groups, thus allowing <strong>the</strong> use <strong>of</strong> chlorides such as 3-chloroenone for <strong>the</strong><br />

cross-coupling reaction<br />

Reductive elimination <strong>of</strong> organic partners reproduces <strong>the</strong> palladium(0)<br />

complex. The reaction takes place directly from cis-complex, and <strong>the</strong> transcomplex<br />

reacts after its isomerization to <strong>the</strong> corresponding cis-complex (eqs<br />

1 and 2). The order <strong>of</strong> reactivity is diaryl- > (alky1)aryl- > dipropyl- ><br />

diethyl- > dimethylpalladium(II), suggesting participation by <strong>the</strong> n-orbital<br />

<strong>of</strong> aryl group during <strong>the</strong> bond formation (eq 1) [3].<br />

• A general catalytic cycle for cross-coupling <strong>Heck</strong> reaction<br />

7


Ionic liquid <strong>Heck</strong> reaction<br />

In <strong>the</strong> presence <strong>of</strong> an ionic liquid a <strong>Heck</strong> reaction proceeds in absence <strong>of</strong> a<br />

phosphorus ligand. In one modification palladium acetate and <strong>the</strong> ionic liquid<br />

(bmim)PF6 are immobilized inside <strong>the</strong> cavities <strong>of</strong> reversed-phase silica gel<br />

[4a]. In this way <strong>the</strong> reaction proceeds in water and <strong>the</strong> catalyst is re-usable.<br />

<strong>Heck</strong> oxyarylation<br />

In <strong>the</strong> <strong>Heck</strong> oxyarylation modification <strong>the</strong> palladium substituent in <strong>the</strong> synaddition<br />

intermediate is displaced by a hydroxyl group and <strong>the</strong> reaction product<br />

contains a tetrahydr<strong>of</strong>uran ring.[4b]<br />

8


Amino-<strong>Heck</strong> reaction<br />

In <strong>the</strong> amino-<strong>Heck</strong> reaction a nitrogen to carbon bond is formed. In one example,[4c]<br />

an oxime with a strongly electron withdrawing group reacts intramolecularly with <strong>the</strong><br />

terminal end <strong>of</strong> a diene to a pyridine compound. The catalyst is<br />

tetrakis(triphenylphosphine)palladium(0) and <strong>the</strong> base is triethylamine<br />

3. Recent Developments on <strong>Heck</strong> reaction<br />

3.1. Palladacycles: Efficient Catalyst Precursors for Homogeneous Catalysis for<br />

<strong>Heck</strong> reaction<br />

Palladacycles are a popular and thoroughly investigated class <strong>of</strong> organopalladium<br />

compounds. The vast majority <strong>of</strong> <strong>the</strong>se complexes possess anionic four-electron<br />

(bidentate) or six-electron (tridentate) donor ligands, with five-membered nitrogencontaining<br />

rings being <strong>the</strong> most common. Such systems have been known since <strong>the</strong><br />

1960s [6].<br />

Examples <strong>of</strong> palladacycles<br />

Their syn<strong>the</strong>sis is facile and it is possible to modulate <strong>the</strong>ir electronic and steric properties simply<br />

by changing (i) <strong>the</strong> size <strong>of</strong> <strong>the</strong> metallacyclic ring (ii) <strong>the</strong> nature <strong>of</strong> <strong>the</strong> metallated carbon atom<br />

(aliphatic, aromatic, vinylic, etc.), (iii) <strong>the</strong> type <strong>of</strong> donor group (N-, P-, S-, O containing group,<br />

etc.) and its substituents (alkyl, aryl, etc.), or (iv) <strong>the</strong> nature <strong>of</strong> <strong>the</strong> X ligands (halide, triflate, or<br />

solvent, e.g. THF, H2O). These factors determine whe<strong>the</strong>r <strong>the</strong> complex is dimeric, monomeric,<br />

neutral, or cationic.However, as already mentioned, palladacycles <strong>of</strong>fer a much wider variety <strong>of</strong><br />

catalytic applications and several new and efficient complexes based on different ligand<br />

backbones have since been reported<br />

9


Phosphorus-, nitrogen-, and sulfur-containing palladacycles are among <strong>the</strong> most active<br />

catalyst precursors for <strong>the</strong> promotion <strong>of</strong> such reactions reported to date. All <strong>of</strong> <strong>the</strong><br />

palladacycles shown in this scheme promote <strong>the</strong> <strong>Heck</strong> reaction <strong>of</strong> aryl iodides with<br />

acrylic esters (Table 1).<br />

10


All <strong>of</strong> <strong>the</strong> palladacycles shown in <strong>the</strong> Scheme promote <strong>the</strong> <strong>Heck</strong> reaction <strong>of</strong> aryl<br />

iodides with acrylic esters (Table 1). In this reaction, <strong>the</strong> highest catalytic activity<br />

observed to date was achieved with <strong>the</strong> palladacycle 6 (Table 1, Entry 9). It is<br />

interesting to note that this palladacycle only promotes <strong>the</strong> <strong>Heck</strong> reaction between aryl<br />

iodides and acrylic esters.<br />

Only a few palladacycles are active catalyst precursors for electron-rich aryl bromides<br />

(Table 1). Moreover, in <strong>the</strong> case <strong>of</strong> aryl chlorides, complexes 1, 7, 8, and 11 were<br />

found to be <strong>the</strong> only active palladacycles (Entries 24-30, Table 1).<br />

In particular, 7 promotes <strong>the</strong> <strong>Heck</strong> reaction <strong>of</strong> electron-rich aryl chlorides, such as 4-<br />

chloroanisole (Entry 29, Table 1).<br />

In contrast with <strong>the</strong> o<strong>the</strong>r catalyst systems presented in Table 1, <strong>the</strong> PCP-pincer complex 7 does<br />

not require <strong>the</strong> use <strong>of</strong> additives to show good activity, even with electronically, deactivated aryl<br />

chlorides. This may imply that here <strong>the</strong> active catalyst is not colloidal palladium but ra<strong>the</strong>r is a<br />

well defined molecular species, and currently it is not possible to rule out a Pd(II)/Pd(IV)<br />

catalytic manifold. It was proposed a catalytic cycle that proceeds via <strong>the</strong> C- H activation <strong>of</strong> <strong>the</strong><br />

olefin at a Pd(II) centre(s) followed by oxidative addition.<br />

Complex 7<br />

Proposed mechanism for <strong>Heck</strong><br />

coupling catalysed by complex<br />

7.<br />

11


3.2. Palladium coordination compounds as Catalysts for <strong>the</strong> <strong>Heck</strong> reaction<br />

The complexes used as catalysts are usually based on phosphorus ligands. These<br />

catalysts are <strong>of</strong>ten water- and air-sensitive.<br />

Therefore, catalysis under phosphane-free conditions is a challenge <strong>of</strong> high importance,<br />

and a number <strong>of</strong> phosphane-free ligands as well as ligand-free palladium catalysts for<br />

<strong>the</strong> <strong>Heck</strong> reaction have been reported up to now.<br />

We have tried to evaluate phosphane-free systems in <strong>the</strong> <strong>Heck</strong> reaction, a substituted<br />

salicylaldehyde thiosemicarbazone was chosen for this purpose.<br />

The chemistry <strong>of</strong> thiosemicarbazones has been an extremely active area <strong>of</strong> research<br />

primarily because <strong>of</strong> <strong>the</strong> beneficial biological (viz. antiviral and antitumor) activities <strong>of</strong><br />

<strong>the</strong>ir transition-metal complexes. Salicylaldehyde thiosemicarbazone is a multidentate<br />

ligand with five potential coordination sites: three N, one O and one S atoms. Usually, it<br />

is bonded to a transition-metal leaving some potential donor sites unused, and it could<br />

be play a constructive role for specific purposes, e.g. <strong>the</strong> construction <strong>of</strong><br />

heteropolynuclear complexes.<br />

This phosphane-free system attracted our attention due to <strong>the</strong> presence <strong>of</strong> additional<br />

potential N-donors, since it is known that an additional coordination site as stabilizing<br />

group during <strong>the</strong> course <strong>of</strong> a metal-mediated reaction could improve <strong>the</strong> catalytic<br />

efficiency <strong>of</strong> <strong>the</strong> complex.<br />

The syn<strong>the</strong>sis <strong>of</strong> <strong>the</strong> palladium complex 3 is outlined in Scheme 1. 2-Salicylaldehyde-N(4)ethylthiosemicarbazone<br />

(H2Sal4Et) (2) was prepared by treatment <strong>of</strong> salicylaldehyde (1) with<br />

N-ethylthiosemicarbazide in ethanol. The syn<strong>the</strong>sis <strong>of</strong> complex 3 was achieved by <strong>the</strong> reaction<br />

<strong>of</strong> ligand 2 with Li2PdCl4, prepared in situ from PdCl2 and LiCl. The microanalytical data are<br />

consistent with <strong>the</strong> formula C10H14ClN3O2PdS which indicates <strong>the</strong> structure<br />

[Pd(HSal4Et)Cl] H2O<br />

12


The monoanionic HSal4Et ligand is coordinated to palladium in a tridentate fashion via <strong>the</strong><br />

phenoxy oxygen, <strong>the</strong> azomethine nitrogen N(1) and <strong>the</strong> sulfur atom, forming one six- and one fivemembered<br />

chelate rings. The ligand shows a Z, E, Z configuration for <strong>the</strong> donor centres oxygen,<br />

nitrogen and sulfur, respectively. The S–C(9) bond distance <strong>of</strong> 1.713(3) Å is consistent with a<br />

double-bond character, while both thioamide C–N distances (N(2)-C(9), 1.338(4) Å; N(3)-C(9),<br />

1.331(4) Å) indicate an increased single bond character, in accordance with a molecule protonated<br />

on N(2).15 The displacement from coplanarity is indicated by <strong>the</strong> dihedral angle between <strong>the</strong><br />

phenol ring, and <strong>the</strong> plane defined by <strong>the</strong> five-membered chelate ring Pd-S-C(9)-N(2)-N(1) being<br />

3.45(12)º, and <strong>the</strong> dihedral angle between <strong>the</strong> phenol ring and <strong>the</strong> plane defined by <strong>the</strong> sixmembered<br />

chelate ring Pd-N(1)-C(8)-C(7)-C(2)-O(1) being 2.21(13)º.<br />

Complex 3 was applied to <strong>the</strong> <strong>Heck</strong> reaction <strong>of</strong> styrene with some representative 4-substituted<br />

aryl bromides (from electron-rich to electron-poor) in DMF at 150ºC for 24 h, using AcONa as<br />

base, and in <strong>the</strong> absence <strong>of</strong> any promoting additive (Scheme 2, Table 1).<br />

The reaction was first performed using a 1:1000 catalyst:aryl bromide molar ratio to ensure a<br />

higher yield process, and <strong>the</strong>n by decreasing <strong>the</strong> ratio to 1:100000. There was good selectivity<br />

towards trans-stilbenes 6, ranging from 92.0 – 96.4%, and it is noteworthy that side-products<br />

were absent or present only in traces. As expected, <strong>the</strong> catalytic activity depends on <strong>the</strong> halide,<br />

while electron-withdrawing groups on <strong>the</strong> aryl ring increase <strong>the</strong> reaction rate. The activity<br />

follows in <strong>the</strong> order NO2 > CHO > H > OMe, suggesting that <strong>the</strong> rate-determining step in <strong>the</strong><br />

<strong>Heck</strong> reaction is <strong>the</strong> oxidative addition <strong>of</strong> <strong>the</strong> aryl bromide to <strong>the</strong> palladium catalyst. Under<br />

argon, <strong>the</strong> catalyst is <strong>the</strong>rmally stable and <strong>the</strong> reaction mixture retains yellow or colorless,<br />

depending <strong>of</strong> <strong>the</strong> palladium concentration. A catalyst:substrate ratio <strong>of</strong> 1:1000 leads to total<br />

yields <strong>of</strong> 45.8 and 53.9% for <strong>the</strong> very inactive 4-bromoanisole and <strong>the</strong> relatively inactive<br />

bromobenzene, respectively. At lower conversions, for <strong>the</strong>se substrates <strong>the</strong> reaction proceeds<br />

with TONs up to 17100 and 18000, respectively. High activity was observed for <strong>the</strong> activated 4bromobenzaldehyde<br />

and 1-bromo-4-nitrobenzene. A catalyst:substrate ratio <strong>of</strong> 1:1000 leads to<br />

high conversion (95%) <strong>of</strong> <strong>the</strong> aryl bromide, and with a ratio <strong>of</strong> 1:100000, <strong>the</strong> reaction proceeds<br />

with TONs up to 42700.<br />

13


Table 1. <strong>Heck</strong> reaction <strong>of</strong> aryl bromides with styrene catalysed by palladium complex 3<br />

aTotal GC yield <strong>of</strong> all isomers, based on <strong>the</strong> aryl bromide using decane as internal standard.<br />

bTurnover no. (TON) = fraction <strong>of</strong> products (6 + 7 + 8) × substrate/Pd ratio.<br />

performed under argon. However, in air and at a very low palladium concentration, for a<br />

catalyst:4-bromobenzaldehyde ratio <strong>of</strong> 1:100000, <strong>the</strong> yield was diminished.<br />

This phosphorus-free complex with additional potential N-donors is <strong>the</strong>rmally stable<br />

under argon, and efficiently catalyses <strong>the</strong> <strong>Heck</strong> reaction <strong>of</strong> aryl bromides with<br />

styrene, with good turnover numbers and a good selectivity towards trans-stilbenes.<br />

This system efficiently catalyses <strong>the</strong> <strong>Heck</strong> reaction <strong>of</strong> aryl bromides (from electronrich<br />

to electron-poor) with styrene under argon, with turnover numbers <strong>of</strong> up to<br />

42700, at 150ºC after 24 h, and with a selectivity towards trans-stilbenes ranging<br />

from 92.0 to 96.4 %. In air, for activated aryl bromides and for a palladium<br />

concentration <strong>of</strong> 1 mM, <strong>the</strong> yields are essential <strong>the</strong> same to those obtained when <strong>the</strong><br />

reaction performed under argon.<br />

These phosphine-free catalysts <strong>of</strong>fer <strong>the</strong> advantage <strong>of</strong> <strong>the</strong> successful coupling <strong>of</strong> aryl<br />

halides and <strong>the</strong> syn<strong>the</strong>sis <strong>of</strong> biaryls under aerobic conditions [7].<br />

14


3.3 Palladium nanoparticles as catalysts for <strong>Heck</strong> reaction<br />

A palladium-nanoparticle-cored G-3 dendrimer, characterized by TEM, TGA,<br />

absorption, and IR spectroscopies, has approximately 300 Pd atoms in <strong>the</strong> metallic core<br />

and an average diameter <strong>of</strong> 2.0 nm, to which are attached fourteen G-3 dendrons.<br />

Nearly 90% <strong>of</strong> <strong>the</strong> metal nanoparticle surface is unpassivated and available for<br />

catalysis. The dendrons inhibit metal agglomeration without adversely affecting<br />

chemical reactivity. Thus, investigations have shown that Pd-G-3 can efficiently<br />

catalyze <strong>Heck</strong> and Suzuki reactions [8]<br />

Syn<strong>the</strong>sis <strong>of</strong> Pd-G-3, in Which Seven <strong>of</strong> <strong>the</strong> Fourteen G-3 Wedges Are Shown (see<br />

text)<br />

Pd-G-3 was prepared by <strong>the</strong> Brust reaction<br />

(K2PdCl4 was phase transferred into toluene using<br />

tetraoctylammonium bromide (TOAB). Fre´chettype<br />

dendritic polyaryl e<strong>the</strong>r disulfide3 <strong>of</strong><br />

generation 3 (G-3S) was <strong>the</strong>n used. The mixture<br />

was cooled in ice (0-2 °C) and excess <strong>of</strong> NaBH4<br />

was added. The Pd-G-3 thus obtained was a black<br />

powder, freely soluble in methylene chloride,<br />

chlor<strong>of</strong>orm, toluene and THF, but insoluble in<br />

e<strong>the</strong>r and alcohol. Pd-G-3 was stable for several<br />

months both as a powder and as a dilute solution<br />

in CH2Cl2. FT-IR and TGA experiments showed<br />

that Pd-G-3 incorporated some amount <strong>of</strong> TOAB<br />

and this could not be removed by <strong>the</strong> standard<br />

purification procedure. Figure 5 shows a highresolution<br />

TEM image <strong>of</strong> Pd-G-3 and <strong>the</strong><br />

corresponding core-size histogram. It can be seen<br />

from Figure 1 that <strong>the</strong> particles exhibit a relatively<br />

wide size distribution (1-5 nm). The mean core<br />

diameter obtained from <strong>the</strong> histogram plot was 2.0<br />

nm<br />

TEM image and core-size histogram <strong>of</strong><br />

Pd-G-3.<br />

15


It was used Pd-G-3 [8,9] as catalyst in <strong>the</strong> <strong>Heck</strong> reactions shown in Scheme 2. Typically, 10<br />

mM each <strong>of</strong> <strong>the</strong> reactants and 20 mM triethylamine were taken up in toluene and <strong>the</strong> mixture was<br />

heated to reflux for 24 h in <strong>the</strong> presence <strong>of</strong> 10 mg (2- 10-3 mol %) Pd-G-3. Pd-G-3 is soluble in<br />

toluene and, hence, reactions 1 and 2 occur under homogeneous catalytic conditions. After <strong>the</strong><br />

reaction, toluene was removed in a rotavapor and <strong>the</strong> residue extracted with e<strong>the</strong>r.<br />

The turnover numbers (TON ) mol product/mol catalyst) and turnover frequencies (TOF ) mol<br />

product/mol catalyst/hour) given in Scheme 2 were calculated from isolated product yields. The<br />

reaction with ethyl acrylate proceeded with good yield, although that with styrene was low. The<br />

TON and TOF, however, are very high for both reactions. This, coupled with <strong>the</strong> absence <strong>of</strong> side<br />

products, indicates that <strong>the</strong> yields could be improved by using slightly higher amounts <strong>of</strong> Pd-G-3.<br />

In most Pd catalyzed reactions, 0.5-2 mol % catalyst is generally used and this is 100-1000 times<br />

higher than what we have employed here.<br />

The catalytic properties <strong>of</strong> <strong>the</strong> Pd nanoparticles dispersed in BMI. PF6 were tested in<br />

<strong>the</strong> coupling <strong>of</strong> aryl halides with n-butylacrylate at different temperatures (Table 1).<br />

First, we tested different bases (NaOAc, NEt(iPr)2, DABCO, and Na2CO3) in <strong>the</strong><br />

reaction <strong>of</strong> iodobenzene with n-butylacrylate at 80 oC ([PhI]/[Pd] ) 1000). A mixture <strong>of</strong><br />

<strong>the</strong> aryl halide (1.0 mmol) and n-butyl acrylate (1.2 mmol) with <strong>the</strong> Pd nanoparticles<br />

dispersed in 0.5 mL <strong>of</strong> ionic liquid gave a light-yellow solution in <strong>the</strong> presence <strong>of</strong><br />

NEt(iPr)2 and suspensions in <strong>the</strong> presence <strong>of</strong> <strong>the</strong> o<strong>the</strong>r bases.<br />

16


Figure 2. Possible pathways involved in <strong>the</strong> <strong>Heck</strong> reaction promoted by Pd nanoparticles<br />

dispersed in imidazolium ionic liquids.<br />

Almost complete iodobenzene conversion was observed in <strong>the</strong> reaction experiment performed<br />

with NEt(iPr)2, whereas only 60%). A similar trend was recently observed in <strong>Heck</strong> reactions promoted by a<br />

heterogeneous catalyst precursor where it was proposed that Pd dissolution and<br />

reprecipitation are inherent parts <strong>of</strong> <strong>the</strong> catalytic cycle.<br />

References<br />

[1] (a) R. B. Bedford, C. S.J. Cazin, D. Holder, The development <strong>of</strong> palladium catalysts for C–C and C–heteroatom<br />

bond forming reactions <strong>of</strong> aryl chloride substrates, Coord. Chem. Rev. 248 (2004) 2283–2321;<br />

(b) Applied Homogeneous Catalysis with Organometallic Compounds; Edited by B. Cornils, W.A. Hermann, VCH,<br />

Weinheim-New York, 1996;<br />

(c ) http://en.wikipedia.org/wiki/<strong>Heck</strong>_reaction<br />

[2] (a) R. F. <strong>Heck</strong>, Jr., J. P. Nolley, Palladium-catalyzed vinylic hydrogen substitution reactions with aryl, benzyl, and<br />

styryl halides, J. Org. Chem. 37(14) (1972) 2320–2322. doi:10.1021/jo00979a024.<br />

(b) T.Mizoroki, K. Mori, A. Ozaki, Arylation <strong>of</strong> Olefin with Aryl Iodide Catalyzed by Palladium, Bull. Chem. Soc. Jap.<br />

44 (1971) 581. doi:10.1246/bcsj.44.581.<br />

[3] F.Ozawa, A.Kubo, T.Hayashi, Generation <strong>of</strong> Tertiary Phosphine-Coordinated Pd(0) Species from Pd(OAc)2 in <strong>the</strong><br />

Catalytic <strong>Heck</strong> <strong>Reaction</strong>, Chemistry Lett. (1992) 2177–2180. doi:10.1246/cl.1992.2177.<br />

[4] (a) Hagiwara, Hisahiro, Sustainable Mizoroki–<strong>Heck</strong> reaction in water: remarkably high activity <strong>of</strong> Pd(OAc)2<br />

immobilized on reversed phase silica gel with <strong>the</strong> aid <strong>of</strong> an ionic liquid, Chemical Communications. 23(2005) 2942–<br />

2944. doi:10.1039/b502528a.<br />

(b) L. Kiss, T. Kurtan, S. Antus, H. Brunner, Fur<strong>the</strong>r insight into <strong>the</strong> mechanism <strong>of</strong> <strong>Heck</strong> oxyarylation in <strong>the</strong> presence<br />

<strong>of</strong> chiral ligands, Arkivoc (2003) GB–653J. http://www.arkat-usa.org/ark/journal/2003/I05_Bernath/GB-653J/GB-<br />

653J.asp.<br />

(c) M. Kitamura, D. Kudo, K. Narasaka, Palladium(0)-catalyzed syn<strong>the</strong>sis <strong>of</strong> pyridines from β-acetoxy-γ,δ-unsaturated<br />

ketone oximes, Arkivoc (2005) JC–1563E. http://www.arkat-usa.org/ark/journal/2006/I03_Coxon/1563/1563.asp.<br />

[5] J. G. De Vries, The <strong>Heck</strong> reaction in <strong>the</strong> production <strong>of</strong> fine chemicals, Canadian Journal <strong>of</strong> Chemistry 79 (2001)<br />

1086. doi:10.1139/cjc-79-5-6-1086.<br />

[6] J.Dupont, M.Pfeffer, J. Spencer, Palladacycles - An Old Organometallic Family Revisited: New, Simple, and<br />

Efficient Catalyst Precursors for Homogeneous Catalysis, Eur. J. Inorg. Chem. (2001) 1917-1927<br />

[7]D. Kovala-Demertzi, P. N. Yadav, M. A. Demertzis, J. P. Jasiski, F. J. Andreadaki, I. D. Kostas, First use <strong>of</strong> a<br />

palladium complex with a thiosemicarbazone-based ligand as catalyst precurs or for <strong>the</strong> <strong>Heck</strong> reaction, Tetrahedron<br />

Letters ,45 (2004) 2923–2926.<br />

[8] C. C. Cassol, A. P. Umpierre, G. Machado, S. I. Wolke, J. Dupont, The Role <strong>of</strong> Pd Nanoparticles in Ionic Liquid in<br />

<strong>the</strong> <strong>Heck</strong> <strong>Reaction</strong>, J. Am. Chem. Soc. 127 (2005) 3298-3299, DOI: 10.1021/ja0430043<br />

17

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!