You are on page 1of 7

Journal of Materials Chemistry

Cite this: J. Mater. Chem., 2012, 22, 2279 www.rsc.org/materials

View Article Online / Journal Homepage / Table of Contents for this issue

Dynamic Article Links <

PAPER

Ligand effects on the air stability of copper nanoparticles obtained from organometallic synthesis
Published on 16 December 2011. Downloaded by Mahidol University on 16/07/2013 08:34:23.

Cl ement Barri ere,a Kilian Piettre,ab Virginie Latour,a Olivier Margeat,a C edric-Olivier Turrin,a c ad Bruno Chaudret* and Pierre Fau*
Received 3rd October 2011, Accepted 22nd November 2011 DOI: 10.1039/c2jm14963j Low-polydispersity copper nanoparticles (NPs) are prepared through hydrogenolysis of various organometallic copper precursors in an organic medium at moderate temperature. The effects of the precursor composition and the nature of the additional surfactants on the structure and stability of NPs are characterized by TEM and UV-Vis analysis. The improved air stability of copper NPs originating from amidinate copper and stabilized by an alkylamine compound (hexadecylamine) is evidenced and compared with the effect of a long chain carboxylic acid (oleic acid).

Introduction
Nanosized metal particles synthesis is a very active research area since such objects may display enhanced properties derived from both quantum connement effects and high surface to volume ratio. With regards to their high chemical activity and air sensitivity, nanoparticles are preferably prepared with noble metals like gold, silver, platinum1 or more directly as oxide compounds. Because copper is prone to oxidation, the literature on copper nanoparticles (NPs) synthesis is less developed than for other metals. Copper is nonetheless a versatile material which nds various applications in various elds of research like advanced microelectronics for conductive lines, medicine as an antibacterial agent, catalytic and biocatalytic transformations2 or for surface enhanced Raman spectroscopy applications.3 The chemistry of copper nanoparticles is dominated by synthetic procedures based on the temperature controlled decomposition of copper salts in water, followed by a reduction step (sodium borohydrate, hydrazine), stabilization of the metal clusters inside protective shells (reverse micelles, polymers or long alkane chains) and eventually transfer into an organic phase.4 Besides these methods, there are also examples of formation of copper nanoparticles by gamma irradiation,5 microwave,6 electron beam,7 thermal decomposition,8 vapor deposition,9 or chemical reductions.10 However, the resulting nanoparticles were reported with diameters ranging between 3 and 100 nanometres, with a broad size distribution. Size dispersity is a key parameter to
a LCC-CNRS, 205 route de Narbonne BP44099, 31077 Toulouse Cedex 4, France. E-mail: pierre.fau@lcc-toulouse.fr b STMicroelectronics, rue Pierre et Marie Curie BP 7155, 37071 Tours Cedex 2, France c LPCNO, CNRS, INSA, 135 avenue de Rangueil, 31077 Toulouse Cedex 4, France. E-mail: chaudret@insa-toulouse.fr d  de Toulouse, Universite  Paul Sabatier, 31077 Toulouse, France Universite

obtain reproducible and controllable chemical and physical properties of nano-objects. In the most favorable cases the standard deviation of the size distribution is presently 30% of the average diameter of nanoparticles. Bunge et al.11 have prepared coinage (Cu, Ag, Au) metal nanoparticles through the reduction of a coinage-mesityl precursor injected into pure organic solvent (hexadecylamine) at 300  C. This method was formerly developed by Fischer and co-workers.12 In all cases, the control on both the size and the size dispersion of copper nanoparticles is a pivotal issue, and these nanoparticles are subjected to the action of oxygen from ambient air and undergo at least a partial surface oxidation. The strategy based on the reduction of an organometallic precursor in an organic solution by means of a reducing gas (hydrogenolysis) is a procedure that has been developed in our group in the early nineties.13 This soft chemistry method offers a wide range of multiple parameters including temperature, ligand content and nature, and reducing gas, which have strong inuences on the size, shape and surface of nanoparticles for a large variety of metals.14 Previous work of our group has reported the preparation of Cu NPs from the precursor CpCutBuNc.15 Moreover we recently described the possible use of [Cu(Mes)]5 for the preparation of copper lms as well as preliminary results for the synthesis of NPs. Here we extend this study to various precursors and ligands in order to evaluate the possibility of modulating size, shape, and oxidation ability of NPs.16 The present paper describes the use of various precursors for the formation of highly controlled copper colloidal solutions obtained from low temperature hydrogenolysis in an organic solvent of copper(I) and copper(II) organic precursors. The resulting copper nanoparticles of this study are characterized by imaging (TEM) and spectroscopic (UV-Vis) techniques. According to synthesis conditions (precursor, temperature, solvent, and stabilizer), the mean size, size dispersion and air stability are described and discussed.
J. Mater. Chem., 2012, 22, 22792285 | 2279

This journal is The Royal Society of Chemistry 2012

View Article Online

Materials and reagents


Isobutyrate copper(II) and amidinate copper(I) were respectively supplied by Sigma Aldrich and Nanomeps S.A. and used without further purication. Mesityl copper(I) precursor was either synthesized in our laboratory or purchased from Nanomeps S.A. Alkylamines (ethylenediamine C2H4(NH2)2, EDA; octylamine CH3(CH2)7NH2, OA; dodecylamine CH3(CH2)11NH2, DDA; hexadecylamine CH3(CH2)15NH2, HDA), and carboxylic acid (oleic acid C18H34O2, OLA) were received from Sigma Aldrich and used without further purication. Toluene and THF solvents were obtained by freezing followed by vacuum pumping in a Schlenk vessel. All precursors and surfactants were stored in a glove box in order to avoid any traces of moisture or air contamination of the chemicals. All experiments were performed with a vacuum/argon ramp and glassware was oven-dried at 110  C prior to use. Fisher Porter techniques have been employed for all reduction steps involving H2 gas which was purchased from Air Liquide (Alphagaz 1 grade, purity 99.999%).

Published on 16 December 2011. Downloaded by Mahidol University on 16/07/2013 08:34:23.

carbon-covered copper grid. The TEM experiments were performed on a JEOL JEM 1011 transmission electron microscope  High resolution operating at 100 kV with a resolution of 4.5 A. TEM (HR-TEM) analysis was carried out using a JEOL JEM 2100F eld emission gun transmission electron microscope  (FEG-TEM) operating at 200 kV with a resolution of 2.3 A  (point-point) and 1 A (line). Scanning electron microscopy (SEM) analysis was carried out using a JEOL JSM 6700F microscope operated with a 0.530 kV scale acceleration voltage. The UV-Visible spectra of copper colloidal solutions were obtained on a Perkin-Elmer Lambda 25 spectrophotometer, at room temperature, between 400 and 800 nm and using a quartz cell (L 1 cm). Samples were prepared by adding 40 mL of the copper colloidal solution to 2 mL of the solvent inside the quartz cell.

Results and discussion Synthesis of copper nanoparticles


In a typical synthesis of copper nanoparticles, a copper precursor (0.18 mmol) was weighed in a glove box atmosphere and introduced into a Fisher Porter reactor. The solvent (5 mL) was added to the copper precursor and gently stirred until full dissolution of the compound. The alkylamine and/or carboxylic acid ligands were separately weighed and introduced into the as-prepared homogeneous copper solution. The Fisher Porter reactor was then tightly closed and connected to the argon/vacuum ramp. All tubing was purged under vacuum/argon cycles before the reactor was pressurized with 3 bars H2. Finally, the reactor was placed in an oil bath and the temperature was raised to the desired level. The reaction was left running overnight. The resulting colloidal solutions were generally red or deep red and some slight copper metallization of the reactor walls was observed only for very low ligands molar content (<0.5 eq.). The copper nanoparticles synthesized in this study have been prepared from various organometallic copper precursors including isobutyrate copper Cu[CHCOO(CH3)2)]2 (1), mesityl copper Cu5[C9H11]5 (2) and (N,N0 -diisopropylacetamidinato) copper Cu2[(i-Pr NH)2(CHCH3)]2 (3). The physical properties of selected copper precursors have made them suitable for vapor phase deposition processes (CVD). This is the case for compound 3 which has recently been highlighted by R. G. Gordon et al. as a precursor of choice for the Atomic Layer Deposition (ALD) of thin copper lms.17 Whereas volatility and stability are key parameters in these vapor processes in order to insure a constant mass transport, these parameters are of low interest for our liquid phase decomposition process. The ligands present in the coordination sphere of the precursor may play a role in the stabilization of the resulting NPs, either directly, or upon reaction with the other species in the solvent. In the case of copper isobutyrate and copper amidinate, the organic moieties released in the medium during decomposition, i.e. isobutyric acid and ethanediamine (respectively hydrogenated isobutyrate and amidinate moiety), can play a role as active ligands for the metal nanoparticles in solution. All colloidal solutions obtained from these precursors are noted col. x (x 1 to 12) in the present

Instrumentation
Transmission electron microscopy (TEM) specimens were prepared under the argon atmosphere of a glove box by slow evaporation of a drop of colloidal solution deposited onto a thin

Table 1 Resume of synthesis conditions, and nanoparticles size and dispersion obtained from this study Synthesis Col. 1 Col. 2 Col. 3 Col. 4 Col. 5 Col. 6 Col. 7 Col. 8 Col. 9 Col. 10 Col. 11 Col. 12 Copper precursor 1 1 2 2 2 2 2 3 3 3 3 3 Solvent Toluene Anisole THF THF THF Toluene Anisole THF Toluene Toluene Toluene Toluene Temperature 110  C 150  C 80  C 80  C 80  C 110  C 150  C 80  C 110  C 110  C 110  C 110  C Eq. ligand 2 OLA, 2 HDA 0.1 OA 0.1 EDA 0.05 DDA 0.1 DDA 2 OLA, 1 HDA 0.1 HDA 0.1 DDA 0.1 HDA 0.5 HDA 0.5 OLA Nanoparticles 3.8 nm, s 0.6 nm 2.17 nm, s 0.44 nm 550 nm poor stabilization 550 nm poor stabilization 7.8 nm, s 1.8 nm 5 nm and 10 nm 1.8 nm, s 0.3 nm 8.89 nm, s 1.4 nm 7.9 nm, s 2.5 nm 6.9 nm, s 1.8 nm 8.4 nm, s 0.9 nm Bimodal 2.5 nm, s 0.49 nm 8 nm, s 1.11 nm

2280 | J. Mater. Chem., 2012, 22, 22792285

This journal is The Royal Society of Chemistry 2012

View Article Online

study, and the basic information related to these colloidal systems are gathered in Table 1. Copper isobutyrate Copper isobutyrate is a rather stable copper(II) compound and is expected to react at higher temperature than copper(I) precursors in the presence of dihydrogen. The decomposition temperature is an important parameter during the formation of copper colloids. According to the reaction temperature, nucleation or growth processes can be favored and lead to different particle sizes and size distributions. The decomposition of compound 1 in solution releases 2 molar equivalents (eq.) of a carboxylic acid (butanoic acid) which can act as a capping agent according to Scheme 1. Isobutyrate copper was rst decomposed in toluene at 110  C without any further stabilizing agent (col. 1). The initial green solution turns red after a few hours, indicating the formation of copper(0) nanoparticles. Since the reactor walls are slightly covered with a bright copper lm, the stabilization of the copper particles is therefore only partial under these conditions. The TEM image of these particles is shown in Fig. 1a. The particle size distribution is quite narrow (mean size is 3.8 nm, s 0.6 nm, inset of Fig. 1a), and their shape presents multiple edges. The particles are well separated and neither agglomeration nor coalescence occurs on the carbon grid, indicating that the carboxylic acid moiety is strongly adsorbed on the copper surface. The UV-Vis spectrum of this solution is shown in Fig. 1b. A smooth plasmon peak can be evidenced at 572 nm which is typical for nanosized copper surface plasmon resonance (SPR).18 Upon air exposure of the sample, the plasmon peak is slightly red-shifted (up to 580 nm) and its intensity decreases rapidly as the nanoparticles surface is oxidized, resulting in a decrease of the metallic core size as well as an aggregation/precipitation of the particles. In a second experiment, a mixture of extra ligands OLA and HDA (2 equivalent ligand/copper precursor molar ratio: 2 eq.) is added to the precursor, and the decomposition is run in anisole solution at 150  C (col. 2). The combination of these long chain

Published on 16 December 2011. Downloaded by Mahidol University on 16/07/2013 08:34:23.

Fig. 2 (a) TEM image of col. 2 (OLA 2 eq., HDA 2 eq.) and particle size distribution (inset), and (b) UV-Vis spectra of col. 2 before (t 0) and as a function of time exposure to air (t min).

acid and amine ligands has been found to allow the growth of anisotropic structures (nanorods) in the case of cobalt precursor decomposition.19 These parameters (high temperature, high ligands content) allow the formation and stabilization of small isotropic copper nanoparticles (mean size 2.17 nm, s 0.44 nm), as observed by TEM imaging (Fig. 2a). On the carbon grid the particles are disposed as an organized monolayer with interparticle distances of ca. 2 nm, reecting the length of the aliphatic part of the stabilizing ligands (2.3 nm for HDA and 2.6 nm for OLA). The UV-Vis spectrum of col. 2 (Fig. 2b) is even smoother than the one of col. 1. Almost no plasmon peak can be seen and only a weak shoulder indicates its presence. This can be a consequence of the particle size (down to 2 nm) for which the plasmon effect vanishes.20 At such a particle size, Liz-Marzan describes the possible interband transitions that can efciently damp surface plasmon resonances through dephasing of the optical polarization associated with the electron oscillation.21 Upon air exposure, the decrease of the signal which has already been mentioned for col. 1 solutions may also be due to the oxidation and aggregation/ precipitation of the particles in solution. Copper isobutyrate is a very stable organometallic precursor and requires rather high temperatures (110  C) to be decomposed in toluene to form NPs. The release of butanoic acid as a byproduct favors the formation of small copper nanoparticles, but even in the presence of a high content of additional ligands, the air stability of these colloidal solutions is limited to a few minutes.

Scheme 1 Hydrogenolysis of the carboxylic acid copper precursor.

Mesityl copper Mesityl copper has been prepared according to the procedure described by Tsuda.22 The copper precursor appears as a pale yellow powder which has been characterized by single crystal X-ray diffraction which conrmed the presence of a crystalline pentameric phase, as previously reported.23 The hydrogenolysis of mesityl copper 2 in solution starts at 50  C. We have already reported on the modication of mesityl copper precursor in

Fig. 1 (a) TEM image of col. 1 (no extra ligand) and particle size distribution (inset), and (b) UV-Vis spectra of col. 1 before (t 0) and as a function of time exposure to air (t min).

Scheme 2 From copper mesityl to amino-copper complexes as copper nanoparticle precursors.

This journal is The Royal Society of Chemistry 2012

J. Mater. Chem., 2012, 22, 22792285 | 2281

View Article Online

Published on 16 December 2011. Downloaded by Mahidol University on 16/07/2013 08:34:23.

Fig. 3 (a) TEM image of col. 3 (OA 0.1 eq.), (b) TEM image of col. 4 (EDA 0.1 eq.), and (c) UV-Vis spectra of col. 4 before (t 0) and as a function of time exposure to air (t min).

solution in the presence of primary or secondary amines. The reaction proceeds smoothly and quantitatively to form new amino-copper precursors.16 This reactivity has been used to produce copper nanoparticles by in situ hydrogenolysis of pre-formed amino-copper precursors from copper mesityl in toluene solutions (Scheme 2). When a very low fraction (down to 0.1 eq.) of a short aliphatic chain amine agent (EDA or OA) is added to the THF reaction medium a clear stabilization effect is observed. After 8 hours of reaction at 80  C, the solution displays a red wine color and a colloidal aspect (col. 3 OA 0.1 eq., col. 4 EDA 0.1 eq.). In both cases, TEM imaging reveals the presence of rather small (5 nm) nanoparticles along with larger ones (50 nm) with large size dispersity (Fig. 3a and b). No control of the particle size could be achieved under these synthesis conditions. Moreover, the kinetic stability of this solution with time is poor since the solution starts to settle after a few days of storage under an inert atmosphere. The stability of the colloidal solutions can be related to several reaction parameters. Firstly, the low quantity of a ligand which is not tightly bound at the surface but can exchange with free amine ligands in solution may not be sufcient to stabilize the nanoparticles and cannot prevent their rapid aggregation. Secondly, THF is also a weak coordinating ligand, and its possible adsorption on the surface of the nanoparticles is most probably reversible, leaving free surface areas allowing growth and coalescence of the particles. The UV-Vis spectra of col. 4 show a rather intense SPR peak centered at 586 nm. This intense signal is due to the presence of large crystals (ca. 50 nm) of copper in the solution.24 When exposed to air, the peak shifts within a few minutes to 590 nm and then vanishes in a few hours, because of oxidation processes followed by aggregation and sedimentation of the large particles.

The red shift during oxidation is attributed to several factors including a modication of the dielectric constant of the material surrounding the Cu domains. The growth of a Cu2O shell also greatly modies the interactions between the copper core and the coordinated ligands and/or the solvent on the surface.25 Amines with a longer alkyl chain like DDA (0.05 eq.) offer a much better stabilization under the same decomposition conditions (col. 5). The mean particle size is 7.8 nm (s 1.8 nm), the shape of the particles tends to be homogeneously spherical and no aggregation is evidenced by TEM imaging (Fig. 4). Compared to col. 3 and col. 4 this result indicates the importance of the alkyl chain length for the stabilization efciency, by sterically preventing the aggregation of metallic cores. It is known26 that an alkylamine molecule is weakly bound to the metal surface thanks to the electron pair of the nitrogen atom, and is in rapid exchange with the solvent. The aliphatic moiety is pointing out from the metal surface and the resulting steric repulsion is the major mechanism for the effective particles dispersion. Shorter aliphatic chains like EDA (C2) or OA (C8) result in a less effective protection effect and therefore lead to the rapid coalescence of copper NPs through Ostwald ripening. This phenomenon is even amplied by a low ratio of alkylamine surfactants. Good particle stabilization requires a sufcient excess of available ligands in the solution due to the continuous exchange of the weak coordinating ligand with the metal surface. When the hydrogenolysis temperature is raised to 110  C (toluene solution), with the same DDA ligand (0.1 eq.), the stabilization of the particles is less efcient and a large distribution of particle sizes from 2 to 15 nm is obtained (col. 6). With such reaction conditions, the TEM image reveals a population centered around 5 nm and a larger one around 10 nm (Fig. 5). This distribution can be due to the distinct nucleation and growth of the amino-copper fraction aside from the mesitylcopper one. Actually, one can expect that such a low ligand level

Fig. 4 TEM image of col. 5 (DDA 0.05 eq.) and particle size distribution.

Fig. 5 TEM image of col. 6 (DDA 0.1 eq.) and particle size distribution.

2282 | J. Mater. Chem., 2012, 22, 22792285

This journal is The Royal Society of Chemistry 2012

View Article Online

Published on 16 December 2011. Downloaded by Mahidol University on 16/07/2013 08:34:23.

Fig. 6 (a) TEM image of col. 7 (OLA 2 eq., HDA 1 eq.) and particle size distribution (inset), and (b) UV-Vis spectra of col. 7 before (t 0) and as a function of time exposure to air (t min).

Fig. 7 (a) TEM image of col. 8 (HDA 0.1 eq.), particle size distribution and higher magnication (insets) and (b) UV-Vis spectra before (t 0) and as a function of time exposure to air (t min).

does not favor the stabilisation of homogeneously distributed copper particles. In the presence of a large quantity of coordinating agents (OLA 2 eq., HDA 1 eq.) and when the reaction temperature is increased up to 150  C in anisole, the nucleation of the particles is favored versus their growth in the reaction medium (col. 7, Fig. 6). The TEM grid reveals a regular pavement of small copper clusters (mean size 1.8 nm, s 0.3 nm). Their shape is rather irregular and many edges on the particles are evidenced. This is similar to the aspect of col. 2 and can be related to the high content of carboxylic acid in the reaction solution. This acid content must be responsible for a nucleation/corrosion mechanism for the copper nanoparticles. A similar phenomenon has already been evidenced in the case of the growth of iron nanoparticles.27 As for col. 2, the UV-Vis analysis of such small particles only displays a broad shoulder as a reection of SPR of copper particles and a very rapid decay in intensity upon air exposure (Fig. 6b). Copper mesityl allows the formation of a variety of copper colloids according to different decomposition temperatures and ligand contents. The low temperature decomposition (80  C) of this copper precursor makes possible the controlled growth of homogeneous copper particles with low ligand content (down to 0.05 eq.). At a higher temperature (110  C) the colloidal stability of such solutions is not compatible with low ligand ratios. The use of high quantities of mixtures of amine and acid ligands (up to 2 eq.) allows the formation of small size copper nanoparticles, but they remain highly sensitive to air oxidation.

Amidinate copper Beside mesityl copper, amidinate copper is a copper(I) organometallic complex, which is easy to reduce. Moreover, diamine species will be released in the reaction medium during the decomposition process and may provide a source of an active capping agent for the NPs stabilization according to Scheme 3.

The decomposition of copper amidinate in THF at 80  C in the presence of a small amount of extra ligands (0.1 eq.) produces an effective stabilization of the nanoparticles (col. 8, Fig. 7a). Low HDA levels (0.1 eq.) have proven to allow the stabilization of a homogeneous collection of spherical copper particles (mean size: 8.89 nm, s 1.4 nm). This high stabilization effect must result from a synergetic phenomenon due to the combination of the diamine by-product released by the precursor decomposition and a small content of additional ligands. In this case we obtain a SPR signal centered at 575 nm which indicates the presence of metallic copper particles and the solution is light red (Fig. 7b). Upon air exposure of the solution the absorbance peak intensity increases markedly (rise by a factor 1.3 after 17 h) while the usual red-shift due to oxidation of the particle shell is observed. Saunders et al.28 have reported a similar peak exaltation during the oxidation process of oleic acid capped copper nanocrystals in toluene. These authors propose that solvents with p-bonds strongly interact with electrons of the copper surface, resulting in a damping effect of the surface plasmon resonance. Upon oxidation p-bonds interaction is shielded, the damping effect is lost and the resonance from the copper core sharply increases. They also observed the plasmon peak decrease after only a few minutes of exaltation. In our case, THF does not contain p-bonds so that the observed damping effect may be linked to both HDA and diamine ligands coordinated around the metal core. This combination seems to form a very effective protection against the growth rate of the oxidized shell on the particle, since the exaltation effect lasts more than 17 hours. The solution nally turns green after 24 hours of oxidation, indicating the presence of a thick Cu2O shell around the copper core.28 In toluene at 110  C and with 0.1 eq. DDA (col. 9) or HDA (col. 10) the decomposition of copper amidinate leads to a slight

Scheme 3 Amidinate copper precursor hydrogenolysis.

Fig. 8 TEM image and particle size distribution (inset) of (a) col. 9 (DDA 0.1 eq.) and (b).col. 10 (HDA 0.1 eq.).

This journal is The Royal Society of Chemistry 2012

J. Mater. Chem., 2012, 22, 22792285 | 2283

View Article Online

Published on 16 December 2011. Downloaded by Mahidol University on 16/07/2013 08:34:23.

Fig. 9 (a) TEM image of col. 11 (HDA 0.5 eq.) and particle size distribution (inset), and (b) UV-Vis spectra of col. 11 before (t 0) and as a function of time exposure to air (t min).

metallization of the reactor walls and to the formation of a red wine colored colloidal suspension (Fig. 8). The mean particle sizes of col. 9 and col. 10 are respectively 7.9 nm (s 2.5 nm) and 6.9 nm (s 1.8 nm). Increasing the alkyl chain length from DDA to HDA leads to a signicant decrease in the metallization level of the reactor walls. This indicates that the presence of a longer alkyl chain allows a better stabilization of the particles, as previously observed with the mesityl copper precursor. An increase of the ligand content produces an even better stabilization effect as evidenced by the narrower particle size distribution (HDA 0.5 eq., col. 11, Fig. 9a) even if there is not a direct relation to the mean particle size. In this case, we have obtained nanoparticle sizes centered at 8.4 nm (s 0.9 nm). These synthesis conditions allow very narrow particle size dispersions with a ratio of sigma/mean size as low as 20%. The UV-Vis spectrum of this copper colloid shows a quite large SPR band centered at 569 nm and the red solution is even clearer compared to col. 8 with THF solvent (Fig. 9b). Upon air exposure of the solution, the SPR intensity increases by a factor close to 3 after 24 hours and the solution presents then a deep red color (Fig. 9b). Similarly to the previous case with THF, the observed SPR exaltation must be linked to the partial oxidation of the copper surface. Initially the SPR damping effect from toluene pbonds and amine coordination participates in the low SPR intensity as described by Saunders et al.28 The observed SPR increase and the darkening of the colloidal solution are related to the shielding of the copper core from direct interaction with solvent and amine ligands by the Cu2O shell growth. In addition, SPR intensity is very sensitive to the dielectric constant (3s) of the particle shell since an increase of 3s leads to a net increase in the SPR absorbance.29 Initially the copper surface is surrounded by alkylamine species with a low dielectric constant close to 2.7 and after oxidation the Cu2O shell presents a higher dielectric constant around 9.30 Moreover, an oxidized surface is an even better electron acceptor for the amine donor. It must result in a denser amine shell coordinated to the oxidized copper surface. This denser capping layer forms an organic protective shell that slows down the oxidation rate of the particle. The use of 0.5 eq. carboxylic acid (OLA) instead of alkylamine as ligand under the same decomposition conditions (toluene 110  C) leads to a stable dark red solution within a few hours. A distinct bimodal particle distribution (col. 12, Fig. 10a) is evidenced from TEM analysis. The smallest particles are centered at 2.5 nm whereas the largest ones are close to 8 nm with respective size dispersions of 18% and 14%. This result indicates the
2284 | J. Mater. Chem., 2012, 22, 22792285

differential stabilization effect from two distinct capping agents. The carboxylic acid is a strong stabilizer for copper NPs and thus induces the formation of very small copper clusters. When no more acid is available in the solution for further stabilization, the initial copper clusters are allowed to grow through an Ostwald ripening before being stabilized by the diamine compound released by the precursor decomposition. In order to support this mechanism, it can be observed that only larger particles are subjected to a certain level of coalescence as a proof of the amine weak coordination on the metal surface. This phenomenon has been observed previously in our group in the case of the stabilization of ruthenium or platinum nanoparticles by amine ligands.26 Worm-like objects resulting from the coalescence of spherical nanoparticles are typical for labile ligand interactions with the NPs surface. The UV-Vis spectrum (Fig. 10b) reveals an initial plasmon peak at 570 nm attributed to metallic copper particles. The solution is dark red compared to other cases (col. 8 and col. 11). This deep color suggests the possible presence of a thin oxidized shell around the particles (carboxylic acid coordinated on the copper surface) which protects the copper surface from SPR damping. When the sample is exposed to air, the peak position shifts to 589 nm and is accompanied by a modest increase in intensity. The p-bonds damping effect due to toluene solvent as described by Saunders et al.28 appears only slightly in the present case since there is a weak peak increase before the plasmon decay (t1 min). After 16 minutes, the SPR signal is already shifted to 611 nm, and it decreases continuously in intensity before it disappears within a few hours. During this process, the initial red wine solution turns green after a few minutes. This evolution in the presence of carboxylic acid ligands suggests the rapid reconstruction of the copper surface due to an acidication of the medium in the presence of humidity intake during air exposure. This mechanism is supported by the very rapid formation of a clear green solution (possible formation of carboxylate Cu2+ complexes and/or Cu2O nanoparticles). These results demonstrate the importance of the choice of both the precursor and the coordinating ligands for the stabilization of colloidal copper nanoparticles. The choice of amidinate copper in addition to extra ligands results in the formation of lowpolydispersity copper nanoparticles. In contrast to the effect of carboxylic acid ligands, the use of alkylamine ligands allows the slowest growth of the oxidation shell upon air exposure of copper nanoparticles.

Fig. 10 (a) TEM image of col. 12 (OLA 0.5 eq.) and particle size distribution (inset), and (b) UV-Vis spectra before (t 0) and as a function of time exposure to air (t min).

This journal is The Royal Society of Chemistry 2012

View Article Online

Conclusion
In summary, we have demonstrated the interest of the organometallic hydrogenolysis for the controlled synthesis of highly pure copper NPs. High temperature (150  C) decomposition of copper precursors (carboxylate or mesityl copper) in anisole favors a high nucleation rate in solution and therefore leads to the formation of very small copper nanoparticles (ca. 2 nm), highly sensitive to air oxidation. The use of more reactive precursors like amidinate copper allows decomposition at lower temperature and the formation of larger copper nanoparticles (ca. 8 nm) with a good control of size dispersion (ratio s/mean size down to 20%). The oxidation of the particles can be followed by the evolution of their surface plasmon resonance under UVVis excitation. We have evidenced a strong damping effect of the plasmon intensity in the case of the stabilization with alkylamine ligands versus long chain carboxylic acids. This damping effect decreases when the colloidal solutions are exposed to ambient air and the plasmon peak increases by a factor of 3 before it vanishes with time. Finally, oleic acid is an efcient ligand for the stabilization of bare copper nanoparticles but is less efcient in order to prevent copper oxidation. The precursor composition is also of primary importance for the air stability of the copper nanoparticles. In the case of the amidinate precursor, a synergetic effect of the released diamine moiety from the precursor with the additional HDA ligand is suggested since the slowest growth rate of copper oxide shell from this study has been obtained. A more detailed characterization of the role of ligands on the copper surface upon air oxidation is currently under investigation by liquid NMR techniques.
3 4 5 6 7 8 9 10 11 12 13 14

15 16 17 18 19 20 21 22 23 24 25

Acknowledgements
This work has been developed thanks to the French S2E2 competitiveness cluster (r egions Centre-Limousin) and the S eSAME CAPI project. We thank FCE, DGE, ANRT and ST Microelectronics for nancial support. We also thank CNRS and Universit e Paul Sabatier (Toulouse III).

Notes and references


1 A. R. Tao, S. Habas and P. D. Yang, Small, 2008, 4, 310. 2 (a) Y. F. Wang, A. V. Biradar, G. Wang, K. K. Sharma, C. T. Duncan, S. Rangan and T. Asefa, Chem.Eur. J., 2010, 16, 10735; (b) H. Meng, Z. Chen, G. M. Xing, H. Yuan, C. Y. Chen, F. Zhao, C. C. Zhang and Y. L. Zhao, Toxicol. Lett., 2007, 175, 102; (c) A. Luisier, I. Utke, T. Bret, F. Cicoira, R. Hauert, S. W. Rhee, P. Doppelt and P. Hoffmann, J. Electrochem. Soc.,

26 27 28 29 30

2004, 151, C535; (d) Y. H. Wei, S. Chen, B. Kowalczyk, S. Huda, T. P. Gray and B. A. Grzybowski, J. Phys. Chem. C, 2010, 114, 15612. J. Cejkova, V. Prokopec, S. Brazdova, A. Kokaislova, P. Matejka and F. Stepanek, Appl. Surf. Sci., 2009, 255, 7864. S. Magdassi, M. Grouchko and A. Kamyshny, Materials, 2010, 3, 4626. S. I. Bin Ahmad, M. S. B. H. Ahmad and S. Bin Radiman, AIP Conf. Proc., 2009, 1136, 186. M. Blosi, S. Albonetti, M. Dondi, C. Martelli and G. Baldi, J. Nanopart. Res., 2011, 13, 127. L. Q. Pham, J. H. Sohn, J. H. Park, H. S. Kang, B. C. Lee and Y. S. Kang, Radiat. Phys. Chem., 2011, 80, 638. Y. Wei, S. Chen, B. Kowalczyk, S. Huda, T. P. Gray and B. A. Grzybowski, J. Phys. Chem. C, 2010, 114, 15612. A. G. Nasibulin, L. I. Shurygina and E. I. Kauppinen, Colloid J., 2005, 67, 1. A. Sarkar, T. Mukherjee and S. Kapoor, J. Phys. Chem. C, 2008, 112, 3334. S. D. Bunge, T. J. Boyle and T. J. Headley, Nano Lett., 2003, 3, 901. J. Hambrock, R. Becker, A. Birkner, J. Weiss and R. A. Fischer, Chemical Communications, Cambridge, UK, 2002, p. 68. A. Duteil, R. Queau, B. Chaudret, R. Mazel, C. Roucau and J. S. Bradley, Chem. Mater., 1993, 5, 341. (a) B. Chaudret, Topics in Organometallic Chemistry, 2005, vol. 16, p. 233; (b) C. Desvaux, F. Dumestre, C. Amiens, M. Respaud, P. Lecante, E. Snoeck, P. Fejes, P. Renaud and B. Chaudret, J. Mater. Chem., 2009, 19, 3268; (c) O. Margeat, F. Dumestre, C. Amiens, B. Chaudret, P. Lecante and M. Respaud, Prog. Solid State Chem., 2006, 33, 71. D. de Caro, V. Agelou, A. Duteil, B. Chaudret, R. Mazel, C. Roucau and J. S. Bradley, New J. Chem., 1995, 19, 1265. C. Barriere, G. Alcaraz, O. Margeat, P. Fau, J. B. Quoirin, C. Anceau and B. Chaudret, J. Mater. Chem., 2008, 18, 3084. Z. Li, A. Rahtu and R. G. Gordon, J. Electrochem. Soc., 2006, 153, C787. M. Singh, I. Sinha, M. Premkumar, A. K. Singh and R. K. Mandal, Colloids Surf., A, 2010, 359, 88. D. Ciuculescu, F. Dumestre, M. Comesana-Hermo, B. Chaudret, M. Spasova, M. Farle and C. Amiens, Chem. Mater., 2009, 21, 3987. T. Ghodselahi, M. A. Vesaghi and A. Shaekhani, J. Phys. D: Appl. Phys., 2009, 42, 015308/1. I. Pastoriza-Santos, A. Sanchez-Iglesias, B. Rodriguez-Gonzalez and L. M. Liz-Marzan, Small, 2009, 5, 440. T. Tsuda, T. Yazawa, K. Watanabe, T. Fujii and T. Saegusa, J. Org. Chem., 1981, 46, 192. H. Eriksson and M. Hakansson, Organometallics, 1997, 16, 4243. D. Mott, J. Galkowski, L. Wang, J. Luo and C.-J. Zhong, Langmuir, 2007, 23, 5740. P. Kanninen, C. Johans, J. Merta and K. Kontturi, J. Colloid Interface Sci., 2008, 318, 88. E. Ramirez, L. Erades, K. Philippot, P. Lecante and B. Chaudret, Adv. Funct. Mater., 2007, 17, 2219. L.-M. Lacroix, S. Lachaize, A. Falqui, M. Respaud and B. Chaudret, J. Am. Chem. Soc., 2009, 131, 549. K. P. Rice, E. J. Walker, M. P. Stoykovich and A. E. Saunders, J. Phys. Chem. C, 2011, 115, 1793. T. Ghodselahi and M. A. Vesaghi, Phys. B, 2011, 406, 2678. W.-Y. Yang, W.-G. Kim and S.-W. Rhee, Thin Solid Films, 2008, 517, 967.

Published on 16 December 2011. Downloaded by Mahidol University on 16/07/2013 08:34:23.

This journal is The Royal Society of Chemistry 2012

J. Mater. Chem., 2012, 22, 22792285 | 2285

You might also like