You are on page 1of 16

Talanta 59 (2003) 1173 /1188 www.elsevier.

com/locate/talanta

Analytical investigation of the chemical reactivity and stability of aminopropyl-grafted silica in aqueous medium
Mathieu Etienne, Alain Walcarius *
Laboratoire de Chimie Physique et Microbiologie pour lEnvironnement, Unite Mixte de Recherche UMR 7564, CNRS-Universite H. Poincare Nancy I, 405 Rue de Vandoeuvre, F-54600 Villers-les-Nancy, France Received 9 September 2002; received in revised form 12 December 2002; accepted 20 December 2002

Abstract Various samples of aminopropyl-functionalized silica (APS) have been prepared by grafting an organosilane precursor 3-aminopropyl-triethoxysilane (APTES) onto the surface of silica gel. The amine group content of the materials has been adjusted by varying the amount of APTES in the reaction medium (toluene). The grafted APS solids have been characterized with using several analytical techniques (N2 adsorption, X-ray photoelectron spectroscopy, infrared spectrometry) to determine their physico-chemical properties. Their reactivity in aqueous solutions was studied by acid-base titration, via protonation of the amine groups, and by way of complexation of these groups by HgII species. APS stability in aqueous medium was investigated at various pH and as a function of time, by the quantitative analysis of soluble Si- or amine-containing species that have been leached in solution upon degradation of APS. The ' chemical stability was found to increase when decreasing pH below the pKa value corresponding to the RNH3 /RNH2 couple, but very low pH values were necessary to get long-term stability because of the high local concentration of the amine groups in the APS materials. Adsorption of mercury(II) ions on APS was also performed to confirm the longterm stability of the grafted solid in acidic medium. Relationship between solution pH and APS stability was discussed. For sake of comparison, the stability of APS in ethanol and that of mercaptopropyl-grafted silica (MPS) in water have been briefly considered and discussed with respect to practical applications of silica-based organic /inorganic hybrids, e.g., in separation science or in the field of electrochemical sensors. # 2003 Elsevier Science B.V. All rights reserved.
Keywords: Aminopropyl-grafted silica; Acid-base reactivity; Chemical stability; Dissolution kinetics; Analytical investigation

1. Introduction The application of organic /inorganic hybrid materials in various fields of chemistry, and

* Corresponding author. Fax: '/33-3-83-27-54-44. E-mail address: walcariu@lcpe.cnrs-nancy.fr (A. Walcarius).

especially in analytical sciences, is a current area of research [1 /5]. Indeed, these solids have the advantage to combine in a single material the properties of both components: the rigid threedimensional inorganic skeleton imparts mechanical stability, while adequately chosen organic functions bring a specific chemical reactivity. In this variety of materials, the organically modified

0039-9140/03/$ - see front matter # 2003 Elsevier Science B.V. All rights reserved. doi:10.1016/S0039-9140(03)00024-9

1174

M. Etienne, A. Walcarius / Talanta 59 (2003) 1173 /1188

silicas have recently attracted considerable attention and gave rise to a wide range of applications, mainly in analytical separations, electrochemistry and sensors [5 /10]. Such success is partly due to the versatility of the sol /gel process to prepare a wide range of materials with predesigned composition and structure, in various shapes, and with many different properties [11,12]. For example, this process was exploited for enzyme immobilization in inorganic matrices without activity loss and permitting, therefore, biosensor development [13], or to produce ceramic-carbon composite electrodes [14], or associated to the screen-printed technology to get disposable sensors [15]. Incorporation of as-synthesized materials into carbon paste electrodes for preconcentration analysis of metal ions was also reported [16 /18]. Moreover, the use of template molecules associated to the sol /gel process has allowed the huge development of ordered mesoporous silicas during the 1990s [19], for which the organically modified forms are very promising, e.g., for solid-phase extraction of heavy metal species from diluted solutions [20 /22]. Amine-functionalized silicas have been widely studied in their solid phase [23], and silica-based materials containing either aminopropyl groups or more complex ligands bearing amine functions, which were either covalently attached to the inorganic network or simply impregnated on a silica surface, have been often proposed as solid extractants for heavy metal species [24 /32]. Examples are available for CuII [24 /30], CdII and HgII [31,32], and some other such as CoII, NiII, ZnII or PbII [24,27,30 /32]. The binding ability of amine-bearing silicas was also exploited in electroanalysis, e.g., for the voltammetric detection of trace CuII after accumulation at electrodes modified with such solids [16,17]. To be efficient, all these applications would require the chemical stability of the hybrid material, at least in the particular conditions and within the time scale of the experiments. This aspect in relation to trace metal extraction from aqueous medium was, however, sparingly considered in the past, and most often not at all. The chemical stability of silica in aqueous medium is pH-dependent and decreases significantly in alkaline solutions [33]. When the silica

surface is functionalized with amine groups, it is expected that the basic character of these functions would affect the overall chemical stability (and reactivity) of the resulting hybrid aminopropylfunctionalized silica (APS) material. Covalent coupling between a silica network and aminopropyl groups usually proceeds with using the APTES precursor that is either grafted on an as-synthesized silica in organic solvent [23] or co-condensated with another silica precursor (e.g., tetraalkoxysilane) in hydroalcoholic medium leading to the one-step formation of aminopropylsiloxane gels [34]. Interest in the covalent linkage between the inorganic structure and the organic groups arises from the non-hydrolyzable Si /C bond in the organosilane, which prevents from leaching of the immobilized reagent in the external solution contrary to impregnation [25]. However, this advantage is only valid if no other degradation pathway (i.e., alkaline attack) is liable to transfer gradually the organic modifier into the solution. This may occur with APS materials in aqueous medium via the hydrolysis of siloxane bonds owing to the basic properties of the amine functions [34]. The acid-base properties of APS have been previously characterized by Zhmud et al. [34 /36]. It was especially shown that hydrogen bonds between amine groups and residual silanols (hydroxyl groups present on the silica surface) can arise from sprawling aminopropyl tails on the surface [35]. In the presence of water, this interaction promotes proton transfer from silanols to amine groups, which leads to the formation of zwitterion-like moieties ( /SiO(, 'H3N /) on the silica surface [35,36]. The amine groups can be protonated in acidic medium, this process being, however, rather slow owing to restricted diffusion in the porous material [37]. Moreover, they are soluble to some extent in aqueous medium (owing to hydrolysis of siloxane bonds, which is favored at high pH values), pointing out the lack of stability of such materials in water [34]. This relative chemical instability was also mentioned in some other reports [38 /40]. Except these few investigations, and despite the large record of works dealing with the use of amine-functionalized silicas for removal of heavy metal ions [24 /32], no

M. Etienne, A. Walcarius / Talanta 59 (2003) 1173 /1188

1175

detailed studies on the main parameters affecting the chemical stability of these materials in aqueous medium are available. In this work, we have thus examined the acidbase reactivity and chemical stability of silica gel samples grafted with aminopropyl groups in aqueous solutions. Various APS samples have been prepared, containing different amounts of grafted ligands. They have been characterized in solution by acid-base titration and quantitative analysis of their degradation products. Effects of pH and contact time with water were thoroughly investigated and critically discussed by taking into account the high local concentration of amine groups in APS. A mercaptopropyl-grafted silica gel (MPS) was used for comparison purpose. Better understanding the basic chemistry of APS in solution in a wide range of experimental conditions would contribute to better defining those required for optimal applications of these materials as solid-phase extractants or as electrode modifiers.

groups and an appropriate organosilane (APTES or MPTMS) in dry toluene [23,32]. Typically, 5 g of silica sample are dispersed in 50 ml dry toluene and stirred for a few minutes at room temperature; selected amounts of APTES (ranging between 20 and 0.060 ml), or 5 ml MPTMS, is then slowly added to the suspension and refluxed for 2 h (APTES) or 24 h (MPTMS). After slow cooling, the resulting solids are filtered, washed with toluene, and dried under reduced pressure for 24 h. The aminopropyl-grafted samples are heated at 120 8C for 12 h. The grafted materials are called afterwards APS (aminopropyl-silica) and MPS (mercaptopropyl-silica). 2.3. Apparatus Modified and unmodified silica materials have been characterized by various techniques. The total pore volume and specific surface area of grafted silica gels were estimated on the basis of nitrogen adsorption/desorption isotherms at the temperature of liquid nitrogen (BET method). These measurements were performed using the Coulter SA 3100 apparatus. The amine content of the APS samples was determined by acid-base titration [32,37], which was monitored with the Metrohm 691 pHmeter (electrode No. 6.0222.100) and by elemental analysis (Central Service for Analysis, CNRS, Lyon). APS samples were also characterized by X-ray Photoelectron Spectroscopy (XPS) and Infrared spectrometry (IR). XPS measurements have been performed at a residual pressure lower than 10 (9 mbar, with using a VSW HA150 MCD electron energy analyzer operating with a Mg Ka non-monochromatic source. IR experiments were carried out in the diffuse reflectance mode, with the aid of a Perkin /Elmer 2000 apparatus, by reflection on a KBr powder containing 10% of the silica-based material. Quantitative analysis of silicon in aqueous solution was made by inductively coupled plasma-atomic emission spectroscopy (ICP-AES, plasma 2000, Perkin /Elmer), and total amounts of amine in solution were quantified by pH-metry. Zeta potentials were measured by Doppler velocimetry (Malvern Instruments) on APS suspensions

2. Experimental 2.1. Chemicals and solutions All solutions were prepared with high-purity water (18 MV cm) from a Millipore milliQ water purification system. Nitric acid (min. 65%) and sodium acetate were purchased from Riedel de Hae n, HCl was obtained from Prolabo. Silica gel was the chromatographic grade Kieselgel Geduran 60 from Merck (average particle size: 70 mm). The reactants 3-aminopropyl-triethoxysilane (APTES) 99% and 3-mercaptopropyl-trimethoxysilane (MPTMS) 95% were, respectively, purchased from Aldrich and Lancaster. Hg(NO3)2 (Fluka), BuNH2 ( !/99%, Aldrich), dry toluene (99%, Merck), and ethanol (95 /96%, Merck) were used as received. 2.2. Synthesis of grafted silicas Grafting the silica surface by covalently attaching aminopropyl or mercaptopropyl functional groups proceeds via a reaction between silanol

1176

M. Etienne, A. Walcarius / Talanta 59 (2003) 1173 /1188

adjusted at selected pH values by HNO3 or NaOH additions. A potential of 50 mV was applied between two palladium-plated electrodes in a quartz suprasil chamber of 75 mm effective length. Measurements were performed after some minutes equilibration. Solution-phase HgII was determined by anodic stripping differential pulse voltammetry on gold electrode, using a m-Autolab potentiostat associated to the GPES electrochemical analysis system (Eco Chemie). Measurements were performed in a conventional single-compartment cell assembled with a rotating gold electrode, a Ag/ AgCl reference electrode (Metrohm, No. 6.0733.100), and a Pt wire auxiliary electrode. 2.4. Procedures 2.4.1. pH-metric titration Direct titration of each APS material was carried out by automatic addition of 10(2 M HCl in a reactor containing 40 ml of deionized water and 100 mg of solid sample. The titration speed was adjusted in order to neutralize all the amine groups in the modified silica material. A speed of typically 0.03 ml min (1 was selected as a good compromise to ensure diffusion of the reactant to all the active centers in the porous solid while keeping a reasonable experiment time. Complete neutralization was checked by back titration of 100 mg APS in an excess HCl, by a standardized NaOH solution, after 24 h reaction and filtration of the solid phase. The differences observed between direct and back titration were less than 2%. The amine group content was also determined by elemental analysis for confirmation purpose. 2.4.2. Monitoring the stability/degradation of APS in solution 0.1 g of silica was added to 200 ml of aqueous solution. After selected reaction times, 1 ml of the solution was taken out with a syringe and filtered off with a 0.45 mm HV Millipore filter. The silicon concentration in solution was then directly determined by ICP-AES. For each figure presented in this paper, measurements of all data points have been performed in a single set of experiments and with the same ICP-AES parameters. The etalon

curve was prepared with tetraethoxysilane ( !/ 98%, Merck). Loss of amine groups in solution was also measured to evaluate the rate and extent of APS instability. Typically, 0.1 g APS was placed in 50 ml of aqueous or ethanolic (96%) solution. After a selected equilibration time, the solid was filtered off and the filtrate was titrated by the pHpotentiometric method with using a standardized HCl solution. 2.4.3. Mercury uptake by APS in HCl solutions 0.1 g of the APS material was placed in 50 ml of aqueous solution containing initially 10 (4 M Hg(NO3)2 and 0.10 M (or 0.02 M) HCl. After a known time in solution, the solid was filtered off and the solution was analyzed. The mercury(II) determination has been performed by anodic stripping differential pulse voltammetry on rotating gold disk electrode ( 0/4 mm, v 0/500 min(1, electrolysis at 0.3 V for 30 s) in 100 ml of electrolytic solution (72 mM NaCl, 12 mM disodium-EDTA, 2.8 M HClO4), according to a published procedure [41].

3. Results and discussions 3.1. Grafting and solid-phase characterization The procedure used for the chemical modification of silica gel by APTES is referred to the work of Vansant et al. [23] and that of Waddell et al. [38]. It is schematically represented in Fig. 1 and involves two successive steps. During the first step, APTES is allowed to react with the silica surface in toluene under nitrogen atmosphere and constant stirring while refluxing. These conditions must be kept for 2 h in order to permit the diffusion of the organosilane molecule in all the pores of the material; even those located deeper in the interior of the porous solid [37]. During this step, the surface hydroxyl groups of silica (silanol groups) are condensing with the ethoxy groups of APTES, liberating EtOH in the medium. The second part of synthesis is a curing step at elevated temperature, which is required to increase the degree of condensation of the grafted layer [23]. Appropriate

M. Etienne, A. Walcarius / Talanta 59 (2003) 1173 /1188

1177

Fig. 1. Schematic representation of the silica surface modication by grafting APTES.

cross-linking is achieved after 2 h curing at 120 8C [38]. Usually, the surface modification of silica by APTES is performed after complete dehydration (by thermal treatment). Indeed, adsorbed water on the silica surface has a great effect on the grafted layer because it can participate to the hydrolysis of ethoxy groups carried by APTES. Moreover, the presence of water often leads to increasing the extent of condensation between neighboring aminopropylsilane molecules and could also generate hierarchical polymerization of the grafting agent with itself. On the other hand, it can contribute to increase the stability of the grafted layer [23]. For this reason, we have chosen to perform the surface modification without thermal pre-treatment because great stability is useful for applications in which the material is in contact with aqueous solutions. If operating in excess APTES, the quantity of grafted ligands is directly related to the amount of silanol groups on the silica surface because they are primarily involved in the grafting process. The number of silanols can be varied by changing the degree of hydroxylation of the silica surface, e.g., decreasing upon thermal treatment to condense adjacent /SiOH groups into siloxane bonds /Si / O /Si /, which requires high temperatures ( !/ 200 8C) [23], or increasing by rehydration in acidic medium [20]. In the present case, a single state of the silica surface was used throughout, containing 4.2 mmol OH g(1 (as measured by thermogravimetry), and the quantity of grafted amine groups was adjusted by changing the amount of APTES introduced in the reaction vessel containing dry toluene as the solvent. Fig. 2A depicts the relation-

Fig. 2. (A) Variation of the amount of grafted groups on the silica surface, determined by acid-base titration as a function of the initial quantity of APTES in the medium. (B) Corresponding variation of the total pore volume of grafted materials expressed with respect to the amount of grafted groups.

1178

M. Etienne, A. Walcarius / Talanta 59 (2003) 1173 /1188

ship that was found between the surface loading (number of amine groups measured by acid-base titration) and the initial quantity of organosilane, both expressed in mmol g(1 with respect to the mass of the final product. A linear relationship was observed between these two parameters up to a ligand concentration of about 1.35 mmol g(1. Moreover, the quantity of grafted aminosilane was always less than the initial quantity of APTES in the synthesis medium, indicating the occurrence of an equilibrium between physically adsorbed and non-adsorbed APTES molecules on silica in toluene. Taking into account that one aminopropylsilane 2 [23] and the surface covers approximately 50 A (1 coverage of 1.35 mmol g (i.e., the higher value on the linear relationship between surface loading and APTES concentration, see Fig. 2A), it can be calculated that the surface area covered by the organic layer is 406 m2 g(1 ((50 )/10(20 m2) )/ (1.32 )/10(3 mol g(1) )/(6.02 )/1023 mol (1) 0/ 397 m2 g(1). This value is very close to the specific surface area of the initial material, 390 m2 g(1 (determined by N2 adsorption, according to the BET method). It seems, therefore, that the presence of adsorbed water on the silica surface does not lead to multilayer adsorption, but only to a small contribution of hierarchical polymerization. Beyond this limit of 1.35 mmol g(1, it is necessary to use very large excess of APTES to observe only a small increase in the ligand loading up to about 1.7 mmol g(1 at the maximum (difference between these values is owing to hierarchical polymerization arising from residual water). Fig. 2B depicts the variation of the total pore volume of the materials as a function of the quantities of grafted amine groups. It exhibits clearly a linear relationship between these two parameters from the lowest toward the highest amine contents. No difference is observed between loadings lower and higher than 1.35 mmol g(1. This result indicates that the grafting process affects regularly the pore volume of the organically modified silica gel and does not lead to a sudden blockage of access to the mesopores, even for aminopropyl group contents higher than 1.35 mmol g(1, for which APTES polymerization could occur in addition to monolayer formation upon grafting.

Fig. 3. (A) Titration curves obtained for 0.1 g APS with adding 1.00 )/10 ( 2 M HCl at various speeds: (a) 1 ml min ( 1; (b) 0.5 ml min ( 1; (c) 0.1 ml min( 1; (d) 0.03 ml min ( 1. (B) Relative position of the rst pH jump (rst equivalent point) with respect to the end of reaction (second equivalent point) for all the APS samples as in Fig. 2.

The grafted products were also characterized by XPS and IR spectroscopy. Similar to previous works, they confirm the presence of the grafted aminopropyl groups. Diffuse Reflectance IR spectroscopy carried out on silica gel before and after grafting with APTES indicates a significant decrease in the silanol bands (isolated silanols at 3740 cm (1 and bridged (hydrogen bonded) silanol in the range 3550 /3670 cm (1) with a concomitant increase of new bands characteristics of the

M. Etienne, A. Walcarius / Talanta 59 (2003) 1173 /1188

1179

immobilized aminopropyl groups. These bands were attributed to both the symmetric and asymmetric stretching of CH3 and CH2 groups (nas(CH3) 0/2975 cm (1 (small intensity), nas (CH2) 0/2928 cm (1, ns(CH3) 0/2886 cm (1 (small intensity), ns(CH2) 0/2870 cm (1), and of NH2 (nas 0/3358 cm (1, ns 0/3279 cm(1) [42 /44]. Presence of bands of CH3 (small intensity) indicates the presence of some remaining ethoxy groups that have not been hydrolyzed. The XPS analysis of the APS surface shows the presence of silicon (Si 2p at 102.7 eV, 27.4%), oxygen (O 1s at 531.9 eV, 55.8%), carbon (C 1s at 284.6 eV, 11.7%), and nitrogen (N 1s at 399.0 eV, 5.2%). These results are in good agreement with those previously reported in the literature [45]. 3.2. Characterization in aqueous solution The chemical modification of silica by grafting APTES dramatically changes its surface properties. The surface of unmodified silica is intrinsically acid owing to the presence of silanol groups that are readily deprotonated in alkaline medium (pKa 0/6.89/0.2, as determined by Schindler and Kamber [46] at 25 8C in 0.1 M NaClO4). This reaction, however, is not quantitative as the apparent pKa values of silanols are increasing significantly in proportion as deprotonation is going on [47,48]. When grafted with aminopropyl groups, the silica surface is expected to display basic properties. Fig. 3A shows illustrative titration curves for an APS sample by HCl at various speeds of reactant addition, from 1 ml min (1 down to 0.02 ml min (1, corresponding to experiment times ranging from about 15 min to 12 h. It is

clearly noticeable in this figure that a fast addition of hydrochloric acid leads to lowering pH in the medium in proportion to the quantity of added protons. At so short experiment times, all the added protons have no enough time to be consumed by the aminopropyl-modified silica (curves a /c in Fig. 3A). This limitation is owing to restricted diffusion inside the porous structure of APS, which requires rather long times for the protons to reach all the basic sites located deeper in the material [37]. For titration rates lower than 0.03 ml min (1, identical potentiometric curves were obtained, corresponding to steady-state situation: a speed of proton addition of 0.03 ml min(1 is sufficiently slow to allow the reactant to diffuse in the APS material and to reach all the active sites without creating an unsteady low pH in solution. Under conditions of complete titration, two different acid-base reactions are observed (two successive pH jumps, as shown in curve d in Fig. 3A). The first one is characterized by a pKa of about 9.6, and is consistent with the protonation of the free amine groups (into their corresponding ammonium form, Eq. (1)). The second one has a pKa of about 6.7 and is attributed to the protonation of the zwitterion-like species /SiO (, ' H3NC3H6 /Si / (Eq. (2)) that are known to exist in APS materials [34,35]. These two pKa values are in agreement with those observed for titration of pure APTES in aqueous medium, which forms octameric species, displaying a 91:9 percent ratio between the two successive steps [49]. In the case of APS, however, the free amine-to-zwitterion ratio was close to 60:40. It seems, therefore, that 40% of the aminopropyl groups in APS strongly interact with

(1)

1180

M. Etienne, A. Walcarius / Talanta 59 (2003) 1173 /1188

(2)

residual silanols to enable proton transfer (formation of /SiO(, 'H3NC3H6 /Si /) while 60% of them remain in the free amine form, which is probably hydrogen-bonded with a surrounding silanol group, as suggested by the shift IR bands of NH2 (nas 0/3358 cm (1, ns 0/3279 cm (1) towards values lower than those corresponding to free n -propylamine (nas 0/3365 cm (1, ns 0/3297 cm (1). All the APS materials containing various quantities of ligands have been titrated at low speed (equilibrium conditions). Fig. 3B shows that a linear relationship was observed between the first and the second equivalent points for all these materials, whatever the amine loading in the range 0.015 /1.71 mmol g(1. Therefore, the two distinct forms of aminopropyl groups grafted on silica gel are coexisting in aqueous suspension at a constant ratio of about 60:40 (owing to the slope of the straight line in Fig. 3B), independently on the grafting extent. However, it is difficult to draw an exact representation of the APS surface at this stage as long as the stability of this material in solution is not better understood (see Section 3.3). On the other hand, electrophoretic mobility measurements (Zeta potentials) carried out from aqueous suspensions of APS subjected to an electric field have brought additional information on the surface properties of this material. A major difference between unmodified and amine-grafted silica gels was indeed observed in the variation of the surface charge of particles with pH. The isoelectric point (IEP) of silica gel is close to pH 2; above this value the silica surface is negatively charged owing to the presence of silanolate groups [33]. On the opposite, the surface of APS samples

was found to be positive on a wider pH range, as explained by the fact that the great part of the amine population is protonated at pH lower than 10. When measuring Zeta potentials of APS as a function of pH, directly (i.e., a few min) upon dispersion of particles in solution to avoid significant degradation of the material, an IEP close to pH 10 was measured and the APS surface was positive at lower pH values (e.g., '/60 mV at pH 8). Such Zeta potentials and IEP are consistent with the pKa value of the grafted aminopropyl groups that are mainly protonated at pH B/10. This IEP value is, however, noticeably higher than those reported for other APS materials when awaiting for equilibration before starting the Zeta potential measurements (IEP 0/8 [50] and IEP /7 [35,36]). In these latter cases, very long times were required to reach steady-state values (i.e., 3 days [36]) so that significant degradation of the APS materials is expected to have occurred [38 /40]. In addition, local pH in the porous APS structure might be different as that in solution as a consequence of the high concentration of pHsensitive ligands in a confined environment whose accessibility to the external solution is time-dependent [37]. This indicates that the surface properties of APS are exposed to variation over prolonged contact with an aqueous phase, as discussed hereafter. 3.3. Stability in solution 3.3.1. Inuence of pH on the dissolution of APS materials in aqueous medium A first characterization of the APS stability in aqueous medium was provided by monitoring the

M. Etienne, A. Walcarius / Talanta 59 (2003) 1173 /1188

1181

Fig. 4. (A) Extent of dissolution (soluble Si) of 0.1 g APS in three different media (200 ml): (a) 0.1 M HNO3; (b) acetate buffer at pH 5.7; (c) pure water (pH is imposed by the intrinsic basicity of the material, which was 9.5 at the end of the experiment). Inset: same experiment prolonged over 4 days. (B) Long-term (in)stability of 0.1 g silica-based materials in 200 ml acetate buffer at pH 5.7: (a) APS containing 1.7 mmol amine groups per gram; (b) unmodied silica gel in the presence of butylamine in solution in the same quantity as the amine groups in (a). Inset: difference between curves (b) and (a). Data are expressed in the form of variation of soluble silicon concentrations in solution.

total soluble silicon-containing species that have leached in solution as a function of time. Their concentration is expected to increase as a result of partial hydrolysis of the silica network as well as from the liberation in solution of aminopropylsilane moieties arising from the hydrolysis of the chemical bond between the organosilane and the

silica surface. This latter reaction can be catalyzed by amine groups [39]. This preliminary stability study has been performed at various pH with an APS material grafted with 1.7 mmol g(1 aminopropyl groups; similar results were obtained for solids characterized by lower capacity. Fig. 4A shows the evolution over 2 h (inset: over 4 days) of the silicon concentration in solution when 0.1 g APS was placed in suspension into 200 ml of three different solutions: 0.1 M HNO3 (pH 1), acetate buffer (pH 5.7) and pure water (in this last medium the solution pH is compelled by the intrinsic basicity of the amine-bearing APS material at a value of about 9.5). It is clearly shown in this figure that initial pH of the suspension has a dramatic effect on the quantity of silicon species liberated in solution, even during the first minutes. The extent of solubilization is very low at pH 1 (Fig. 4A, curve a), much higher at pH 5.7 (Fig. 4A, curve b), and maximal when the solution pH is not controlled (Fig. 4A, curve c), enabling the material to express its total basic power. Rationalization of these results is quite easy at pH 1 (where the basic action of amine groups is prevented because they are totally protonated) and at pH 9.5 (where silica is expected to dissolve to significant extent), but is rather surprising at pH 5.7 as unmodified silica is usually stable at this pH value and as most of the amine groups of APS are expected to be protonated ( !/99.99%). Despite these latter facts, the APS material displays a significant rate of dissolution during the first hour of suspension in a buffered solution at pH 5.7 (Fig. 4A, curve b). Note that in this case (pH 5.7) the dissolution extent, as expressed by mass ratio with respect to the mass of starting material, remains low: about 1% degradation after 1 h, and 6% after 10 h in suspension. It seems, therefore, that the grafted aminopropyl groups at the silica surface are playing a key role in the instability of the APS material in aqueous medium, even in non-basic environment. This is further confirmed by comparing the behavior of silica-based materials in the absence and in the presence of amine groups, either dispersed in solution (soluble base) or immobilized at the silica surface (grafted bases). Fig. 5A summarizes the results of three experiments car-

1182

M. Etienne, A. Walcarius / Talanta 59 (2003) 1173 /1188

Fig. 5. Dissolution kinetics of 0.1 g silica-based materials in various conditions in (A) 200 ml acetate buffer at pH 5.7 and (B) 200 ml pure water: (a) unmodied silica gel alone; (b) APS containing 1.7 mmol amine groups per gram; (c) unmodied silica gel in the presence of butylamine in solution in the same quantity as the amine groups in (b). Other conditions as in Fig. 4.

ried out on the same time scale as in Fig. 4A, in acetate buffer at pH 5.7. Curve a shows the hydrolysis of 0.1 g of unmodified silica gel in this medium, which is very low and slow. Curve b illustrates the hydrolysis of 0.1 g of APS material in the same conditions and, as seen previously in Fig. 4A, the quantity of soluble silicon increases immediately after the material was suspended in solution. The last curve c shows the behavior of a suspension containing 0.1 g of the unmodified silica in the presence of soluble BuNH2, at an

amount corresponding to about the molar quantity of the amine grafted on the surface of 0.1 g of APS. Although the quantity of amine in solution is the same as that immobilized within the APS material, the hydrolysis of silica in the presence of free amine (non-grafted) is comparable to that of the same material without base (comparison between curves a and c in Fig. 5A). The acetate buffer at pH 5.7 is thus able to neutralize the basicity of soluble BuNH2 distributed into the whole volume of solution (0.85 )/10(3 M), but it is not sufficient to counterbalance the basicity arising from the APS material. In this last case, the amine groups are not dispersed into the whole solution but they are confined within the APS material at a high concentration (2.2 M, as estimated from the amine loading of 1.7 mmol g(1 and total pore volume of 0.76 ml g(1). When performing a similar set of experiments as in Fig. 5A, but in pure water (i.e., floating pH) instead of buffer solution, the dissolution of unmodified silica was still very slow (Fig. 5B, curve a), that of APS was very fast (Fig. 5B, curve b), but that of unmodified silica in the presence of BuNH2 led to a sharp increase in the silicon concentration in solution, contrarily to what happened in the buffer. This demonstrates the ability of free amine in solution to dissolve silica when it is not neutralized by an acid buffer, in agreement with the increase in the hydrolysis of silica when rising pH [33]. By comparing Figs. 4A and 5A and B, it appears that APS dissolution in aqueous medium is mainly due to the presence of amine groups, but their quantity is not the only parameter explaining the high rate of dissolution of the material in acetate buffer. Indeed, the high local concentration of amine groups in APS materials can not be buffered efficiently to avoid hydrolysis. It was only possible to counter efficiently the high basicity of the concentrated aminopropyl groups in APS by dispersing the material into a 0.1 M nitric acid solution (Fig. 4A, curve a); for such an external medium, the concentration of residual unprotonated amine groups in the material was as low as 5 )/10(9 M. The behavior of APS in buffered solution (pH 5.7) was also investigated at longer equilibration times. Fig. 4B compares the evolution of the

M. Etienne, A. Walcarius / Talanta 59 (2003) 1173 /1188

1183

silicon concentration measured as a function of time in suspensions containing, respectively, an unmodified silica gel in the presence of BuNH2

free aminopropylsilane. This is expected to occur to the substrate which is either singly, doubly, or triply bonded to APTES; the last step being

(3)

(curve a) and the APS material (curve b), during 2 days. In agreement with what was observed previously (Fig. 5A), a sharp increase of the soluble silicon appeared during the first 250 min ( /4 h) when the APS material was suspended in the buffer solution, while a continuous slow dissolution of the unmodified silica was observed over the entire time range (Fig. 4B, curve b). After the first 4 h, the degradation rate of APS was slower, displaying a speed of dissolution comparable to that of unmodified silica gel. The dissolution of the unmodified silica gel in this medium is owing to the high ionic strength generated by the buffer [33,51], while this parameter is not ratedetermining in the dissolution of APS. The inset in Fig. 4B depicts the difference between curves a and b. It allows clearly to distinguish a breakthrough during APS dissolution in aqueous medium: a fast degradation at short time due to the presence of amine groups in the material, followed by a slow dissolution at longer times similar to that of unmodified silica. Steady state situation appears after several hours. Because of possible destruction of the Si /O /Si bond by nucleophilic attack of amine groups (catalyzed by water molecules), one can suggest that liberation of silicon in solution arises from the deterioration of chemical bonds between the silane layer and the silica surface leading to leaching of

illustrated by the following equation: Indeed, the amount of aminopropylsilane that has leached out of APS after 4 h equilibration in aqueous medium (i.e., just before the breakthrough in curve b of Fig. 4B) is about 0.9 mmol g(1, which is less than the initial loading of the APS material used for this experiment (1.7 mmol g(1). Aminosilane liberation, however, cannot be the sole mechanism involved in the degradation process as the quantity of silicon in solution after 48 h equilibration is higher than 2 mmol g(1, which exceeds the amount of APTES that has been grafted on the material. Some other silicon-containing species originating from the bulk material have also passed in solution. To distinguish between these two processes, the quantity of amine liberated in pure water (drastic conditions concerning the stability) has been determined by potentiometric titration (Fig. 6, curve b). By this way it is possible to characterize quantitatively the extent of leaching of aminopropylsilane in solution, as a function of time, and to compare it with the amount of total soluble silicon. As shown, a fast liberation of aminosilane was observed during the first 2 h with the APS material in solution. A steady state was reached after typically 4 h and did not change later on, even after several days (data not shown). The quantity of amine liberated at the equilibrium does not correspond to all the amino-

1184

M. Etienne, A. Walcarius / Talanta 59 (2003) 1173 /1188

Fig. 6. Inuence of solvent on the degradation rate of APS (expressed through the variation of concentration of aminopropyl groups that have leached in the external solution with time): (a) ethanol at 96%; (b) pure water. Data were obtained from 0.1 g solid in 50 ml solution.

Fig. 7. Extent of mercury(II) adsorbed by 0.1 g APS, as a function of time, in solutions containing initially 10 ( 4 M Hg(NO3)2 and two different HCl concentrations: (a) 0.02 M; (b) 0.10 M.

silanes available within the material. At the maximum, only 67% of the total aminopropylsilane content of APS were liberated in solution: onethird of them was still remaining at the surface of the silica material after several days in closed reactor. A possible interpretation of APS degradation involves a fast initial step resulting from the leaching of aminopropylsilane species in solution, and a subsequent slow degradation event which is essentially owing to the attack of the bulk silica that is catalyzed by the basic conditions generated by the solution-phase aminopropylsilane (hydroxide anions produced by hydrolysis of amine groups). Interestingly, the maximal amount of aminopropylsilane species that pass in solution when suspending APS particles in pure water (i.e., 67% of them) is close to the amount of those groups that are in the form of free amine in the material (those having reacted in the first part of titration curve; see Section 3.2 and Fig. 3A). It seems, therefore, that zwitterion-like species ( /SiO (, ' H3NC3H6 /Si /) are much more stable and less subject to leaching in solution, in agreement with the fact that ammonium functions do not possess the pair of electrons of amine functions that is responsible for the hydrolysis of the grafted aminopropylsilanes (Eq. (3)).

3.3.2. Inuence of proton concentration on the longterm stability of APS in acidic solutions and restricted uses in alkaline medium */interactions with metal ion species As shown above, APS particles exhibit noticeably long stability when immersed in acidic solutions. This stability was further studied with respect to the binding of the negatively charged chloro-complexes of mercury(II) to protonated APS in acidic medium, which can occur via electrostatic interaction with the ammonium ' groups /NH3 ,Cl ( that are formed in the ( presence of HCl. In this medium, both HgCl3 2( and HgCl4 species are liable to exist in a proportion depending on the chloride ion concentration. These anionic complexes are liable to exchange chloride ions in protonated APS (see ( Eq. (4), as an illustrative case for HgCl3 ). The percentage of accumulated mercury(II) within the protonated APS material has been followed over several weeks in two different acidic media, 0.02 M HCl (Fig. 7, curve a) and 0.1 M HCl (Fig. 7, curve b). Immobilization of the mercury(II) complexes by ion exchange in the material was rather fast and an equilibrium was observed after some hours in both solutions, corresponding to the consumption of about 65% of the initial solution-phase metal ion concentration. These HgII-loaded APS particles were then

M. Etienne, A. Walcarius / Talanta 59 (2003) 1173 /1188

1185

allowed to equilibrate further in the same solutions while performing a continuous monitoring of the

surface. Depending on the time of contact required for a target application of this kind of material in

(4)

solution-phase metal ion concentrations over several days, in order to compare the stability of HgIIloaded APS for these two different acidic strengths. As shown in Fig. 7, the materials capacity did not change significantly during more than 5 weeks equilibration in 0.1 M HCl, while a slow decrease of the accumulated mercury(II) was recorded when working in less diluted HCl (0.02 M). This latter behavior is explained by a slow loss of ligands in solution due to some remaining free amine groups that are not fully eliminated in 0.02 M HCl because of the high local concentration of ligands in the APS material; a higher HCl concentration (e.g., 0.1 M) is required to entirely compensate this remaining local basicity. Long-time immobilization of metal ions by APS requires, therefore, a rather strong acidic medium. At this point of the discussion, it is of interest to compare Figs. 4B, 6 and 7. In pure water, the release of aminosilane occurs very quickly and a steady state is observed after several hours in the solution (Fig. 6, curve b). In acetate buffer at pH 5.7, a steady state seems to appear after 2 days in solution (inset of Fig. 4B). Finally, in 0.02 M HCl solution, a very slow decrease of the ligand loading in APS is observed within several weeks (Fig. 7, curve a). These three experiments show the relationship existing between pH of the aqueous solution and the stability of the chemical bonds relying on the aminopropylsilane and the silica gel

aqueous medium, the results presented in this study enable to select carefully the experimental conditions that must be used in order to keep a sufficient reactivity and to limit the APS degradation during the time scale of the experiment. For example, we have proposed recently the use of an APS-modified carbon paste electrode for the electrochemical detection of copper(II) ions in aqueous medium at pH 7 [17]. In spite of the

Fig. 8. Effect of grafting the silica surface with mercaptopropyl groups on its chemical stability in aqueous medium as a function of pH. Extent of dissolution (soluble Si) of 0.1 g MPS in 0.1 M acetate buffer at pH 5.0 (a), and in 0.1 M phosphate buffer at pH 8.3 (b); curve (c) depicts the case of 0.1 g of unmodied silica gel in phosphate buffer at pH 7.9, for comparison purpose.

1186

M. Etienne, A. Walcarius / Talanta 59 (2003) 1173 /1188

long-term instability of this material at such a pH value, the analysis was possible because of the short contact times of APS particles with the external solution. Following the same logic, the use of APS material would be possible during several days in 0.02 M HCl, and several weeks in 0.1 M HCl. 3.3.3. Improved stability of silica-based organic / inorganic hybrids in solutions The main process inducing degradation of APS is the breaking of the chemical bonds between the aminopropylsilane layer and the silica surface, according to a catalytic hydrolysis of siloxane bonds involving the amine function and a water molecule (Eq. (3)). It has been aforementioned that protonation of the amine groups enables a significant increase of the stability of the material in aqueous solution. Another possibility is to limit the contact of water with the surface of the modified silica material. Indeed, dispersion of APS in ethanol instead of water reveals a significant increase of stability of the material in this organic solvent (comparison of curves a and b in Fig. 6). While aminosilane is immediately released in aqueous solution (curve b), this degradation process remains very low (curve a) on the same time scale in ethanol solution, demonstrating a better stability of this material in a medium containing only a low percentage of water ( /5%). Chemical deterioration of silica-based materials can also be circumvented via functionalization of their surface with hydrophobic groups in order to limit the contact with water, which is known to promote the nucleophilic attack of siloxane bonds [52]. This procedure has been recently exploited to reduce the destructive effect of moisture in ordered mesoporous silica materials (MCM-48) by grafting their internal surfaces with either trimethylchlorosilane or n -octyldimethylchlorosilane moieties [53,54]. In the goal to evaluate the interest of this approach to limit the dissolution of silica-based materials, a silica gel surface was grafted with mercaptopropyltrimethoxysilane (MPTMS). An MPS was obtained, containing 1.5 mmol g (1 thiol groups, which is close to the amine groups content of the APS material used for stability experiments (1.7 mmol g(1). The variation of specific surface

area upon modification with MPS ((/19%) was very close to the decrease observed for APS ((/ 22%). Then, the difference between APS and MPS is mainly due to the nature of functional groups supported by the propyl chain of the organosilane. The dissolution of MPS has been followed and compared with that of pure silica gel. Fig. 8 depicts the evolution of the solution-phase silicon concentration when the material was immersed in acetate buffer. At pH 5, no significant silicon leaching in solution was observed (Fig. 8, curve a), indicating a high stability of this material in aqueous solutions at such pH. The same material was then placed in phosphate buffer at pH 8.3 (curve b), for which a slow increase of the silicon concentration was observed. This dissolution rate was, however, much lower than that of the pure silica gel dispersed in phosphate buffer at pH 7.9 (curve c). Although pH in this last case was slightly lower than that experienced by the mercaptopropyl-modified material, the rate of degradation was higher for the unmodified material than for the thiol-modified material. So, the hydrophobic property of mercaptopropyl groups generates significant protection against degradation of the silica network even in the presence of diluted hydroxyl anions in the external solution. The association of such modification with amine-bearing silica materials could lead to extend the range of applications of these organic /inorganic hybrid materials in aqueous solutions.

4. Conclusions Aminopropyl groups grafted on the surface of silica gel do exist under two different forms, free unprotonated amine ( /Si /C3H6 /NH2) and zwitterion-like species ( /SiO(, 'H3NC3H6 /Si /), which can be distinguished by pH-metric titration. These latter are much more stable in aqueous solution (several hours) while the first ones contribute to a fast hydrolysis reaction of Si /O /Si bonds, leading to the leaching of aminopropylsilane in the external solution. The high local concentration of these ligands in the APS material enables this degradation to occur even in buffered solutions at pH 5.7, and to a lesser extent down to

M. Etienne, A. Walcarius / Talanta 59 (2003) 1173 /1188

1187

pH 2. Long-term stability of the grafted solid in acidic medium was only observed at pH as low as 1. Surface properties of APS are strongly dependent on pH and equilibration time in aqueous solution. Exploitation of the ligand properties of such a material for solid /liquid extraction processes should take into account the above considerations; in particular, the experimental conditions should be adjusted for the target application by taking into account the required chemical form of aminopropyl groups and the time scale of the experiment. APS materials can be used in neutral or slightly alkaline for applications requiring the exploitation of the complexation properties of amine groups only if short-time experiments (i.e., few min) are involved. They can be utilized in strong acidic media over prolonged periods of time (several days) in their protonated form. Chemical stability of APS is noticeably improved in organic solvents and might be even better in aqueous medium after grafting hydrophobic groups on its surface.

References
[1] [2] [3] [4] P. Gomez-Romero, Adv. Mater. 13 (2001) 163. A.D. Pomogailo, Russ. Chem. Rev. 69 (2000) 53. Special issue of Chem. Mater. 13 (2001). M.M. Collinson, in: H.S. Nalwa (Ed.), Handbook of Advanced Electronic and Photonic Materials and Devices, vol. 5, Academic Press, 2001, pp. 163 /194. M.M. Collinson, Mikrochim. Acta 129 (1998) 149. M. Tsionsky, G. Gun, V. Glezer, O. Lev, Anal. Chem. 66 (1994) 1747. O. Lev, Z. Wu, S. Bharathi, V. Glezer, A. Modestov, J. Gun, L. Rabinovich, S. Sampath, Chem. Mater. 9 (1997) 2354. (a) A. Walcarius, Electroanalysis 10 (1998) 1217; (b) A. Walcarius, Electroanalysis 13 (2001) 701. A. Walcarius, Chem. Mater. 13 (2001) 3351. M.M. Collinson, Trends Anal. Chem. 21 (2002) 30. L.L. Hench, J.K. West, Chem. Rev. 90 (1990) 33. C.J. Brinker, G.W. Scherer, Sol /Gel Science, Academic Press, San Diego, CA, 1990. J. Wang, Anal. Chim. Acta 399 (1999) 21. L. Rabinovich, O. Lev, Electroanalysis 13 (2001) 265. J. Wang, P.V.A. Pamidi, D.S. Park, Anal. Chem. 68 (1996) 2705. A. Walcarius, N. Lu thi, J.-L. Blin, B.-L. Su, L. Lamberts, Electrochim. Acta 44 (1999) 4601.

[5] [6] [7]

[8] [9] [10] [11] [12] [13] [14] [15] [16]

[17] M. Etienne, J. Bessie ` re, A. Walcarius, Sens. Actuators B 76 (2001) 531. [18] S. Sayen, M. Etienne, J. Bessie ` re, A. Walcarius, Electroanalysis 14 (2002) 1521. [19] J.S. Beck, J.C. Vartuli, Curr. Opin. Solid State Mater. Sci. 1 (1996) 76. [20] X. Feng, G.E. Fryxell, L.-Q. Wang, A.Y. Kim, J. Liu, K.M. Kemmer, Science 276 (1997) 923. [21] L. Mercier, T.J. Pinnavaia, Adv. Mater. 9 (1997) 500. [22] A. Stein, B.J. Melde, R.C. Schroden, Adv. Mater. 12 (2000) 1403. [23] E.F. Vansant, P. Van der Voort, K.C. Vrancken, Characterisation and Chemical Modication of the Silica Surface, Elsevier, The Netherlands, 1995. [24] D.E. Leyden, G.H. Luttrell, Anal. Chem. 47 (1975) 1612. [25] I.S. Khatib, R.V. Parish, J. Organomet. Chem. 369 (1989) 9. [26] G.V. Kudryavtsev, D.V. Miltchenko, V.V. Yagov, A.A. Lopatkin, J. Colloid. Interf. Sci. 140 (1990) 114. [27] E.M. Soliman, Anal. Lett. 30 (1997) 1739. [28] A. Klonkowski, B. Grobelna, T. Widernik, A. JankowskaFrydel, W. Mozgawa, Langmuir 15 (1999) 5814. [29] C.A. Borgo, R.T. Ferrari, L.M.S. Colpini, C.M.M. Costa, M.L. Baesso, A.C. Bento, Anal. Chim. Acta 385 (1999) 103. [30] E.M. Soliman, M.E. Mahmoud, S.A. Ahmed, Talanta 54 (2001) 243. . Ko [31] U klu , Chim. Acta Turc. 12 (1984) 267. [32] J.J. Yang, I.M. El-Nahhal, I.-S. Chuang, G.E. Maciel, J. Non-Cryst. Solids 209 (1997) 19. [33] R.K. Iler, The Chemistry of Silica, Wiley, New York, 1979. [34] B.V. Zhmud, J. Sonnefeld, J. Non-Cryst. Solids 195 (1996) 16. [35] A.A. Golub, A.I. Zubenko, B.V. Zhmud, J. Colloid. Interf. Sci. 179 (1996) 482. [36] B.V. Zhmud, A.B. Pechenyi, J. Colloid. Interf. Sci. 173 (1995) 71. [37] A. Walcarius, M. Etienne, J. Bessie ` re, Chem. Mater. 14 (2002) 2757. [38] T.G. Waddell, D.E. Leyden, M.T. DeBello, J. Am. Chem. Soc. 103 (1981) 5303. [39] M. Gimpel, K.K. Unger, Chromatographia 16 (1982) 117. [40] J.R. Jezorek, K.H. Faltynski, L.G. Blackburn, P.J. Henderson, H.D. Medina, Talanta 32 (1985) 763. [41] Y. Bonl, M. Brand, E. Kirowa-Eisner, Anal. Chim. Acta 424 (2000) 65. [42] C.-H. Chiang, H. Ishida, J.L. Koenig, J. Colloid. Interf. Sci. 74 (1980) 396. [43] R.S.S. Murthy, D.E. Leyden, Anal. Chem. 58 (1986) 1228. [44] K.C. Vrancken, P. van der Voort, I. Gillis-DHamers, E. Vansant, P. Grobet, J. Chem. Soc. Faraday Trans. 88 (1992) 3197. [45] K.M.R. Kallury, P.M. Macdonald, M. Thompson, Langmuir 10 (1994) 492. [46] P. Schindler, H.R. Kamber, Helv. Chim. Acta 51 (1968) 1781.

1188

M. Etienne, A. Walcarius / Talanta 59 (2003) 1173 /1188 [51] S.S. Jorgensen, Acta Chem. Scand. 22 (1968) 335. [52] A. Cauvel, G. Renard, D. Brunel, J. Org. Chem. 62 (1997) 749. [53] K.A. Koyano, T. Tatsumi, Y. Tanaka, S. Nataka, J. Phys. Chem. B 101 (1997) 9436. [54] A. Matsumoto, K. Tsutsumi, K. Schumacher, K.K. Unger, Langmuir 18 (2002) 4014.

[47] J. Sonnefeld, J. Colloid. Interf. Sci. 155 (1993) 191. [48] C. Despas, A. Walcarius, J. Bessie ` re, Langmuir 15 (1999) 3186. [49] L. Coche-Gue rente, V. Desprez, P. Labbe , J. Electroanal. Chem. 458 (1998) 73. [50] S. Kondo, T. Ishikawa, N. Yamagami, K. Yoshioka, Y. Nakahara, Bull. Chem. Soc. Jpn. 60 (1987) 95.

You might also like