Вы находитесь на странице: 1из 13

Polymer 45 (2004) 74497461 www.elsevier.

com/locate/polymer

Block copolysiloxanes and their complexation with cobalt nanoparticles


M.L. Vadalaa, M. Rutnakornpituka, M.A. Zalicha,b, T.G. St Pierreb, J.S. Rifea,*
a

Department of Chemistry and the Macromolecules and Interfaces Institute, Virginia Polytechnic Institute and State University, 2018 Hahn Hall, Mail Code 0212, Blacksburg, VA 24061-5976, USA b School of Physics, The University of Western Australia, Crawley, WA 6009, Australia Received 1 December 2003; received in revised form 11 August 2004; accepted 31 August 2004

Abstract Block copolysiloxanes have been prepared and utilized to form complexes of cobalt nanoparticles encased in the copolymers. The coated nanoparticles could be dispersed in polydimethylsiloxane (PDMS) to afford PDMS ferrouids. Poly(dimethylsiloxane-b-methylvinylsiloxane) (PDMS-b-PMVS) diblock copolymers were synthesized via anionic living polymerization with controlled molecular weights and narrow molecular weight distributions. The PMVS blocks were functionalized with trimethoxysilethyl or triethoxysilethyl pendent groups to yield poly(dimethylsiloxane-b-(methylvinyl-co-methyl(2-trimethoxysilethyl)siloxane) (PDMS-b-(PMVS-co-PMTMS)) or poly(dimethylsiloxane-b-(methylvinyl-co-methyl(2-triethoxysilethyl)siloxane) (PDMS-b(PMVS-co-PMTES)) copolymers, respectively. Stable suspensions of mostly superparamagnetic cobalt nanoparticles were prepared in toluene in the presence of PDMS-b-(PMVS-coPMTMS) and PDMS-b-(PMVS-co-PMTES) copolymers via thermolysis of Co2(CO)8. TEM micrographs showed non-aggregated cobalt nanoparticles with mean particle diameters ranging from z1015 nm. Specic saturation magnetizations of the cobaltcopolymer complexes ranged from w40110 emu gK1 of cobalt. q 2004 Published by Elsevier Ltd.
Keywords: Polysiloxane; Polymethylvinylsiloxane; Cobalt

1. Introduction Ferrouids are liquid dispersions of ferromagnetic or ferrimagnetic particles, usually in hydrocarbons, esters or water. Polymers or low molecular weight surfactants adsorbed or bonded on the particle surfaces can prevent agglomeration by either electrostatic or steric repulsion. The particles are often ferrimagnetic iron oxides such as magnetite or maghemite since methods for their synthesis in small sizes (e.g. z10 nm in diameter) are established and these materials are relatively stable against oxidation. If particle diameters are in this range, the materials can be superparamagnetic. When such uids are placed in gradient magnetic elds, they typically respond by moving as an entity toward the direction of highest eld [1,2]. When the applied elds are removed, the magnetic moments of the
* Corresponding author. Tel.: C1 540 231 6717; fax: C1 540 231 8517. E-mail address: anynn@vt.edu (J.S. Rife). 0032-3861/$ - see front matter q 2004 Published by Elsevier Ltd. doi:10.1016/j.polymer.2004.09.001

particles randomize rapidly and the net magnetic moment of the uid returns to zero. Our interest has been to design biocompatible, polydimethylsiloxane (PDMS)-based ferrouids for possible use in therapies for treating retinal detachment disorders [36]. Cobalt nanoparticles have inherently higher magnetizations relative to magnetite or maghemite. The specic saturation magnetization of bulk cobalt is 162 emu gK1 while magnetite is about 92 emu gK1 [7,8]. Cobalt nanoparticles can be prepared via chemical reduction of CoCl2 in organic media [9] or reduction with hydrides in inverse micellar media [1012]. Other methods utilize block copolymers that stabilize cobalt nanoparticles in micelle cores. These systems typically involve thermolysis of organo-cobalt precursors [5,6,13,14]. Thermolysis reactions of dicobalt octacarbonyl in the presence of dispersion stabilizers can provide superparamagnetic cobalt nanoparticles free from contamination by side products [1,1518].

7450

M.L. Vadala et al. / Polymer 45 (2004) 74497461

Previous investigations of dicobalt octacarbonyl reactions in our laboratories utilized poly(dimethylsiloxane-b(3-cyanopropyl)methylsiloxane- b-dimethylsiloxane) (PDMS-b-PCPMS-b-PDMS) triblock copolymers in toluene, D4, or PDMS solvents to generate cobalt nanoparticles in situ [5,6,14]. The central block containing the nitriles was only sparingly soluble in the solvents employed, whereas the PDMS endblocks were soluble. The reaction solutions contained micelles with the PCPMS forming the micelle cores and the PDMS endblocks protruding into the solvents. It was reasoned that the lone pairs of electrons on the nitrogens complexed with the cobalt, and the PDMS tails provided steric stabilization of these nanoparticle dispersions. These cobalt dispersions had specic saturation magnetizations of 90110 emu gK1 of cobalt and TEM micrographs showed well-dispersed cobalt nanoparticles [6,14]. The oxidative durability of the cobalt nanoparticles was poor, however, and their magnetization decreased slowly with time under ambient conditions [14,19]. Cobalt carbonyl reactions in solutions of pentablock copolymer steric dispersion stabilizers (Fig. 1) were also investigated to form cobalt nanoparticles [19]. The motivation for this approach was to protect the cobalt surfaces against oxidation with silica-like shells formed around the nanoparticles by hydrolyzing and condensing trialkoxysilyl pendent groups [2]. These reactions resulted in stable dispersions of z10 nm cobalt with specic magnetizations of 90110 emu gK1 of cobalt. The pendent triethoxysilyl groups were hydrolyzed in the presence of dibutyltin diacetate (catalyst) with stoichiometric concentrations of water at room temperature. This paper describes diblock polysiloxane dispersants wherein trialkoxysilyl-functional blocks complexed with cobalt and PDMS blocks provided dispersion stability in toluene. Surprisingly, the trialkoxysilethyl-functional blocks served as anchor segments for the cobalt. These block copolymer dispersion stabilizers are of particular interest because of their relative simplicity as compared to the previously studied pentablock stabilizers. Additionally, either trimethoxysilethyl (more reactive) or triethoxysilethyl pendent groups could be employed to prepare cobalt dispersions with the copolymers described herein, and it was

reasoned that this might allow more latitude in designing post-crosslinking conditions to yield oxygen impermeable shells.

2. Experimental 2.1. Materials Hexamethylcyclotrisiloxane (D3, Gelest, Inc.) was puried by stirring it over calcium hydride at 80 8C overnight and was subsequently fractionally vacuumdistillation under nitrogen into a pre-weighed, amedried ask. The initiator, n-butyllithium, was kindly donated by the Lithco Division of the FMC Corporation and was approximately 2.45 M in cyclohexane. It was titrated with diphenylacetic acid in THF until a yellow color persisted [20] and used as received. 1,3,5-trivinyl1,3,5-trimethylcyclotrisiloxane (DV 3 ; Gelest, Inc.) was fractionally distilled under vacuum into a pre-dried ask, purged with nitrogen, sealed with a septum, and stored in a dessicator. Cyclohexane (Aldrich, 99%) was stirred with concentrated sulfuric acid for 48 h, washed with water, dried over MgSO4, then over sodium, and distilled just prior to use. Tetrahydrofuran (THF, 99.5%, E.M. Sciences) was dried over calcium hydride overnight, then reuxed over sodium in the presence of benzophenone until the solution was a deep purple. The THF was distilled just prior to use. Trimethylchlorosilane (Gelest, Inc.) was used as the terminating reagent for the diblock copolymers, and was distilled before use. A Pt0(1,3-divinyl-1,1,3,3-tetramethyldisiloxane)1.5 complex catalyst in xylene (2.12.4 wt% Pt0, Karstedts catalyst) (Gelest, Inc.) was used as received. Triethoxysilane and trimethoxysilane (Gelest, Inc.) were used as received. Toluene was washed twice with concentrated sulfuric acid and neutralized with water. It was dried over MgSO4 for 1 h, then over calcium hydride overnight and distilled just before use. Co2(CO)8 (Alpha Aesar) stabilized with 15% hexane was stored under argon in the freezer without further purication.

Fig. 1. Pentablock copolymers were investigated as steric dispersion stabilizers for cobalt nanoparticles. The nitrile containing central block functioned as the anchor block to bind to cobalt; PDMS tail blocks provided dispersion stability; PMTES blocks were precursors for solgel reactions to form silica-like shells around the nanoparticles [2].

M.L. Vadala et al. / Polymer 45 (2004) 74497461

7451

2.2. Synthesis 2.2.1. Synthesis of poly(dimethylsiloxane-b-methylvinylsiloxane) diblock copolymers (PDMS-b-PMVS) with controlled molecular weights A synthetic procedure for preparing a diblock copolymer with a targeted number average molecular weight (Mn) of 7000 g molK1 comprised of 5000 g molK1 PDMS and 2000 g molK1 PMVS is provided. A series of diblock copolymers with different molecular weights were prepared by varying the ratio of n-butyllithium to D3 to control the Mn of the PDMS blocks, and by varying the amounts of DV 3 relative to D3 to control the Mn of the PMVS blocks. The rst part of the synthesis involved preparing a PDMS block with one terminal lithium siloxanolate. D3 (15.64 g, 0.003 mol) was distilled into a ame-dried, 250 ml, roundbottom ask equipped with a magnetic stir bar, purged with dry nitrogen and sealed with a septum. Cyclohexane (18 ml) was added via syringe as a solvent to dissolve the D3. Once dissolved, 1.26 ml of a 2.45 M solution of n-butyllithium (0.003 mol) in cyclohexane was charged to the reaction ask via syringe, and stirred for 15 min at 25 8C. THF (18 ml) was added as a promoter and the reaction was allowed to proceed at room temperature until the complete conversion of D3 monomer to polymer as determined by 1H NMR. This required approximately 8 h reaction time for this composition. DV 3 (6.2 ml, 0.003 mol) was charged to the reaction ask and the copolymerization was allowed to proceed at room temperature until conversion reached approximately 95% (approx. 8 h) as measured by 1H NMR. The diblock copolymer was terminated with an excess of trimethylchlorosilane (0.76 ml, 0.006 mol) via syringe and stirred for approximately 30 min. The solution clouded upon addition of the trimethylchlorosilane due to precipitation of insoluble LiCl salt. The polymer solution was diluted with chloroform (300 ml) and placed in a separatory funnel. It was washed with water several times to remove LiCl. Approximately 80% of the chloroform was removed by rotary evaporation and the polymer was twice precipitated and puried by pouring the remaining solution into stirred methanol (200 ml). The clear liquid copolymer sank to the bottom of the vessel. The methanol was decanted and the polymer was dried under vacuum at approximately 500 mTorr at 60 8C for 24 h. 2.2.2. Synthesis of poly(dimethylsiloxane-b-methyl (2-triethoxysilethyl)siloxane] (PDMS-b-PMTES) via hydrosilation Functionalized diblock copolymers were prepared from the PDMS-b-PMVS diblock copolymers. The molar ratios of triethoxysilane or trimethoxysilane relative to the number of vinyl groups determined the degree of hydrosilation. A representative procedure for quantitative hydrosilation is described. Firstly, 0.3 g of the polymer precursor, a 5000 g molK1 PDMS and 2500 g molK1 PMVS diblock copolymer, was weighed into a ame-dried, 6-dram vial

equipped with a magnetic stir bar. The vial containing the copolymer was sealed with a septum under nitrogen. Toluene (10 ml) was added via syringe and the mixture was stirred until the polymer dissolved. Triethoxysilane (0.21 ml, 0.0012 mol) was added via syringe to the reaction vial followed by addition of 10 ml of Pt0 (Karstedts) catalyst in xylene. The reaction was stirred at 60 8C until complete hydrosilation occurred as evidenced by 1H NMR. The solvent was evaporated under vacuum and the PDMSb-PMTES diblock copolymer was stored under nitrogen. 2.2.3. Synthesis of poly(dimethylsiloxane-b-(methylvinylsiloxane-co-methyl(2-trimethoxysilethyl)siloxane)) (PDMSb-(PMVS-co-PMTMS)) via hydrosilation An exemplary synthesis to prepare a diblock copolymer wherein only half the pendent vinyl groups on the polymer precursor were functionalized is provided. Firstly, 0.3 g of the polymer precursor, a 5000 g mol K1 PDMS and 2500 g molK1 PMVS diblock copolymer, was weighed into a ame-dried, 6-dram vial equipped with a magnetic stir bar. The vial containing the copolymer was sealed with a septum under nitrogen. Toluene (10 ml) was added via syringe and the mixture was stirred until the polymer dissolved. Trimethoxysilane (0.08 ml, 0.0006 mol) was added via syringe to the reaction vial followed by addition of 10 ml of the Pt0 (Karstedts) catalyst in xylene. The amount of catalyst was based on z10K3 mol of Pt0 per mole vinyl. The reaction was stirred at 60 8C until complete hydrosilation occurred as evidenced by 1H NMR. The solvent was evaporated under vacuum and the PDMS-b[PMVS-co-PMTMS] diblock copolymer was stored under nitrogen. 2.2.4. Hydrolysis and condensation of poly(dimethylsiloxane-b-methyl(2-triethoxysilethyl)siloxane) (PDMS-bPMTMS) A series of reactions were conducted wherein the concentration of water per equiv. of methoxysilyl groups was varied from stoichiometric (1 mol water:2 mol methoxy), two times stoichiometric (2 mol water:2 mol methoxy), and four times stoichiometric (4 mol water:2 mol methoxy). A solgel hydrolysis and condensation reaction of a PDMS-b -PMTMS block copolymer having a 10,000 g molK1 PDMS block connected to a 4800 g molK1 PMTMS block with four times the stoichiometric concentration of water is described. A 3-neck, 250 ml, roundbottom ask equipped with a mechanical stirrer, condenser, and nitrogen purge was charged with 1.0 g (0.0016 equiv. of trimethoxysilethyl groups, 0.0047 equiv. methoxy) of the PDMS-b-PMTMS copolymer. Toluene (15 ml) was transferred via syringe to the reaction vessel to dissolve the copolymer. Water (0.17 ml, 0.0096 mol, 12 equiv. per trimethoxysilane, four times stoichiometric) was added via syringe. The reaction was stirred at 95 8C. The reaction progress was monitored via 1H NMR.

7452

M.L. Vadala et al. / Polymer 45 (2004) 74497461

Fig. 2. Anionic block copolymerization to obtain diblock PDMS-b-PMVS copolymers.

2.2.5. Synthesis of a cobalt nanoparticle uid in the presence of a PDMS-b-(PMVS-co-PMTES) diblock copolymer A representative reaction is described utilizing a 5300 g molK1 PDMS-b-3900 g molK1 (PMVS-co-PMTES) copolymer. The PMVS-co-PMTES block had 50% of the repeat units hydrosilated with triethoxysilane. Dispersions with other copolymers were prepared in an analogous manner. A 500 ml, 3-neck, roundbottom ask equipped with a condenser, mechanical stirrer with a vacuum ready adapter, and argon purge was ame-dried under argon. The apparatus was placed in a temperature controlled silicone oil bath over a hot plate (without a magnetic stirrer). The PDMSb-(PMVS-co-PMTES) copolymer (1.0 g, 0.0013 equiv. triethoxysilethyl groups) in 10 ml toluene was transferred to the reaction vessel via cannula. An additional 10 ml of toluene was added via syringe to the reaction ask. Dicobalt octacarbonyl (1.0 g, z0.0035 mol) was weighed into the reaction vessel and dissolved under argon. A greenish-brown gas lled the ask immediately upon adding the dicobalt octacarbonyl. The reaction temperature was raised to approximately 110 8C (toluene reux) and maintained at that temperature for approximately 2 h or until the complete disappearance of the Co4(CO)12 intermediate as determined by FT-IR. After cooling, a stable magnetic dispersion of cobalt nanoparticles resulted. 2.3. Characterization of copolymers and copolymer solutions All 1H and 29Si NMR spectra were obtained on a Varian Unity 400 MHz NMR spectrometer operating at 400 and 80 MHz, respectively. The samples for 29Si NMR were

prepared with 0.63 g copolymer, 0.52 g Cr(Acac)3, and 2.4 ml d-CHCl3. Quantitative silicon NMR spectra were obtained with the aid of the relaxation agent, Cr(Acac)3, with a pulse width of 168.08 and a relaxation delay of 10 s. Gel permeation chromatograms were obtained in chloroform at 30 8C on a Waters Alliance Model 2690 chromatograph equipped with a Waters HR 0.5CHR 2CHR 3CHR 4 styragel column set. A Viscotek viscosity detector and a refractive index detector were utilized with polystyrene calibration standards to generate a universal molecular weight calibration curve for absolute molecular weight analyses. Thermal properties of the copolymer precursors (PDMS-b-PMVS) were analyzed by differential scanning calorimetry using a TA Instruments DSC Q1000 under constant ow of helium. The samples (1015 mg) were cycled from K15025 8C using hermetically sealed DSC pans. Two scans were performed on each sample where the Tgs were taken as the inection points on the second scans. 2.4. Characterization of magnetic uids FTIR spectra were obtained using a Nicolet Impact 400 FTIR spectrometer with two drops of the samples run neat between salt plates. Transmission electron micrographs were acquired using a Philips 420T TEM run at 100 kV. TEM samples containing cobalt stabilized with copolymers were prepared by diluting toluene dispersions with additional toluene until they had the appearance of weak tea. A drop was syringed onto a carbon coated copper grid and the toluene was evaporated. A Standard 7300 Series Lakeshore vibrating sample magnetometer (VSM) was used to determine the magnetic properties of the cobalt samples including saturation magnetization and any hysteresis.

M.L. Vadala et al. / Polymer 45 (2004) 74497461

7453

Fig. 3. Gel permeation chromatograms of the PDMS-b-PMVS diblock copolymers conrmed narrow molecular weight distributions, suggesting the living nature of these polymerizations.

The magnetic moment of each dried sample was measured over a range of applied elds from K8000C8000 Oe with a sensitivity of 0.1 emu. A Quantum Design magnetic properties measurement system using a superconducting quantum interference device sensor was used to make measurements of cobalt specic magnetization (s) in applied magnetic elds (H) over the range of K70,000C 70,000 Oe at room temperature and 5 K. Low temperature measurements were made both after cooling the sample in zero applied eld and in an applied eld at 70,000 Oe. These measurements were used to investigate the presence of any: (1) residual cobalt carbonyl species, and (2) cobalt oxide on the surfaces of metallic cobalt particles. Elemental analyses for cobalt were performed by Desert Analytics Laboratory (Tucson, AZ) by treating the samples with hot concentrated nitric acid followed by concentrated perchloric acid until complete dissolution was achieved. These solutions were analyzed by inductively coupled plasma to determine
Table 1 Mns and molecular weight distributions of PDMS-b-PMVS diblock copolymers Targeted Mn (g molK1) 7000 7000 8000 12,000 12,000 16,000 18,000 18,000 18,000 Mn by 1H NMR (g molK1) 6000 7000 8900 12,700 13,000 15,000 18,400 19,500 16,500

cobalt. Cobalt was calculated from the sample responses relative to standards and blanks.

3. Results and discussion 3.1. Copolymer synthesis and characterization The PDMS-b-PMVS copolymers were prepared at room temperature in cyclohexane utilizing THF as a promoter (Fig. 2). It is well-known that D3 polymerizations can be living under the conditions utilized in this work [21,22]. These diblock copolymers had narrow molecular weight distributions suggesting that the polymerizations of the PMVS blocks were also living. Recent work by Kickelbick supports this premise by demonstrating a linear dependence of molecular weight with monomer conversion for DV 3 [23].

Mn by 29Si NMR (g molK1) 6000 6300 9000 12,500 12,500 16,000 18,500 20,000 17,000

Mn by GPC (g molK1) 7800 9000 8700 13,000 14,700 13,700 17,500 22,000 17,200

Molecular weight distribution 1.03 1.05 1.01 1.02 1.01 1.06 1.01 1.09 1.03

7454

M.L. Vadala et al. / Polymer 45 (2004) 74497461

Fig. 4. The living anionic polymerization of D3 and the formation of PDMS blocks were monitored by 1H NMR.

Ring-opening reactions of cyclosiloxane trimers are promoted by enthalpic and entropic effects. The enthalpic contribution to ring-opening D3 has been attributed to relief of ring strain [24,25], and it is reasonable to hypothesize that 1,3,5-trimethyl-1,3,5-trivinylcyclotrisiloxane DV 3 is also strained. The entropy change for D3 polymerization is also favorable (positive), resulting in an overall KDG. The diblock copolymers were prepared by sequential anionic ring-opening polymerizations of D3 and DV 3; respectively. The synthesis of the PDMS block rst was important for controlling the chemical structure of these polymers. For example, analogous polymerizations wherein DV 3 was polymerized rst produced polymers with multimodal molecular weight distributions. This was attributed to a side reaction of the alkyllithium initiator reacting with vinylsilanes, and such reactions have been documented previously [2628]. Cason et al. demonstrated that alkyllithiums (n-butyllithium and phenyllithium) add to vinyltriphenyl and vinyltrimethylsilanes. Stober et al. have reported polymers and copolymers via anionic

polymerization of the vinyl groups initiated with n-butyllithium. By contrast, when D3 then DV 3 were polymerized sequentially, narrow molecular weight distributions suggested well-dened polymers. GPC chromatograms for these diblock copolymers indicated molecular weight distributions close to one (Fig. 3 and Table 1). It should be noted that Weber [29] and Kickelbick [23] have reported well-dened homopolymerizations of DV 3 by initiating the reactions with a weakly basic lithium silanolate, thus avoiding any reaction of an alkyllithium with the vinyl moieties [23]. The copolymerizations were conducted under rigorously anhydrous conditions to prevent any reaction of the strongly basic initiator with water (which would form LiOH and n-butane). D3 ring-opening was monitored with 1H NMR by observing the decrease of the methyl peak of the cyclic monomer at 0.14 ppm and the concurrent increase of the resonance corresponding to the methyl protons of the linear species at 0.06 ppm (Fig. 4). The D 3 was reacted quantitatively before the second monomer was added to ensure that the second block would be pure PMVS. The molar ratios of linear PDMS to grams of DV 3 controlled the PMVS block molecular weights. This reaction was also monitored via 1H NMR (methyl resonances of the cyclic monomer at 0.27 ppm and linear species at 0.17 ppm). The diblock copolymers were terminated with an excess of trimethylchlorosilane at approximately 90% conversion of the DV 3 . This somewhat early termination was done to avoid any backbiting or intermolecular chainchain substitutions which might occur at extremely low monomer concentrations. The relative compositions of each block in the diblock copolymers after isolation were conrmed by 1H and 29Si NMR. In the 1H NMR spectra, the integrals from the PDMS methyl resonances at 0.06 ppm were compared to the methyl resonances of the polymethylvinylsiloxane at 0.08 ppm (Fig. 5). The number average molecular weights were determined by ratioing the butyl endgroup integrals to the methyl resonance integrals. 29 Si NMR provides a valuable tool for characterizing polysiloxanes because of its wide frequency range. The

Fig. 5. 1H NMR spectra of the PDMS-b-PMVS diblock copolymers afforded a means of conrming the compositions of these copolymers.

M.L. Vadala et al. / Polymer 45 (2004) 74497461

7455

Fig. 6. Quantitative 29Si NMR of the PDMS-b-PMVS diblock copolymers conrmed the expected 1:1 ratios of the terminal silicon atoms.

PDMS silicons were well-separated from the PMVS silicons in the 29Si NMR spectra. The diblock copolymers had one silicon next to a butyl endgroup (8.59 ppm) and one silicon in the trimethylsilyl endgroup (7.86 ppm). Block copolymer compositions and molecular weights were obtained by ratioing the repeat unit integrals corresponding to each block to the endgroups (Fig. 6). In all cases, the block Mns derived from 29Si NMR paralleled the targeted Mns and those determined via 1H NMR (Table 1). 3.2. Copolymer functionalization and characterization The polymethylvinylsiloxane blocks in the PDMS-bPMVS copolymers were quantitatively hydrosilated with

triethoxysilane or trimethoxysilane to yield PDMS-bPMTES or PDMS-b-PMTMS, respectively (Fig. 7A). Alternatively, the PDMS-b-PMVS precursor copolymers were only partially hydrosilated (Fig. 7B). In these cases, the resulting diblock copolymers were comprised of a PDMS block linked to a poly(methylvinylsiloxane-comethyltrialkoxysilethylsiloxane) wherein the sequences of methylvinylsiloxane and methyltrialkoxysilethylsiloxane units in the second blocks were random. These partially hydrosilated diblock copolymers have been designated PDMS-b-(PMVS-co-PMTES) (for the cases of hydrosilation with triethoxysilane) and PDMS- b -(PMVS-coPMTMS) (for the cases of hydrosilation with trimethoxysilane).

Fig. 7. (A) Quantitative functionalization of a PDMS-b-PMVS copolymer precursor with triethoxy- or trimethoxysilanes; (B) PDMS-b-PMVS copolymer precursors partially hydrosilated with triethoxy- or trimethoxysilanes.

7456

M.L. Vadala et al. / Polymer 45 (2004) 74497461

Table 2 Glass transition temperatures of a series of PDMS-b-PMVS (prepolymers) and PDMS-b-PMTES (functionalized polymers) Mn PDMS (g molK1) 4200 5000 5000 10,500 11,000 13,000 16,200 17,000 14,500
a

Mn PMVS (g molK1) 1800 2000 2900 2200 2000 2000 2200 2500 2000

Total Mna (g molK1) 6000 7000 8900 12,700 13,000 15,000 18,400 19,500 16,500

Tg (8C) (pre-polymer) K129 K129 K129 K128 K128 K127 K127 K127 K127

Tg (8C) (functionalized) K124 K125

K126

The total Mns were derived from 1H NMR.

Hydrosilation reactions have been extensively utilized to prepare monomers with SiC bonds or to crosslink polysiloxanes. Karstedts catalysts (Pt0 complexed with divinyltetramethyldisiloxane) are available commercially as solutions in either organic solvents such as xylene or polysiloxane oligomers. Such complexed catalysts are soluble in polysiloxane media and have high reactivity. The PDMS-b-PMVS polymer precursors were rigorously dried to avoid any premature reaction of the alkoxysilyl groups. The hydrosilation reactions were monitored via 1H NMR by observing the disappearance of the SiH peaks (dZ4.5 ppm) and an upeld shift and broadening of the triethoxysilyl methylene peaks. Normal (b) and reverse addition (a) of the SiH across the double bond are known to occur [30]. Both a (13%) and b (87%) addition products were obtained. Two signals [d1.4 CH3 and d1.3 CH]

corresponding to a addition across the double bonds appeared as the reactions proceeded. 3.3. Thermal properties of PDMS-b-PMVS and PDMS-b(PMVS-co-PMTES) Bulk thermal properties of each diblock copolymer were measured via differential scanning calorimetry to probe their phase structures. The PMVS block molecular weight was held as constant as possible at approximately 2000 g molK1 and the PDMS block molecular weights were varied from 5000 to 16,000 g molK1. For each copolymer, only one glass transition temperature (Tg) was observed at approximately K26 8C (Table 2), suggesting that the PDMS and PMVS blocks were miscible. This is in agreement with previous studies of similar PDMS based

Fig. 8. Thermolysis of dicobalt octacarbonyl: (a) initial reaction mixture with peaks at 2020, 2050, and 2070 cmK1 corresponding to terminal CO and 1860 cmK1 attributed to bridging CO; (b) a spectrum representing the intermediate reaction stage showing new peaks at 2065 and 2055 cmK1 assigned to Co4(CO)12; (c) a spectrum showing residual carbonyl peaks after approximately 4 h and the absence of peaks at 2065 and 2055 cmK1 (the reactions were normally terminated at this point); (d) a spectrum after 12 h at toluene reux showing the complete disappearance of all carbonyl peaks, indicating full thermolysis.

M.L. Vadala et al. / Polymer 45 (2004) 74497461

7457

Fig. 9. Transmission electron micrographs of cobalt uids comprised of (a) 16,000 g molK1 PDMS-b-3400 g molK1 (PMVS-co-PMTMS); (b) 5000 g molK1 PDMS-b-3400 g molK1 (PMVS-co-PMTMS); and (c) 16,000 g molK1 PDMS-b-3900 g molK1 (PMVS-co-PMTES).

polymers [29,30]. Semlyen et al. [29,30], Rife et al. [31], and Weber et al. [29] have also reported low glass transition temperatures of PMVS homopolymers. The atactic nature of the stereocenters in the PMVS blocks prevents any crystallization of the PMVS homopolymers. PDMS-b-PCPMS-b-PDMS triblock copolymers previously studied in our laboratories were microphase separated even at low block molecular weights (e.g. 5000 g molK1 PDMS-b-2000 g molK1 PCPMS-b-5000 g molK1 PDMS [14]), and this was attributed to the nitrile containing central block (PCPMS) being signicantly more polar than the PDMS tail blocks. By contrast, the thermal properties of the new dispersion stabilizers suggested that they may not be multi-phase materials. To compare homopolymer transitions, a 1000 g molK1 PMVS homopolymer (Tg K137 8C) was quantitatively functionalized to form a PMTES

homopolymer (Tg K95 8C). The higher Tg in PMTES compared to PMVS probably reects less rotational mobility in the former. In the diblock series, the PMVS block lengths were kept constant at 2000 g molK1 (pre-functionalization) or 5800 g molK1 (post-functionalization). Only one Tg was observed for each of these diblock copolymers (Table 2). 3.4. Synthesis and characterization of PDMS-b-(PMVS-coPMTMS)/cobalt complexes The investigations described herein on the formation of stabilized cobalt nanoparticles utilized thermolyses of dicobalt octacarbonyl in the presence of PDMS-b-(PMVSco-PMTES) or PDMS-b-(PMVS-co-PMTMS) in toluene. The PMTES or PMTMS blocks complexed with cobalt nanoparticles and the PDMS blocks protruded into the

7458

M.L. Vadala et al. / Polymer 45 (2004) 74497461

Fig. 10. Specic magnetization curves for cobaltpolymer complexes prepared with (a) 5000 g molK1 PDMS-b-3900 g molK1 (PMVS-co-PMTES); and (b) 16,000 g molK1 PDMS-b-3900 g molK1 (PMVS-co-PMTES).

toluene solvent. Dicobalt octacarbonyl and the copolymer were dissolved in toluene and the reaction mixtures were heated to toluene reux to displace the carbon monoxide and form cobalt nanoparticles. The reactions were monitored by FTIR by observing changes in structure and the decrease of carbonyl peaks (Fig. 8). The carbon monoxide region around 2000 cmK1 initially showed three absorbance bands at 2020, 2050, and 2070 cmK1 attributed to terminal CO, and a peak at 1860 cmK1 due to the bridging carbonyls (Fig. 8a) [14,18]. Two new peaks at 2065 and 2055 cmK1 appeared in the IR spectra as the reactions proceeded, and these peaks were attributed to a Co4(CO)12 intermediate based on previous assignments [18] (Fig. 8b). Two to four hours of reaction at reux were required for the Co4(CO)12 intermediates to disappear. Residual peaks remained in the carbonyl region of the IR spectra (2075, 2030, 2010 cmK1) when these reactions were terminated which remain unassigned (Fig. 8c), suggesting that some carbon monoxide remains under these conditions. If these reactions were continued at toluene reux for long periods (e.g. 12 h) the carbonyl peaks completely disappeared (Fig. 8d). Unfortunately however, this also caused a considerable amount of cobalt to plate-out on the glass apparatus.

Dispersions prepared with PDMS-b-(PMVS-co-PMTES) or PDMS-b-(PMVS-co-PMTMS) copolymers (wherein half the repeat units in the anchor blocks were functionalized) were dark homogeneous uids without any visible precipitates. By contrast, however, cobalt dispersions with PDMSb-PMTES or PDMS-b-PMTMS copolymers (wherein the anchor block was fully functionalized) exhibited small amounts of precipitates. TEM micrographs of the cobalt polymer complexes with PDMS-b-(PMVS-co-PMTES) or PDMS-b-(PMVS-co-PMTMS) having 30004000 g molK1 anchor blocks showed electron dense, z1015 nm diameter cobalt particles (Fig. 9), and there were no discernible differences in particle sizes. The copolymer sheaths surrounding each particle were visible but it was not possible to determine their thickness due to contrast issues associated with the technique. Further investigations of these in situ syntheses will be conducted to ascertain whether particle sizes can be correlated with anchor block lengths and micelle sizes in the reaction solutions. Compositions of selected cobaltcopolymer complexes were measured by elemental analysis after isolation, and the corresponding saturated magnetizations were derived from vibrating sample magnetometry. This allowed for

M.L. Vadala et al. / Polymer 45 (2004) 74497461

7459

Fig. 11. s vs. H measurements conducted at (a) 300 K and (b) 5 K (zero-eld cooled hysteresis loop (%), eld-cooled hysteresis loop) on cobaltpolymer complexes prepared with 5000 g molK1 PDMS-b-3400 g molK1 (PMVS-co-PMTMS). Inset in (b) is an enlarged region around the origin for 5 K hysteresis loops.

estimating the specic saturation magnetizations of the cobalt species as ranging from w40110 emu gK1 of cobalt. The higher end of this range was consistent with previously reported values for approximately 10 nm diameter cobalt particles generated via thermolyses of dicobalt octacarbonyl [6,14,18]. The lower values observed in this work are likely attributable to the incomplete extent of carbon monoxide evolution. All of these values, however,

are consistently below the expected bulk cobalt magnetizations of 162 emu gK1 and the reasons for this remain unclear. Representative magnetization curves of these materials (Fig. 10) suggest that these copolymercobalt complexes display some magnetic hysteresis. Pileni et al. has reported hysteresis for cobalt particles organized in a two-dimensional lattice on a graphite substrate with mean particle diameters of approximately 68 nm [32].

7460

M.L. Vadala et al. / Polymer 45 (2004) 74497461

Room temperature and 5 K magnetic susceptometry measurements were used to investigate any presence of: (1) a cobalt oxide layer on the surface of the cobalt nanoparticles and (2) unreacted cobalt carbonyl species in the cobaltpolymer complexes (Fig. 11). The cobalt specic magnetization (s) shows a continuous increase with eld at high elds and low temperature. However, at room temperature s almost saturates at high elds. These observations are indicative of the presence of paramagnetic species within the sample. This paramagnetic component is believed to be cobalt carbonyl species that have not been incorporated into the cobalt nano-crystals. This observation is in agreement with the infrared spectra, which suggest residual carbon monoxide ligands are present after formation of the cobalt nanoparticle dispersions. It is anticipated that these residual paramagnetic cobalt species will be reacted in subsequent annealing steps for these materials. In addition, the eld-cooled 5 K hysteresis loop does not show a signicant shift from the zero eld-cooled hysteresis loop, indicating that there is no signicant oxide layer on the surface of the metallic cobalt nanoparticles. If cobalt oxide were present in any signicant amount, an asymmetric eld-cooled hysteresis loop shift, caused by the coupling of an antiferromagnetic layer (CoO) with a ferromagnetic layer (Co), would be expected [33].

not understood, but may be at least partially attributable to residual paramagnetic species and also to a somewhat amorphous cobalt structure. It will be important to determine whether post solgel or pyrolysis reactions can protect the cobalt surfaces against oxidation. Moreover, effects of such post reactions on cobalt magnetic properties should also be carefully studied.

Acknowledgements The authors are grateful for funding from DARPAAFOSR (contracts F49620-02-1-0408 and F49620-03-10332) and from NSF (contract #DMR-0312046), and for the generous donation of n-butyllithium initiator by the Lithco Division of FMC. M.A. Zalich thanks the Australian American Fulbright Commission for a Fulbright Postgraduate Fellowship to conduct research in Australia.

References
[1] Wells S. Preparation and properties of ultrane magnetic particles, PhD dissertation in chemistry. University of North Wales, Bangor, Wales; 1989. [2] Berkovski B, Bashtovoy V. Magnetic uids handbook. New York: Begell Publishing Co; 1996. [3] Rutnakornpituk M, Baranauskas VV, Rife JS, Connolly J, St Pierre TG, Dailey JP. Eur Cells Mater 2002;3:1025. [4] Dailey JP, Phillips J, Li C, Rife JS. J Magn Magn Mater 1999;194: 1408. [5] Rife JS, Rutnakornpituk M, Vadala ML, Wilson K, Hoyt J. Polym Prepr 2000;41:13689. [6] Stevenson J, Rutnakornpituk M, Vadala ML, Esker AR, Rife JS, Charles S, Wells S, Dailey JP. J Magn Magn Mater 2001;225:4758. [7] Sato T, Iijima T, Sekin M, Inagaki N. J Magn Magn Mater 1987;65: 252. [8] Popplewell J, Sakhnini L. J Magn Magn Mater 1995;149:72. [9] Sun Y, Rollins H, Guduru R. Chem Mater 1999;11:79. [10] Lin X, Sorensen C. Langmuir 1998;14:71406. [11] Pileni M. Langmuir 1997;13:326676. [12] Sorensen C, Klabunde K, Hajipanayis G. J Mater Res 1999;14:1542. [13] Platanova OA, Bronstein LM, Solodovnikov SP, Yanovskaya IM, Obolonkova ES, Valetsky PM, Wemz E, Antonietti M. Colloid Polym Sci 1997;275:42631. [14] Rutnakornpituk M, Thompson MS, Harris LA, Farmer K, Esker AR, Rife JS, Connolly J, St Pierre TG. Polymer 2002;43:233748. [15] Hess P, Parker Jr. P. J Appl Polym Sci 1966;10:191527. [16] Yin J, Wang ZL. Nanostruct Mater 1999;11:84552. [17] Smith T, Wychick D. J Phys Chem 1980;84:1621. [18] Tannenbaum R. Inorganica Chimica Acta 1994;227:22340. [19] Rutnakornpituk M. Synthesis of silicone magnetic uids for use in eye surgery, PhD dissertation in chemistry. Virginia Polytechnic Institute and State University, Blacksburg, VA; 2002. [20] Kofron W, Backawski L. J Org Chem 1976;41:187980. [21] Chojnowski J, Scibiorek M, Gladkova N. Polym Bull 2000;44:377. [22] Boileau S. In: McGrath JE, editor. Ring-opening polymerization: kinetics, mechanisms and synthesis. ACS Symposium Series, 286. Washington, DC: American Chemical Society; 1984. p. 2335. [23] Kickelbick G, Husing N, Bauer J. J Polym Sci, Part A 2002;40: 153951.

4. Conclusions and recommendations for further investigations Well-dened PDMS-b-PMVS diblock copolymers were prepared with controlled molecular weights and narrow molecular weight distributions. The PMVS blocks were either partially or fully functionalized with triethoxysilethyl or trimethoxysilethyl pendent moieties using hydrosilations. Thermal characterization suggested that the blocks in the derivatized copolymers were also miscible. Solutions of the partially hydrosilated PDMS-b-(PMVSco-PMTMS) or PDMS-b-(PMVS-co-PMTES) copolymers were used to form sterically stabilized cobalt nanoparticle dispersions without any observed aggregation. Cobalt yields in the dispersed complexes were approximately 6070 wt% of the amount charged. Considerable additional work will be required to understand the parameters that control the maximum cobalt to copolymer ratio in these complexes. It is desirable to maximize the concentration of cobalt to obtain polymercobalt particles with enhanced magnetic responses. To date, formation of cobalt in solutions of the fully functionalized analogues of these polymers has resulted in some aggregation, but the reasons for this behavior are also still unclear. Cobalt nanoparticles prepared in this research were approximately 1015 nm in diameter as determined by TEM. Specic saturation magnetizations of the cobaltcopolymer complexes ranged from w40110 emu gK1 of cobalt. The deviation from the bulk cobalt value of 162 emu gK1 is

M.L. Vadala et al. / Polymer 45 (2004) 74497461 [24] Chojnowski J, Cypryk M. In: Jones R, Ando W, editors. Siliconcontaining polymers. Dordrecht: Kluwer; 2000. [25] Frye C, Salinger R, Fearon R. J Org Chem 1970;35:1308. [26] Stober M, Michael K, Speier J. J Org Chem 1967;32:27403. [27] Nametkine N, Topchiev A, Dourgarian S. J Polym Sci, Part C 1963. Polymer Symposia No. 4 1053. [28] Cason L, Brooks H. J Am Chem Soc 1952;74:45823. [29] Weber W, Cai G. Polymer 2001;43:17539.

7461

[30] Clarson S, Dodgson K, Semlyen J. Polymer 1985;26:930. [31] Hoyt-Lalli J. Synthesis of functionalized polysiloxanes and investigation of highly lled thermally conductive microcomposites, PhD Dissertation in Chemistry, Virginia polytechnic institute and state university, Blacksburg, VA; 2002. [32] Russier V, Petit C, Legrand J, Pileni M. Appl Surf Sci 2000;164: 1939. [33] Nogues J, Schuller IK. J Magn Magn Mater 1999;192:20332.

Вам также может понравиться