You are on page 1of 8

FULL PAPERS

DOI: 10.1002/adsc.200900270

“Click” Polymer-Supported Palladium Nanoparticles as Highly


Efficient Catalysts for Olefin Hydrogenation and Suzuki Coupling
Reactions under Ambient Conditions
Ctia Ornelas,a Abdou K. Diallo,a Jaime Ruiz,a and Didier Astruca,*
a
Institut des Sciences Molculaires, UMR CNRS N85255, Universit Bordeaux 1, 33405 Talence Cedex, France
Fax: (+ 33)-5-4000-6994; e-mail: d.astruc@ism.u-bordeaux1.fr

Received: April 17, 2009; Revised: June 30, 2009; Published online: September 8, 2009

Supporting information for this article is available on the WWW under


http://dx.doi.org/10.1002/adsc.200900270.

Abstract: Complexation of palladium(II) acetate atm hydrogen, with turnover numbers (TONs) of
[PdACHTUNGRE(OAc)2] or dipotassium tetrachloropalladate 200,000. When stabilized by the water-soluble poly-
[K2PdCl4] to “click” polymers functionalized with ACHTUNGRE(sodium sulfonate-triazolylmethyl)styrene, the
phenyl, ferrocenyl and sodium sulfonate groups gave PdNPs (0.01% mol Pd) catalyze the Suzuki–Miyaura
polymeric palladium(II)-triazolyl complexes that coupling between iodobenzene (PhI) and phenylbor-
were reduced to “click” polymer-stabilized palladium onic acid [PhB(OH)2] in water/ethanol (H2O/EtOH)
nanoparticles (PdNPs). Transmission electron micro- at 25 8C with TONs of 8,200. This high catalytic ac-
scopy (TEM) showed that reduction using sodium tivity is comparable to that obtained with “click”
borohydride (NaBH4) produced PdNPs in the 1– dendrimer-stabilized PdNPs under ambient condi-
3 nm range of diameters depending on the nature of tions.
the functional group, whereas slow reduction using
methanol yielded PdNPs in the 22–25 nm range. The
most active of these PdNPs (0.01% mol Pd), stabi- Keywords: catalysis; “click” polymers; hydrogena-
lized by poly(ferrocenyltriazolylmethyl)styrene, cata- tion; palladium nanoparticles; Suzuki–Miyaura reac-
lyzed the hydrogenation of styrene at 25 8C and 1 tion; water-soluble catalyst

Introduction The concept of using nitrogen-based ligands present


on the dendritic backbone for encapsulation of Pd(II)
Over the last decade, metal nanoparticles (NPs) have whereby, upon reduction to Pd(0), the PdNPs formed
appeared as one of the most promising solutions to- are stabilized by the dendrimers, was pioneered by
wards efficient catalysis under mild, environmentally Crooks and colleagues who have used the poly(ami-
benign conditions. The stabilization of NPs for further doamine) dendrimers (PAMAM).[1a,8] Recently we
catalytic use has been achieved by a variety of means have reported that triazole-containing dendrimers,
including ligands, surfactants, ionic liquids, supercriti- synthesized by the “click” reaction of copper-cata-
cal micro-emulsions, polymers and dendrimers. Catal- lyzed azide/alkyne 1,3-dipolar cycloaddition
ysis by polymer-supported metal nanoparticles has (CuAAC),[9] are able to bind Pd(II) by coordination
been the subject of extensive studies since the 1990s of their triazole ligands.[10] We have also shown that,
due to the great efficiency in many reactions with after reduction of the Pd(II) already incorporated in
mechanistic issues.[1–3] In particular, polymer-stabilized the dendritic backbone to Pd(0), the PdNPs formed
palladium nanoparticles (PdNPs) are of special inter- are stabilized by the dendrimers, and that these
est,[4] because they are active in olefin hydrogena- “click” dendrimer-encapsulated nanoparticles (DENs)
tion[4,5] and various C C coupling reactions[6] that are are much more stable and catalytically efficient than
usually catalyzed by palladium complexes.[3,7] In addi- those stabilized by PAMAM dendrimers.[11] The main
tion, it is possible to provide water-soluble as well as advantage of the DENs is that, knowing the ratio of
organo-soluble PdNPs depending on the hydrophilici- Pd(II) per triazole ligand, it is possible to control the
ty of the supporting polymers.[1,4] exact number of metal atoms in the NPs and there-
fore their sizes.[11]

Adv. Synth. Catal. 2009, 351, 2147 – 2154  2009 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim 2147
FULL PAPERS Ctia Ornelas et al.

The aim of the present work is to investigate Results


whether triazole-containing polymers,[12–14] that are
more easily accessible than dendrimers, can also be Synthesis of the “Click” Polymers
catalytically efficient and stable as the “click” DENs.
Although the functionalization of polymers using the Classic radical polymerization of p-chloromethylstyr-
“click” CuAAC reaction has been extensively exploit- ene, using AIBN as initiator, afforded poly(chlorome-
ed over the last five years,[12–14] the presence of the tri- thyl)styrene (1) in 67% yield. The resulting poly-
azole heterocycles on the polymeric structures have ACHTUNGRE(chloromethyl)styrene was analyzed by size exclusion
not been yet exploited for metal coordination purpos- chromatography (SEC) showing an average molecular
es. In this work, we have functionalized the easily ac- weight (MW) of 3,600 that corresponds to a degree of
cessible poly(azidomethyl)styrene with three distinct polymerization (DP) of 24 and a polydispersity index
terminal alkynes (phenylacetylene, ethynylferrocene (PDI) of 1.33. The chlorides in polymer 1 were substi-
and sodium propargylsulfonate) using the “click” tuted by azide groups upon reaction with sodium
CuAAC reaction. The three resulting triazole-con- azide in dimethylformamide (DMF) at 80 8C for 16 h,
taining polymers were used to bind Pd(II) that, upon yielding the poly(azidomethyl)styrene 2. The comple-
reduction to Pd(0), formed the first “click” polymer- tion of the azidation reaction was verified using
1
stabilized PdNPs that are catalytically active in olefin H NMR by the complete disappearance of the
hydrogenation and Suzuki–Miyaura C C coupling re- CH2Cl peak at 4.53 ppm and the appearance of the
actions under mild conditions. CH2N3 peak at 4.27 ppm.
Three polymers containing triazole ligands were
synthesized by “click” reaction[9] between the azido
groups of polymer 2 and phenylacetylene, ethynylfer-
rocene and sodium propargylsulfonate, using copper
sulfate/sodium ascorbate in THF/water (1:1)
(Scheme 1). The completion of the “click” reactions

Scheme 1. Synthesis of the “click” polymers.

2148 asc.wiley-vch.de  2009 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim Adv. Synth. Catal. 2009, 351, 2147 – 2154
“Click” Polymer-Supported Palladium Nanoparticles as Highly Efficient Catalysts FULL PAPERS

was verified using 1H NMR by the total disappear- Table 1. Diameters of the PdNPs stabilized by the “click”
ance of the CH2N3 peak at 4.27 ppm as well as by in- polymers, obtained by TEM.
frared spectroscopy (IR) by complete disappearance
PdNPs Reducing agent Diameter [nm]
of the N3 band at 2090 cm 1. The resulting polymers
poly(phenyltriazolylmethyl)styrene (3), poly(ferro- PdNPs stabilized by 3 NaBH4 1.0  0.2
ACHTUNGREcenyltriazolyl-methyl)styrene (4) and poly(sodium sul- PdNPs stabilized by 3 MeOH 25.2  6.0
fonate-triazolylmethyl)styrene (5) were characterized PdNPs stabilized by 4 NaBH4 1.7  0.4
by 1H and 13C NMR, elemental analysis and SEC, and PdNPs stabilized by 4 MeOH 22.4  4.7
PdNPs stabilized by 5 NaBH4 2.8  0.4
they all disclosed a degree of polymerization of ap-
proximately 24 and a PDI of 1.33, showing that the
azidation as well as the CuAAC reaction did not
change the polymer backbone. Polymers 3 and 4 are
soluble in most common organic solvents such as sequent to reduction using NaBH4, respectively (see
methylene chloride, chloroform and tetrahydrofuran, Supporting Information for all the other TEM images
whereas polymer 5 is soluble in aqueous media due to and histograms).
the presence of the sodium sulfonate groups. The TEM data show that the PdNPs obtained by
reduction with NaBH4 are much smaller (1–3 nm
range) than those obtained by reduction with metha-
Complexation of Pd(II) and Characterization of the nol (22–25 nm range). This result is in accordance
PdNPs Stabilized by the “Click” Polymers with the tendency found for polymer-stabilized
PdNPs already reported.[2,3] For the polymer-stabilized
In our previous reports, we showed that triazole het- PdNPs, the size of the NPs is determined by the metal
erocycles can complex Pd(II), the ratio of triazole reduction rate (the reduction with methanol is much
ring per palladium atom being dependent on the type slower than that using NaBH4) in contrast to the den-
of palladium precursor complex.[10] We also showed drimer-stabilized PdNPs in which the size is governed
that, in organic solvents in presence of methanol, tri- by the number of Pd(II)-triazole units in the dendri-
ACHTUNGREazole ligands coordinate to PdACHTUNGRE(OAc)2 in a Pd:triazole mer.[11]
ring ratio of 1:1,[11c] and in aqueous media the triazole
ligands also coordinate to K2PdCl4 in a Pd:triazole
ring 1:1 ratio.[11d] This ratio is now applied to the Catalytic Efficiency of the “Click” Polymer-Stabilized
“click” polymers in order to introduce Pd(II) into the PdNPs in the Hydrogenation and Suzuki–Miyaura
polymers that, after reduction, would afford polymer- Reactions
stabilized-PdNPs.
For polymers 4 and 5, the polymer-stabilized-NPs The “click” polymer-stabilized PdNPs described
were synthesized in CHCl3/MeOH (2:1) by adding above were tested as catalysts in hydrogenation of
one equivalent of PdACHTUNGRE(OAc)2 per triazole unit (since styrene to ethylbenzene and Suzuki–Miyaura coupling
the click reactions were complete, the mass of poly- between PhI and PhB(OH)2, in order to investigate
mer was divided by the molecular weight of the mon- their catalytic efficiency and stability and eventually
omeric unit, in order to calculate the amount of palla- to compare them with the “click” DENs reported ear-
dium complex to be added). The reduction of Pd(II) lier.[11] The amount of catalyst used in both reactions
to Pd(0) in polymers 4 and 5 was performed either by was 0.01% mol Pd, and the results obtained for the
stirring in the presence of methanol for 16 h or by hydrogenation and Suzuki–Miyaura reactions are
adding 10 equivalents of NaBH4, in order to compare gathered in Table 2 and Table 3, respectively. The re-
the influence of the reducing agents on the NP size actions were followed by GC, and the turnover fre-
and catalytic activity. For polymer 5, the polymer-sta- quency values (TOF) were determined on the basis of
bilized NPs were synthesized in aqueous solution by the yield of formation of final product per hour and
addition of one equivalent of K2PdCl4 per triazole are expressed in mol substrate (mol Pd) 1 h 1. The
unit, followed by addition of 10 equivalents of turnover number values (TON) express the total
NaBH4. After complexation of Pd(II) [PdACHTUNGRE(OAc)2 or amount of substrate catalyzed before degradation and
K2PdCl4] to the triazole ligands of the polymers fol- are expressed in mol substrate/mol Pd (in the case of
lowed by reduction to Pd(0), all the polymer-stabi- the hydrogenation reactions, the substrate was added
lized NPs were characterized by transmission electron several times until complete degradation of the cata-
microscopy (TEM). The TEM data are gathered in lyst).
Table 1. Figure 1 and Figure 2 show the TEM pictures Table 2 shows that the PdNPs, stabilized by poly-
and size-distribution histograms of the PdNPs stabi- mer 3 and soluble in organic solvents, and those stabi-
lized by polymer 4 subsequent to reduction using lized by polymer 4 have comparable activities, but the
methanol and the PdNPs stabilized by polymer 5 sub- latter are much more stable than the former, perform-

Adv. Synth. Catal. 2009, 351, 2147 – 2154  2009 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim asc.wiley-vch.de 2149
FULL PAPERS Ctia Ornelas et al.

Figure 1. PdNPs stabilized by polymer 4 using methanol as the reducing agent: a) TEM image and b) size-distribution histo-
gram.

Figure 2. PdNPs stabilized by polymer 5 using NaBH4 as the reducing agent: a) TEM image and b) size-distribution histo-
gram.

Table 2. Catalytic efficiencies (TOF) and stabilities (TON) Discussion


obtained for all the new catalysts for styrene hydrogenation
at 0.01% mol Pd, 25 8C and 1 atm H2. In the present work, it was anticipated that the advan-
Catalyst Reducing agent TOF [a]
TON [b] tages disclosed in the strategy utilizing the “click”
dendrimer-stabilized PdNPs would still be beneficial,
PdNPs stabilized by 3 NaBH4 1,100 26,000 at least to a certain extent, for catalysis with “click”
PdNPs stabilized by 3 MeOH 1,400 17,000 polymer-stabilized PdNPs. Even if the “click” poly-
PdNPs stabilized by 4 NaBH4 930 200,000 mers, contrary to dendrimers, do not allow the precise
PdNPs stabilized by 4 MeOH 1,740 150,000
control over the NP size, the results show that the
[a]
The catalytic activity of the PdNPs was investigated for “click” polymer-stabilized PdNPs present a catalytic
the hydrogenation of styrene at 0.01% mol Pd, in CHCl3/ efficiency of the same order as that disclosed earlier
MeOH (2:1), 25 8C and 1 atm H2. The reactions were fol- for the “click” dendrimer-stabilized PdNPs. In the
lowed by GC, and the TOF values were determined on present study, two extremely different PdNP sizes
the basis of the yield of ethylbenzene per hour and are were obtained by variation of the nature of the reduc-
expressed in mol H2 (mol Pd) 1 h 1. tant (NaBH4 vs. methanol). As expected, the PdNPs
[b]
The TONs are expressed in mol substrate/mol Pd.
obtained using NaBH4 are much smaller than those
using methanol because the stronger reductant
(NaBH4) produces the smaller PdNPs.[15] While the
ing up to 200,000 catalytic cycles (mol substrate/mol relation between the strength of the reducing agent
Pd) for styrene hydrogenation at 25 8C. and the NP size is easy to understand, the comparison
Table 3 shows that the water-soluble PdNPs stabi- between the catalytic efficiency of the various NPs is
lized by polymer 5 are the most active PdNPs in the not straightforward. It is expected that smaller NPs
Suzuki–Miyaura coupling between iodobenzene and would be much more active (per mol Pd) than larger
phenylboronic acid, presenting a catalytic efficiency NPs. Our results, however, show that the larger
of 900 mol PhI (mol Pd) 1 h 1 and a yield of 82% PdNPs are more active than the smaller ones. This is
after 24 h at 25 8C (TON = 8,200). probably due to the difference of nature of the PdNP

2150 asc.wiley-vch.de  2009 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim Adv. Synth. Catal. 2009, 351, 2147 – 2154
“Click” Polymer-Supported Palladium Nanoparticles as Highly Efficient Catalysts FULL PAPERS

Table 3. Catalytic efficiencies (TOF) and stabilities (TON) obtained for all the new catalysts for the Suzuki–Miyaura cou-
pling between PhB(OH)2 and PhI at 0.01% mol Pd and 25 8C.

Catalyst Reducing agent Time (Yield) Time (Yield) Time (Yield) TOF[c] TON[d]
PdNPs stabilized by 3[a] NaBH4 24 h (14%) 48 h (15%) – 60 1,500
PdNPs stabilized by 3[a] MeOH 24 h (32%) 48 h (41%) 72 h (44%) 130 4,400
PdNPs stabilized by 4[a] NaBH4 24 h (14%) 48 h (15%) – 60 1,500
PdNPs stabilized by 4[a] MeOH 24 h (32%) 48 h (41%) 72 h (44%) 130 4,400
PdNPs stabilized by 5[b] NaBH4 7 h (62%) 24 h (82%) – 900 8,200
[a]
The PdNPs were formed using PdACHTUNGRE(OAc)2, and the catalytic reactions were performed in CHCl3/MeOH (2:1), at 0.01%
mol Pd, 25 8C.
[b]
The PdNPs were formed using K2PdCl4, and the catalytic reactions were performed in H2O/EtOH (1:1) at 0.01% mol Pd,
25 8C.
[c]
The reactions were followed by GC, and the TOF values were determined on the basis of the yield of formation of bi-
phenyl. The TOFs are expressed in mol PhI (mol Pd) 1 h 1.
[d]
The TONs are expressed in mol substrate/mol Pd.

surfaces that result from the different reducing agents Conclusions


used. In particular the borates, that were formed
when NaBH4 was used, are not formed along the Coordination of Pd(II) by triazole-containing poly-
methanol reduction. On the one hand, the PdNPs mers, followed by reduction to Pd(0) using either
produced by reduction with NaBH4 are smaller, NaBH4 or methanol, yields small (1–3 nm) or, respec-
which should have made them more efficient, but on tively, large (21–24 nm) “click” polymer-stabilized
the other hand the borates formed at their surface PdNPs that are soluble in organic solvents or in water
probably reduce their catalytic activity. depending on the functional group of the terminal
For styrene hydrogenation, the PdNPs stabilized by alkyne used in the “click” reaction. The three new
polymer 4 and soluble in organic solvents are much “click” polymer-stabilized PdNPs are active catalysts
more stable (the TON reaching values of 200,000 mol for styrene hydrogenation and Suzuki–Miyaura reac-
H2 {mol Pd} 1) than those stabilized by polymer 3 tions at 25 8C using 0.01% mol Pd. For styrene hydro-
(TON values are approximately eight times higher for genation, the most active among these PdNP catalysts
the former catalyst than for the latter one). It is prob- are those stabilized with “click” polymers containing
able that this higher stability obtained for NPs that ferrocenyl groups on the side chain in which the
are stabilized by polymer 4 is due to the tri-dimen- Pd(II) was reduced to Pd(0) using methanol (styrene
sional bulk of the ferrocenyl groups acting as steric hydrogenation at 1 atm H2 with TON up to 200,000).
protectors of the metal NPs (without coordination to For the Suzuki–Miyaura reaction, the most active cat-
the PdNP surface that would decrease the activity). alyst is the one for which the PdNPs are stabilized by
On the other hand, there is no significant difference “click” polymers containing sodium sulfonate groups
on the catalytic activity of the NPs stabilized by poly- on the side chain providing water solubility [82%
mers 3 and 4 in the Suzuki–Miyaura coupling, proba- yield in 24 h, TOF = 880 mol PhI (mol Pd) 1 h 1 and
bly due to the nature of the leaching mechanism (that TON = 8,200].
is typical of this type of reaction when catalyzed by These new “click” polymer-stabilized PdNPs show
PdNPs).[3,11] Note that the Suzuki–Miyaura coupling a high catalytic activity under ambient conditions that
reaction was successfully performed in aqueous media is comparable to that of the “click” dendrimer-stabi-
(at 0.01% mol Pd, at 25 8C) using the water-soluble lized PdNPs. The remaining challenge is extension of
PdNPs stabilized by polymer 5, reaching 82% yield in this type of catalysts Suzuki–Miyaura coupling with
24 h. The fact that the best overall results for the aryl chlorides for which some molecular Pd-phos-
Suzuki–Miyaura coupling are obtained with the phine catalysts are very efficient.[7]
water-soluble “click” polymer-stabilized NPs in aque-
ous solvent using K2PdCl4 as a source of Pd(II) and
NaBH4 as a reducing agent indicates that the borates
formed do not significantly inhibit the PdNP surface,
probably due to their hydrolysis in the aqueous sol- Experimental Section
vent.
General Data
Ethynylferrocene, phenylacetylene, and all the olefinic sub-
strates were purchased from Sigma–Aldrich. Sodium prop-
argyl sulfonate was synthesized according to ref.[16] Synthe-

Adv. Synth. Catal. 2009, 351, 2147 – 2154  2009 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim asc.wiley-vch.de 2151
FULL PAPERS Ctia Ornelas et al.

ses of the PdNPs were carried out using Schlenk techniques 6300 g·mol 1; PDI: 1.33; DP: 24; anal. calcd. for [C17H15N3]n :
using degassed solvents. C 78.13, H 5.79; found: C 77.35, H 6.12.

Synthesis of Poly(chloromethyl)styrene, Polymer 1 Synthesis of Poly(ferrocenyltriazolylmethyl)styrene,


Polymer 4
p-Chloromethylstyrene (2.17 g, 14.2 mmol) was dissolved in
60 mL of dry toluene and heated to 80 8C. AIBN (0.116 g, Polymer 2 (0.115 g) and ethynylferrocene (0.335 g,
0.701 mmol) was added to the solution, and the reaction 1.60 mmol) were dissolved in 60 mL of tetrahydrofuran
mixture was allowed to stir at 80 8C for 48 h. Toluene was re- (THF), and 60 mL of water weere added (1:1 THF/water).
moved under vacuum, and the crude product was washed At 0 8C, CuSO4·5 H2O was added (0.265 g, 1.06 mmol; 1 M
with methanol and precipitated twice with dichloromethane/ aqueous solution) followed by dropwise addition of a freshly
methanol. The polymer poly(chloromethylstyrene) was ob- prepared solution of sodium ascorbate (0.420 g, 2.12 mmol;
tained as a white powder; yield: 1.44 g (67%). 1H NMR 1 M aqueous solution). The solution was allowed to stir for
(CDCl3, 300 MHz): d = 7.06 and 6.49 (arom. CH of styrenic 16 h at room temperature. After removing THF under
ring), 4.53 (CH2Cl), 1.67 and 1.39 (CH and CH2 of polymer vacuum, dichloromethane and an aqueous solution of am-
chain); 13C NMR (CDCl3, 75 MHz): d = 138.8 and 135.5 monia were added. The mixture was allowed to stir for
(arom. Cq), 127.0 and 126.6 (arom. CH of styrenic ring), 51.5 10 min in order to remove the copper trapped inside the po-
(CH2Cl), 43.1 and 41.3 (CH and CH2 of polymer chain). lymer. The organic phase was washed twice with water,
SEC data: MW: 3600 g·mol 1; PDI: 1.33; DP: 24. dried with sodium sulfate, filtered through paper, and the
solvent was removed under vacuum. The product was
washed with pentane in order to remove the excess of
Synthesis of Poly(azidomethyl)styrene, Polymer 2 alkyne and precipitated twice using dichloromethane/pen-
tane. The poly(p-ferrocenyltriazolylmethylstyrene) was ob-
Polymer 1 (0.700 g) was dissolved in 30 mL of dry DMF,
tained as an orange waxy product; yield: 0.233 g (69%).
NaN3 (0.90 g, 14.0 mmol) was added, and the reaction mix- 1
H NMR (CDCl3, 300 MHz): d = 7.73 (CH of triazole), 6.73
ture was heated at 80 8C for 16 h. The solvent was removed
and 6.24 (arom. CH of styrenic ring), 5.37 (CH2-triazole),
under vacuum, and the product was dissolved in dichloro-
4.78, 4.30 and 4.07 (CH of ferrocenyl group), 1.36 (CH and
methane and washed with water. The organic layer was
CH2 of polymer chain); 13C NMR (CDCl3, 75 MHz): d =
dried with sodium sulfate, filtered, and the solvent was re-
146.4 (Cq of triazole), 132.4–127.5 (aromatics), 119.4 (CH of
moved. Polymer 2 was obtained as a white waxy product;
triazole), 75.1, 69.2, 68.3 and 66.2 (CH and Cq of ferrocenyl
yield: 0.645 g (91%). 1H NMR (CDCl3, 300 MHz), d = 7.01
group), 52.9 (CH2-triazole), 41.3 (CH and CH2 of polymer
and 6.51 (arom. CH of styrenic ring), 4.27 (CH2N3), 1.72 and
chain); SEC data: MW: 9000 g·mol 1; PDI: 1.33; DP:24;
1.45 (CH and CH2 of polymer chain); 13C NMR (CDCl3,
anal. calcd. for [C21H19N3Fe]n : C 68.31, H 5.19; found: C
75 MHz), d = 144.3 and 132.1 (arom. Cq), 127.8 and 127.4
67.80, H 5.45.
(arom. CH of styrenic ring), 53.9 (CH2Cl), 39.7 (CH and
CH2 of polymer chain).
Synthesis of Poly(sodium sulfonate-
triazolylmethyl)styrene, Polymer 5
Synthesis of Poly(phenyltriazolylmethyl)styrene,
Polymer 3 The poly(azidomethylstyrene) polymer (0.200 g) was dis-
solved in 30 mL of tetrahydrofuran (THF), and 20 mL of an
Polymer 2 (0.165 g) and phenylacetylene (0.5 mL) were dis- aqueous solution of sodium propargylsulfonate (0.395 g,
solved in 60 mL of tetrahydrofuran (THF) and 60 mL of 2.78 mmol) were added. At 0 8C, CuSO4 (0.462 g, 1.85 mmol;
water were added (1:1 THF/water). At 0 8C, CuSO4·5 H2O 1 M aqueous solution) was added, followed by the dropwise
was added (0.347 g, 1.39 mmol; 1 M aqueous solution) fol- addition of a freshly prepared solution of sodium ascorbate
lowed by dropwise addition of a freshly prepared solution of (0.733 g, 3.70 mmol, 1 M aqueous solution). Water was
sodium ascorbate (0.551 g, 2.78 mmol; 1 M aqueous solu- added in order to obtain a solution in 1:1 THF/H2O. The so-
tion). The solution was allowed to stir for 16 h at room tem- lution was allowed to stir for 16 h at room temperature.
perature. After removing THF under vacuum, dichlorome- After removing THF under vacuum, the polymer remained
thane and an aqueous solution of ammonia were added. The dissolved in the aqueous solution. The polymer was purified
mixture was allowed to stir for 10 min, in order to remove by dialysis against 1 L of aqueous ammonia solution for
the copper trapped inside the polymer. The organic phase three days [in order to remove Cu(I) that remains trapped
was washed twice with water, dried with sodium sulfate, fil- inside the polymer] and against 1 L water for one day (the
tered, and the solvent was removed under vacuum. The dialysis solutions were substituted twice a day). After re-
product was washed with pentane in order to remove the moving water, the polymer poly(sodium sulfonatetriazolyl-
excess of alkyne and precipitated twice using dichlorome- methylstyrene) was obtained as a light yellow waxy product;
thane/pentane. Polymer 3 was obtained as a white powder; yield: 0.100 g (30%). 1H NMR (D2O, 300 MHz): d = 7.83
yield: 0.237 g (65%). 1H NMR (CDCl3, 300 MHz): d = 8.02 (CH of triazole), 7.10–6.90 (arom. CH of styrenic ring), 4.2
(CH of triazole), 7.79 and 7.30 (CH of phenyl ring), 6.79 (CH2SO3Na), 1.20–0.70 (CH and CH2 of polymer chain);
and 6.31 (arom. CH of styrenic ring), 5.36 (CH2-triazole), SEC data: MW: 8000 g·mol 1; PDI: 1.33; DP: 28.
1.72 and 1.53 (CH and CH2 of polymer chain); 13C NMR
(CDCl3, 75 MHz): d = 147.9 (Cq of triazole), 132.9–125.8
(aromatics), 120.9 (CH of triazole), 53.7 (CH2-triazole), 41.3
(CH and CH2 of polymer chain); SEC data: MW:

2152 asc.wiley-vch.de  2009 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim Adv. Synth. Catal. 2009, 351, 2147 – 2154
“Click” Polymer-Supported Palladium Nanoparticles as Highly Efficient Catalysts FULL PAPERS

Transmission Electron Microscopy (TEM) The PdNPs were prepared in one Schlenk and then divided
4
in several Schlenk flasks containing the reactants for the
Samples were prepared by setting a drop of a 4  10 M so- Suzuki–Miyaura reaction. At given times, one reaction solu-
lution of PdNPs (concentration in mol Pd) on a holey- tion was extracted with dichloromethane (all reactants and
carbon-coated Cu TEM grid. The size of the PdNPs was final product are soluble) and analyzed by GC.
measured using the software sigmascanpro (for each sample,
about 100 nanoparticles were measured).
Supporting Information Available
Gas Chromatography (GC) TEM and size distribution histograms of all the click-poly-
mer-stabilized PdNPs.
GC data were recorded on a Hewlett Packard 5890 Series II
gas chromatograph, equipped with a flame ionization detec-
tor. Helium was used as the carrier gas for all the substrates.
The injector and detector temperatures were 240 8C. Acknowledgements
General Procedure for the Preparation of the PdNPs We are grateful to Dr. Eric Cloutet (LCPO, Bordeaux) for
helpful assistance in the SEC experiments, and to Fundażo
For PdNPs stabilized by polymers 3 or 4, the nanoparticles para a CiÞncia e a Tecnologia (FCT), Portugal (Ph. D. grant
were synthesized in CHCl3/MeOH (2:1) using 1 equiv. of to CO), the Institut Universitaire de France (IUF, DA), the
PdACHTUNGRE(OAc)2 per triazole unit. For PdNPs stabilized by polymer CNRS, the ANR (CATAQ project) and the Universit Bor-
5, the nanoparticles were synthesized in water using 1 equiv. deaux I for financial support.
K2PdCl4 per triazole unit.
1 mL of solution of “click” polymer (1 mg) was intro-
duced in a Schlenk flask. 1 mL of a solution containing
1 equiv. of PdACHTUNGRE(OAc)2 (in CHCl3) or K2PdCl4 (in water) per
References
triazole unit was added. Then NaBH4 (10 equiv. per Pd) was
[1] a) Y. Niu, R. M. Crooks, C. R. Chim. 2003, 6, 1049;
added, and the light-yellow solution turned to black indicat-
b) M-C. Daniel, D. Astruc, Chem. Rev. 2004, 104, 293;
ing nanoparticle formation. For the nanoparticles formed by
c) D. Astruc, F. Lu, J. Ruiz, Angew. Chem. 2005, 117,
methanol reduction, no NaBH4 was added allowing the
8062; Angew. Chem. Int. Ed. 2005, 44, 7852; d) J. G.
yellow CHCl3/MeOH solution containing the polymer-Pd
De Vries, Dalton Trans. 2006, 421; e) Nanocatalysis
complexes to stir for 16 h, under a nitrogen atmosphere, re-
(Eds.: U. Heiz, U. Landman), Springer, Berlin, 2007;
sulting in a dark-brown solution.
f) D. Astruc, Inorg. Chem. 2007, 46, 1884; g) Nanoparti-
cles and Catalysis, (Ed.: D. Astruc), Wiley-VCH, Wein-
Hydrogenation Reactions heim, 2008.
The PdNPs were freshly prepared in a Schlenk flask, and [2] a) H. Bçnnemann, W. Brijoux, in: Active Metals, (Ed.:
10,000 equiv. of substrate (per mol Pd) was added. The A. Frstner), VCH, Weinheim, 1996, p 339; b) M. T.
Schlenk flask was filled with H2 (1 atm), and the solution Reetz, W. Helbig, S. A. Quaiser, in: Active Metals,
was allowed to stir at 25 8C. For re-use of the catalyst, the (Ed.: A. Frstner), VCH, Weinheim, 1996, p 279; c) N.
substrate was added to the reaction solution until catalyst Toshima, Y. Yonezawa, New J. Chem. 1998, 22, 1179;
degradation. Calculation of the turnover frequency (TOF) d) B. F. G. Johnson, Coord. Chem. Rev. 1999, 190, 1269;
was carried out using several samples of the solution that e) A. Roucoux, J. Schulz, H. Patin, Chem. Rev. 2002,
were extracted at different reaction times and analyzed by 102, 3757; f) S. Mandal, P. R. Selvakannan, D. Roy,
GC. Calculation of the turnover number (TON) was carried R. V. Chaudhari, M. Sastry, Chem. Commun. 2002,
out using the sum of substrate that reacted in all the catalyt- 3002; g) J. Dupont, G. S. Fonseca, A. P. Umpierre,
ic cycles, after analyzing the final reaction solution (after P. F. P. Fichner, J. Am. Chem. Soc. 2002, 124, 4228;
catalyst degradation). h) H. Ohde, C. M. Wai, J. Kim, M. Ohde, J. Am. Chem.
Soc. 2002, 124, 4540; i) A. Roucoux, J. Schulz, H. Patin,
Suzuki–Miyaura Reactions Adv. Synth. Catal. 2003, 345, 222.
[3] N. T. S. Phan, M. Van der Sluys, C. J. Jones, Adv. Synth.
The PdNPs were freshly prepared in a Schlenk flask and Catal. 2006, 348, 609.
K3PO4, phenylboronic acid and iodobenzene were succes- [4] a) N. Toshima, K. Kushihashi, T. Yonezawa, H. Hirai,
sively added. The solution was allowed to stir under N2 at Chem. Lett. 1989, 1769; b) M. V. Seregina, L. M. Bron-
25 8C. stein, O. A. Platonova, D. M. Chernyshov, P. M. Valet-
For the reactions performed in organic solvents, the reac- sky, Chem. Mater. 1997, 9, 923; c) B. P. S. Chauhan, J. S.
tants (iodobenzene and phenylboronic acid and final prod- Rathore, T. Bandoo, J. Am. Chem. Soc. 2004, 126,
uct (biphenyl) were soluble in the reaction media, so the re- 8493; d) book on hydrogenation: The Handbook of
action kinetics was followed by simply taking several ali- Homogeneous Hydrogenation, (Eds.: J. G. de Vries,
quots of the reaction solutions and analyzing by GC. C. J. Elsevier), Wiley-VCH, Weinheim, 2006.
In the case of the reaction performed in H2O/EtOH, the [5] Hydrogenation reaction using polymer-stabilized
reactants (iodobenzene and phenylboronic acid were soluble PdNPs: a) H. Hirai, J. Macromol. Sci. Chem. 1979, 13,
in the reaction media, but the final product (biphenyl) was 633; b) N. Toshima, K. Kushihashi, T. Yonezawa, H.
not, precipitating as a white solid. In this case, the catalytic Hirai, Chem. Lett. 1989, 1769; c) M. V. Seregina, L. M.
reaction was performed in several Schlenk flasks in parallel. Bronstein, O. A. Platonova, D. M. Chernyshov, P. M.

Adv. Synth. Catal. 2009, 351, 2147 – 2154  2009 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim asc.wiley-vch.de 2153
FULL PAPERS Ctia Ornelas et al.

Valetsky, Chem. Mater. 1997, 9, 923; d) B. P. S. Chau- Chem. Int. Ed. 2002, 41, 2596; c) C. W. Tornoe, C.
han, J. S. Rathore, T. Bandoo, J. Am. Chem. Soc. 2004, Christensen, M. Medal, J. Org. Chem. 2002, 67, 3057;
126, 8493. d) P. Wu, A. K. Feldman, A. K. Nugent, C. J. Hawker,
[6] a) Y. Li, X. M. Hong, D. M. Collard, M. A. El-Sayed, A. Scheel, B. Voit, J. Pyun, J. M. J. Frchet, K. B.
Org. Lett. 2000, 15, 2384; b) V. Kogan, Z. Aizenshtat, Sharpless, V. V. Fokin, Angew. Chem. 2004, 116, 4018;
R. Popovitz-Biro, R. Neumann, Org. Lett. 2002, 4, Angew. Chem. Int. Ed. 2004, 43, 3928; e) V. D. Bock,
3529; c) R. Narayanan, M. A. El-Sayed, J. Am. Chem. H. Hiemstra, J. H. van Maarseveen, Eur. J. Org. Chem.
Soc. 2003, 125, 8340; d) H. Oyamada, R. Akiyama, H. 2006, 51.
Hagio, T. Naito, S. Kobayashi, Chem. Commun. 2006, [10] C. Ornelas, J. Ruiz, E. Cloutet, S. Alves, D. Astruc,
4297; e) D. Saha, K. Chattopadhyay, B. C. Ranu, Tetra- Angew. Chem. 2007, 119, 890; Angew. Chem. Int. Ed.
hedron Lett. 2008, 49, 1003; f) H.-B. Pan, C-H. Yen, B. Angew. Chem. Int. Ed. Engl. 2007, 46, 872.
Yoon, M. Sato, C. M. Wai, Synth. Commun. 2006, 36, [11] a) C. Ornelas, L. Salmon, J. R. Aranzaes, Chem.
3473; g) K. C. Jin, R. Najman, T. W. Dean, O. Ishihara, Commun. 2007, 4946; b) A. K. Diallo, C. Ornelas, L.
C. Muller, M. Bradley, J. Am. Chem. Soc. 2006, 128, Salmon, J. Ruiz, D. Astruc, Angew. Chem. 2007, 119,
6276; h) L. Z. Ren, L. J. Meng, Express Polym. Lett. 8798; Angew. Chem. Int. Ed. Engl. 2007, 46, 8644; c) C.
2008, 2, 251; i) X. Yang, Z. Fei, D. Zhao, W. H. Ang, Y. Ornelas, L. Salmon, J. Ruiz, D. Astruc, Chem. Eur. J.
Li, P. J. Dyson, Inorg. Chem. 2008, 47, 3292; j) J. Du- 2008, 14, 50; d) C. Ornelas, J. Ruiz, L. Salmon, D.
trand, E. Teuma, F Malbosc, Y. Kihn, M. Gomez, Astruc, Adv. Synth. Catal. 2008, 350, 837.
Catal. Commun. 2008, 9, 273. [12] First report on “click” polymers synthesis: B. Helms,
[7] a) J. Louie, J. F. Hartwig, Angew. Chem. 1996, 108, J. L. Mynar, C. J. Hawker, J. M. J. Frchet, J. Am.
2531; Angew. Chem. Int. Ed. Engl. 1996, 35, 2359; Chem. Soc. 2004, 126, 15020.
b) W. A. Herrmann, V. P. W. Bçlm, C. P. Reisinger, J. [13] Recent reviews on “click” polymers : a) M. Meldal,
Organomet. Chem. 1999, 576, 23; c) C. Amatore, A. Macromol. Rapid Commun. 2008, 29, 1016; b) W. H.
Jutand, Acc. Chem. Res. 2000, 33, 314; d) I. P. Belet- Binder, R. Sachsenhofer, Macromol. Rapid Commun.
skaya, A. V. Cheprakov, J. Organomet. Chem. 2004, 2008, 29, 952; c) P. Lundberg, C. J. Hawker, A. Hult,
689, 4055; e) R. B. Bedford, C. S. J. Cazin, D. Holder, M. Malkoch, Macromol. Rapid Commun. 2008, 29, 998;
Coord. Chem. Rev. 2004, 248, 2283; f) F. Alonso, I. P. d) B. Le Droumaguet, K. Velonia, Macromol. Rapid
Beletskaya, M. Yus, Tetrahedron 2005, 61, 11771. Commun. 2008, 29, 1073; e) J. A. Johnson, M. G. Finn,
[8] a) M. Zhao, R. M. Crooks, Angew. Chem. 1999, 111, J. T. Koberstein, N. J. Turro, Macromol. Rapid
375; Angew. Chem. Int. Ed. 1999, 38, 364; b) R. M. Commun. 2008, 29, 1421.
Crooks, M. Zhao, L. Sun, V. Chechik, L. K. Yeung, [14] a) G. K. Such, J. F. Quinn, A. Quinn, E. Tjipto, F.
Acc. Chem. Res. 2001, 34, 181; c) Y. Niu, L. K. Yeung, Caruso, J. Am. Chem. Soc. 2006, 128, 9318; b) S. Srini-
R. M. Crooks, J. Am. Chem. Soc. 2001, 123, 6840; vasachari, Y. Liu, L. E. Prevette, T. M. Reineke, Bio-
d) L. K. Yeung, R. M. Crooks, Nano Lett. 2001, 1, 14; materials 2007, 28, 2885; c) D. Daz, J. J. Marrero Tella-
e) Y. Niu, R. M. Crooks, Chem. Mater. 2003, 15, 3463; doa, D. Garca Velzqueza, . Gutirrez Raveloa, Tet-
f) R. W. J. Scott, O. M. Wilson, A.-K. Oh, E. A. Kenik, rahedron Lett. 2008, 49, 1340; d) L. Campos, K. Killops,
R. M. Crooks, J. Am. Chem. Soc. 2004, 126, 15583; R. Sakai, J. M. J. Paulusse, D. Damiron, E. Drocken-
g) S.-K. Oh, Y. Niu, R. M. Crooks, Langmuir 2005, 21, mller, B. W. Messmore, C. J. Hawker, Macromolecules
10209; h) O. M. Wilson, M. R. Knecht, J. C. Garcia- 2008, 41, 7063; e) E. Boisselier, A. Chan Kam Shun, J.
Martinez, R. M. Crooks, J. Am. Chem. Soc. 2006, 128, Ruiz, E. Cloutet, C. Belin, D. Astruc, New J. Chem.
4510. 2009, 33, 246; f) K. Itomi, S. Kobayashi, K. Morino, H.
[9] a) H. C. Kolb, M. G. Finn, K. B. Sharpless, Angew. Iida1, E. Yashima, Polym. J. 2009, 41, 108.
Chem. 2001, 113, 2056; Angew. Chem. Int. Ed. 2001, 40, [15] M. Reetz, in: Nanoparticles and Catalysis, (Ed.: D.
2004; b) V. V. Rostovsev, L. G. Green, V. V. Fokin, Astruc), Wiley-VCH, Weinheim, 2008, Chap. 8, p 253.
K. B. Sharpless, Angew. Chem. 2002, 114, 2708; Angew. [16] E.-H. Riu, Y. Zhao, Org. Lett. 2005, ##7##6, 1035.

2154 asc.wiley-vch.de  2009 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim Adv. Synth. Catal. 2009, 351, 2147 – 2154

You might also like