You are on page 1of 13

Journal of Colloid and Interface Science 289 (2005) 249–261

www.elsevier.com/locate/jcis

Silane adsorption onto cellulose fibers:


Hydrolysis and condensation reactions
Marie-Christine Brochier Salon a , Makki Abdelmouleh b , Sami Boufi b ,
Mohamed Naceur Belgacem a,∗ , Alessandro Gandini c
a LGP2, Ecole Française de Papeterie et des Industries Graphiques (INPG), BP 65, Domaine Universitaire, F-38402 St. Martin d’Hères, France
b LMSE, Faculté des sciences de Sfax, BP 802-3018 Sfax, Tunisia
c Instituto de Química de São Carlos, Universidade de São Paulo, Av. Trabalhador São Carlense, 400, CEP 13566-590 São Carlos, SP, Brazil

Received 23 February 2005; accepted 25 March 2005


Available online 23 May 2005

Abstract
The hydrolysis of three alkoxysilane coupling agents, γ -methacryloxypropyltrimethoxysilane (MPS), γ -aminopropyltriethoxysilane
(APS), and γ -diethylenetriaminopropyltrimethoxysilane (TAS), was carried out in an ethanol/water (80/20) solution and followed by 1 H,
13 C, and 29 Si NMR spectroscopy, which showed that its rate increased in the order MPS < APS < TAS. The formation of the silanol groups
was followed by their self-condensation to generate oligomeric structure. APS and MPS only gave soluble products, whereas colloidal parti-
cles precipitated in the medium when TAS was hydrolyzed. Pristine and hydrolyzed MPS were then adsorbed onto a cellulose substrate and
thereafter a thermal treatment at 110–120 ◦ C under reduced pressure was applied to the modified fibers to create permanent bonding of the
coupling agent at their surface.
 2005 Elsevier Inc. All rights reserved.

1. Introduction Usually, the surface treatment is carried out with a silane


water–alcohol solution in a concentration range of 0.5–2%
Functional trialkoxysilanes, R Si(OR)3 , are widely used by weight. These conditions offer several advantages, in par-
in numerous industrial applications as coupling agents, to ticular (i) an increase of silane solubilization in the medium,
enhance adhesion between polymeric matrices and inorganic (ii) better control of the film thickness on the surface, and
solids [1–5]. The mechanisms of these coupling reactions are (iii) more uniform coverage of the surface. The water in-
related to the presence of two types of reactive moieties in duces stepwise hydrolysis of the silane [5,11] to give the cor-
their structure, which respond in different ways, according responding silanol derivative R -Si(OH)3 , which promotes
to the substrates they approach. On one hand, the alkoxy the silane adsorption onto OH-rich substrates through hy-
groups OR enable the silane to be anchored to surfaces drogen bonding. The actual chemical condensation is known
bearing hydroxyl groups [6–8], and on the other hand, the to occur after thermal activation, leading to siloxane bridges
organic functionality R (amine, methacrylic, vinylic, etc.) [4,12,13], as follows:
improves their compatibility, or even copolymerizes with
organic matrices, thus enhancing the interfacial adhesion be-
tween the two phases [9,10].

* Corresponding author. Fax: +33 (0) 4 76 82 69 33.


.
E-mail address: naceur.belgacem@efpg.inpg.fr (M.N. Belgacem).

0021-9797/$ – see front matter  2005 Elsevier Inc. All rights reserved.
doi:10.1016/j.jcis.2005.03.070
250 M.-C. Brochier Salon et al. / Journal of Colloid and Interface Science 289 (2005) 249–261

Subsequently, after solvent evaporation, the residual onto the cellulose substrate. Indeed, the surface of the cellu-
silanol groups may undergo a further condensation re- lose fibers is markedly less reactive than that of the inorganic
action with the substrate hydroxyl groups, but also self- oxides onto which the silanes are usually applied [1,31,32].
condensation to form a polysiloxane network on the surface Furthermore, the cellulose –O–Si–O– linkage may undergo
[12–15]. Moreover, in the aqueous medium, the (partially reversible hydrolysis, which can strip off some of the an-
or totally) hydrolyzed silanes constitute a rather reactive chored silane from the surface. In order to prevent this draw-
system, which evolves with time as a result of the conden- back, it is necessary to ensure a high surface coverage, so
sation of the silanol groups both with each other and with that the silane links to numerous surface sites, and the ad-
alkoxy groups, to form dimeric and oligomeric structures sorbed silanol groups are close enough to each other to un-
[4,12,16–18], viz.: dergo self-condensation. The ensuing bidimensional silox-
ane network will therefore minimize any desorption of the
coupling agent.
, A better knowledge of the structural changes that occur in
the solution during silane hydrolysis and self-condensation
was deemed necessary for the selection of appropriate con-
ditions favoring the physical adsorption of the silane onto an
organic substrate.
.
2. Experimental

Further condensation reactions lead to gel-like networks, 2.1. Materials


which precipitate in the form of colloidal particles. Both
the hydrolysis and the condensation reactions of the silanol The cellulose fibers used in this work were commercial
groups are affected by the structure of the organic part of microcrystalline fibers (TECHNOCEL-150DM). Their av-
the silane and by the medium composition (temperature, pH, erage length was about 50 µm and their specific surface
concentration, amounts of water and catalyst) [19,20]. The area, measured by the BET technique using nitrogen as the
extent of silane hydrolysis and its degree of oligomerization adsorbed gas, was about 2.5 m2 /g. The organofunction-
are known to exert a great influence on the silane adsorp- al trialkoxysilanes γ -methacryloxypropyltrimethoxysilane
tion and on its configuration at the substrate surface [21–23]. (MPS), γ -aminopropyltriethoxysilane (APS), and γ -di-
Therefore, these two parameters must be judiciously con- ethylenetriaminopropyltrimethoxysilane (TAS), were indus-
trolled in order to optimize the surface modification with trial products of high purity used as received from OSI-
silane coupling agents. WITCO. Their structures and NMR chemical shift assign-
Even though the hydrolysis and condensation reactions ments are given in Tables 1 and 2, respectively. All other
of silanes have been investigated using different spectro- reagents and solvents were commercial high-purity prod-
scopic techniques such as FTIR [11,24], Raman [25], and ucts, used as received.
NMR spectroscopy [16,21,23,26], the multitude of reactions
likely to occur and their dependence on solvent composi- 2.2. Hydrolysis of silane coupling agents
tion makes it necessary to control the conduct of these re-
actions carefully to favor silane adsorption on the substrate The hydrolysis and condensation reactions of the three
surface. silanes, carried out in a mixture of ethanol-d6 and deuterated
In previous studies, we showed that the treatment of cellu-
lose fibers with appropriate silane coupling agents improved Table 1
the mechanical properties of the ensuing composite mate- The silane coupling agents used in this work
rials [27–29]. These effects were explained in terms of the
Silane Formula
ability of the functional group borne by the silane moiety to
bridge covalently with the polymeric matrix, thus increasing
the fiber/matrix adhesion. However, the actual quantifica-
APS
tion of silanes adsorbed on lignocellulosic substrates and
the mechanisms of their mutual condensation reactions have
been scantily reported [30].
The present investigation deals with a more fundamental TAS
approach, aiming at establishing the kinetics of the hydroly-
sis and the structural changes occurring in a silane solution
after the addition of water. These modifications are then cor-
MPS
related with the evolution of the amount of silane adsorbed
M.-C. Brochier Salon et al. / Journal of Colloid and Interface Science 289 (2005) 249–261 251

water CD3 CD2 OD/D2 O: 80/20 (w/w) at a concentration of sions (5% (w/w) with respect to the solvent). The ensuing
10% (w/w), were followed in situ at 25 ◦ C by 1 H, 13 C, and suspensions were maintained under stirring for 2 h. Then, af-
29 Si NMR. ter the cellulose particles had been isolated by centrifugation
at 2500 rpm for 20 min, the quantity of adsorbed coupling
2.3. NMR kinetic set up agent was determined using FTIR spectroscopy for MPS and
colorimetry in the case of APS and TAS. The detailed de-
1 H, 13 C,
and 29 Si NMR spectra of the silane solutions scription of these experimental conditions has been reported
were run on a Varian UNITY 400 spectrometer equipped elsewhere [27].
with a 5-mm probe and operating at 399.956, 100.572, and
79.455 MHz, respectively. All chemical shifts, relative to
TMS, were measured with a coaxial insert tube contain- 3. Results and discussion
ing the TMS solution as the external reference. The spectral
widths were calibrated to obtain the best resolution for pro- 3.1. Hydrolysis followed by NMR
ton and carbon. For silicon, a 10-mm BB probe was used
to optimize the signal/noise ratio for minimum acquisition 3.1.1. APS hydrolysis
times in relation to the specific reaction rates. The spectral The peak assignments of the 1 H and 13 C NMR spectra
width was 12 kHz and the relaxation delay 20 s with pro- of 10% (w/w) APS solution in ethanol-d6 are given in Ta-
ton decoupling only applied during acquisition time, to avoid ble 2. The commercial APS was highly pure, since only
negative nOe effects. The T1 measurements were made with traces of nondeuterated free ethanol were detected, probably
the inversion–recovery method. All compounds and solvents arising from minimal hydrolysis of APS when it was mixed
were weighed directly into clean tubes. The silane was first with ethanol-d6 . Immediately after the addition of deuter-
added to the ethanol-d6 and the zero reaction time was set ated water, the APS hydrolysis began, as revealed by the
immediately after the addition of D2 O. The kinetic studies formation of free ethanol-OD, giving rise to the appearance
were carried out directly in NMR tubes by following the evo- of two characteristic peaks at 1.19 and 3.62 ppm. The kinet-
lution of the relevant NMR signals. ics of the reaction was followed by measuring the changes
Solid state 29 Si CP/MAS NMR measurements were car- in the intensity of the peaks in 1 H and 13 C NMR spectra,
ried out with a Bruker Avance 400 operating at 79.490 MHz as shown in Figs. 1 and 2. For 1 H NMR study the evo-
with samples in 7-mm rotors of ZrO2 . Magic angle spin- lution of the peaks at 3.9–4.0 and 3.62 ppm, correspond-
ning was performed at a 3-kHz spinning rate. The construc- ing to –Si–(OCH2 CH3 )3 and CH3 CH2 –OD, respectively,
tor pulse program cp3lev was used with a 25-kHz spec- was set up (see Fig. 1). The plots reported in Fig. 3 show
tral width, 5-s repetition time, and 2-ms contact time. The that the rate of APS hydrolysis was relatively high in the
29 Si scale was calibrated by the external standard M Q
8 8 ethanol/water (80/20) mixture. In fact, 50 and 90% conver-
(−109.80 ppm as the highest field signal) [33]. sions were reached within 40 min and 3 h, respectively.
In addition to the progressive decrease in the intensity of
2.4. Adsorption isotherms the peaks at 1.25 and 3.85 ppm, corresponding to the starting
ethoxysilane, and the corresponding increase in those aris-
The adsorption isotherms were obtained by adding dif- ing from the liberated ethanol-OD at 1.19 and 3.62 ppm, the
ferent amounts of the chosen silane, dissolved in a mixture 1 H NMR spectra presented in Fig. 1 revealed that all of the

of ethanol and water 80/20 (w/w), to the cellulose suspen- protons of the organic part of the silane underwent a con-

Table 2
NMR chemical shifts of the coupling agents used here, in ethanol-d6 at 25 ◦ C
Name δ 29 Si δ 13 C/δ 1 H
R= αCH2 βCH2 γ CH2 O–CH2 –CH3 O–CH3 R
APS −45.3 ppm 8.8 ppm 27.6 ppm 45.6 ppm 59.4 ppm
19.1 ppm
NH2 – 0.63 ppm 1.6 ppm 2.63 ppm 3.84 ppm 4.995 ppm
1.24 ppm
MPS −42.9 ppm 6.32 ppm 23.3 ppm 67.6 ppm 51.1 ppm (a) 126.0 ppm, (b) 137.2 ppm,
(c) 18.0 ppm, (d) 168.3 ppm
H2 C=C(CH3 )–COO– 0.27 ppm 1.3 ppm 3.7 ppm 3.15 ppm (a) 5.18 and 5.68 ppm,
(c) 1.53 ppm
TAS −42.4 ppm 8.0 ppm 23.9 ppm 53.6 ppm 51.2 ppm (a) 42.2 ppm, (b) 53.1 ppm,
(c) 49.9 ppm, (d) 49.8 ppm
NH2 –(CH2 –CH2 –NH)2 – 0.02 ppm 1.0 ppm 1.96 ppm 2.94 ppm (a) 2.11 ppm, (b) 2.05 ppm,
(c) 2.09 ppm, (d) 2.09 ppm,
NH–NH2 4.19 ppm
252 M.-C. Brochier Salon et al. / Journal of Colloid and Interface Science 289 (2005) 249–261

Fig. 1. 1 H NMR spectra of APS hydrolysis.

Fig. 2. 13 C NMR spectra of APS hydrolysis.


M.-C. Brochier Salon et al. / Journal of Colloid and Interface Science 289 (2005) 249–261 253

bons, the signals corresponding to free and partially


hydrolyzed APS were not observed.
(iv) The ethoxysilane groups gave three superimposed sharp
peaks, the most intense corresponding to free APS and
the other two to its partially hydrolyzed structures.

In conclusion, the 1 H and 13 C spectra showed the same


tendencies, except for the transitory partially hydrolyzed
APS, which was clearly detected only in the 13 C NMR spec-
tra.

3.1.2. TAS hydrolysis


Thanks to its particular structure (three amine functions,
one primary and two secondary), TAS is an interesting
silane coupling agent that can interact with OH-rich surfaces
through its hydrolyzed silanol groups, leaving the amino-
containing moieties free to react with the polymeric matrix.
The 1 H and 13 C NMR spectra of pristine TAS (before the
addition of water) revealed that about 20% of the methoxy
Fig. 3. Reaction rate of free alcohol release for silane compounds. groups of TAS had already been hydrolyzed, because of the
well-known high reactivity of this compound toward adven-
tinuous evolution, characterized by the progressive loss of titious water during handling and storage. As with APS, the
the peak multiplicity, accompanied by a broadening of these hydrolysis of TAS was followed by the evolution of the pro-
peak and a slight frequency shift. These changes probably ton peaks corresponding to both the methyl groups attached
arose from silanol self-condensation reactions, leading to the to the silane (–Si–(OCH3 )3 ) and those liberated during the
formation of dimeric and oligomeric structures. In order to hydrolysis (CH3 –OD), appearing at 2.95 and 2.7 ppm, re-
investigate the evolution of these structures, a detailed 13 C spectively. The rate of TAS hydrolysis was considerably
NMR analysis was performed. In addition to the decrease in higher than that of APS, since the reaction was completed
the starting ethoxy peaks (19 and 59.5 ppm) and the increase in 1 h, as shown in Fig. 3. This reaction was also followed
in those coming from free ethanol-OD, at 18 and 57.5 ppm, by 13 C NMR spectroscopy, which displayed the following
already monitored by 1 H NMR, the following changes were features:
observed, as shown in Fig. 2:
(i) The disappearance of the methoxy group of the ini-
(i) A narrow doublet was detected at 8.78 ppm, accom- tial silane at 51.19 ppm and the concomitant formation
panied by several peaks between 9.5 and 12 ppm. The of free methanol, as indicated by the peak growing at
sharp structure of these peaks and their quasi-total dis- 49.79 ppm.
appearance after 3 h reaction were attributed to par- (ii) The emergence of two individual peaks at 8.5 and
tially hydrolyzed silane. On the other hand, the very 9.5 ppm, assigned to monomeric hydrolyzed TAS.
broad featureless signal (between 14 and 11 ppm) prob-
(iii) The presence of a broad peak at 13 ppm, attributed to
ably corresponded to CH2 (α) relative to the condensed
the CH2 (α).
oligomeric structures. Over a long reaction time (24 h
(iv) The broadening of the CH2 (β) and CH2 (γ ) peaks, indi-
or more), the CH2 (α) signal shifted toward a more uni-
cating the occurrence of the oligomerization reaction in
form peak centered at 12.5 ppm.
(ii) The CH2 (β) resonance split into three peaks close parallel to the hydrolysis (see Fig. 4).
to each other. The first two, at 26.3 and 26.85 ppm,
appeared at the beginning of the reaction, then in- The rapid evolution of this system impeded any precise
creased, and finally disappeared. These peaks were as- quantitative analysis using 13 C NMR, as the minimum time
sociated with free and partially hydrolyzed APS. The needed for a spectrum acquisition was 8 min. However,
third large peak, centered at 27.2 ppm, was attributed to 90 min after the water addition, the different monomeric
the oligomeric structures. species had totally disappeared from the solution, and col-
(iii) The same observations applied to the CH2 (γ ) res- loidal particles, corresponding to insoluble networks, started
onance, namely loss of resolution and enlargement to precipitate. Obviously, the presence of three amino func-
of peaks, accompanied by a slight shift in frequency tions within the silane structure greatly enhanced the rate of
for the oligomeric structures. However, for these car- both the hydrolysis and the silanol condensation reactions.
254 M.-C. Brochier Salon et al. / Journal of Colloid and Interface Science 289 (2005) 249–261

Fig. 4. 13 C NMR spectra of TAS hydrolysis.

3.1.3. MPS hydrolysis broad peaks at 11, 23.5, and 67.4 ppm (Figs. 5 and 6, for 1 H
The methacryloxy group of MPS is very interesting be- and 13 C NMR spectra, respectively).
cause it can be copolymerized with alkenyl-based matrices,
commonly used in composite materials. The kinetic mea- 3.1.4. Kinetics of hydrolysis and estimate of the different
surements were made under the same conditions as with products
APS and TAS. As with TAS, the MPS hydrolysis generated The fact that the hydrolysis of MPS, APS, and TAS were
free methanol, giving rise to peaks at 2.9 and 49.71 ppm in performed under the same conditions of temperature, water
the 1 H and 13 C NMR spectra, respectively. The shoulder at concentration, and solvent composition allowed a quanti-
3.1 ppm was assigned to –Si–(OCH3 )3−n (OD)n of the par- tative comparison of their rates to be made. A first-order
tially hydrolyzed MPS. The rate of hydrolysis of MPS was treatment of the data (1) was tested for the removal of the
particularly low, since after 24 h, the conversion had only alkoxy groups and gave straight lines up to conversions ex-
reached 17%, as shown in Fig. 3. This slow reaction can be ceeding 95%, as shown in Fig. 7.
related to the hydrophobic character of MPS and to the rela-
d[alkoxy] [alkoxy]
tive proportion of water in the medium. Attempts to increase − = Kh [alkoxy][H2 O] ⇒ ln
dt [alkoxy]0
the amount of water above 30 wt% led to phase separation
of the solution, whereas the addition of a small amount (1% = −Kh [H2 O]t = −Kh t.
(w/w) with respect to silane) of triethyl amine (TEA) greatly The apparent first-order constants Kh , calculated from the
accelerated the reaction [35,36], as clearly shown in Fig. 3. slope of each straight line, were found to be 0.0085, 0.378,
As with the other silanes, the hydrolysis was accompanied 0.76, and 5.04 min−1 for MPS, MPS-TEA, APS, and TAS,
by a condensation reaction among the silanol groups to give respectively.
oligomeric structures. The evidence of the presence of such With all these reactions, the rate of hydrolysis was deter-
structures was provided by the broadening of the different mined by measuring the changes in the peak intensity of the
peaks in the 1 H and 13 C NMR spectra. The monomeric hy- CH2 of the ethanol released by APS and of the CH3 of the
drolyzed MPS were characterized by sharp peaks at 6.47, methanol liberated by MPS and TAS, taking as reference the
23.34 and 67.67 ppm, whereas the oligomeric species gave corresponding areas of the initial silanes (Fig. 3).
M.-C. Brochier Salon et al. / Journal of Colloid and Interface Science 289 (2005) 249–261 255

Fig. 5. 1 H NMR spectra of TEA-catalyzed MPS hydrolysis.

For APS and MPS, we assessed both the extent of hy- the polycondensation rates, but the concentration of partially
drolysis and the amount of condensation products. Since hydrolyzed species did not exceed the maximum obtained in
MPS is a methoxy derivative, its 1 H NMR spectra display the absence of TEA.
a singlet and, consequently, it was possible to distinguish With TAS, the hydrolysis kinetics could be followed only
the CH3 protons of the initial compound from those of the by 1 H NMR, because the reaction celerity impeded any
free methanol and those of the partially hydrolyzed silane quantitative use of 13 C NMR spectra, as mentioned above.
(Figs. 8 and 9). In the 13 C NMR spectra, the CH2 (α) gave Partially hydrolyzed species were observed, but could not
well separated peaks. The good overlapping of the curves be quantified.
obtained from 1 H and 13 C NMR data indicated that under
our conditions, 13 C NMR spectroscopy was a reliable tech- 3.1.5. 29 Si NMR study
nique for a quantitative study and the hydrolysis of MPS and 29 Si NMR spectroscopy is of great interest to acquire
APS was followed accordingly (Figs. 8 and 9). The prod- further information about the structure of the oligomeric
ucts of the partial hydrolysis could not be observed in the 1 H species. The chemical shift of silicon is determined by the
NMR spectra because of the peak width associated with the chemical nature of its neighbors, namely, the number of
CH2 multiplicity (quadruplet), whereas the absence of this siloxane bridges attached to a silicon atom, and M, D, T, and
overlapping feature in the 13 C spectra allowed these inter- Q structures are the commonly used notation corresponding
mediate products to be quantified through the CH2 (α). The to one, two, three, and four Si–O– bridges, respectively.
same behavior was observed with these two silanes, viz., 29 Si NMR also provides a way to follow the hydrolysis

(i) the amount of partial hydrolysis products increased up and the condensation reaction of silicon alkoxides. For ex-
to a maximum value of 8 and 18% for MPS and APS, re- ample, a T structure bears only one organic Si–R side group
spectively, and then (ii) self-condensation started and its rate and the three siloxane bridges could be differentiated be-
increased progressively, while the fraction of hydrolyzed tween Si–OSi and Si–OR groups. With the nomenclature of
species decreased accordingly. It is worth noting that no Mi , Di , Ti , and Qi , taken from Glasser and Wilkes [37–41],
condensation reaction was observed before the extent of hy- where i refers to the number of –O–Si groups bound to the
drolysis reached its maximum value, and that, with MPS, silicon atom of interest, Ti corresponds to R–Si(–OSi)i –
the presence of TEA greatly enhanced the hydrolysis and OR (3−i) . Four T peaks can be present, namely T0 (−37 to
256 M.-C. Brochier Salon et al. / Journal of Colloid and Interface Science 289 (2005) 249–261

Fig. 6. 13 C NMR spectra of TEA-catalyzed MPS hydrolysis.

Fig. 7. First-order kinetic curves of silane coupling agents hydrolysis.


Fig. 8. MPS hydrolysis kinetics—TEA addition at 24 h, as deduced from
the 1 H and 13 C NMR spectra. Data deduced from 1 H NMR: black forms;
those obtained from the 13 C NMR: gray forms.

−39 ppm), T1 (−46 to −48 ppm), T2 (−53 to −57 ppm),


and T3 (−61 to −66 ppm) [41]. It is also well known that the First, 29 Si NMR measurements were made with a 5-mm
transition from Si–OR to the corresponding silanol Si–OH probe, but the relaxation times of different species were
gives a shift to lower fields (because of the electron-donating too long to give well-resolved spectra during kinetic ob-
properties of OR) and the extent of the shift increases with servations and only long-lived species could be detected.
the alkyl group molecular weight [42–44], e.g., 5–6 and The 29 Si NMR spectra of APS in ethanol/water solution
2–3 ppm for ethyl and methyl groups, respectively. are shown in Fig. 10. Based on the literature data, the sin-
M.-C. Brochier Salon et al. / Journal of Colloid and Interface Science 289 (2005) 249–261 257

(a)

Fig. 9. APS hydrolysis kinetics, as deduced from the 1 H and 13 C NMR


spectra. Data deduced from 1 H NMR, black forms; those obtained from
13 C NMR, gray forms.

gle peak at −40 ppm observed after 25 min reaction was


assigned to monomeric hydrolyzed silane. Usually, the hy-
drolysis leads to a small downfield shift of about 5–6 ppm.
After 8 h hydrolysis, the spectrum showed two peaks cen-
tered at −55 ppm, assigned to T2 units, followed by a small
peak at −61 ppm and a set of peaks ranging from −60
to −64 ppm, which were assigned to T2 and T3 structures
(b)
[16,34]. After the much longer reaction time of 4 days, the
29 Si NMR spectrum displayed the same peaks, but the in-

tensity of the T3 units at −63 ppm was more uniform and


had increased considerably, whereas the intensity of the T2
units at −55 ppm had decreased correspondingly. This evo-
lution resulted from the condensation reaction between the
silanol groups, which generated structures with higher mole-
cular weight. This result is in agreement with the observation
that the APS solution did not show any precipitation of col-
loidal particles of polysiloxane over a period of several days.
With MPS in the presence of TEA (1% (w/w) with re-
spect to silane), the reaction rates were still too high for the
detection of intermediate short-lived species. Therefore, the
MPS hydrolysis was followed using a ten fold decrease in
TEA, which produced a considerable reduction in the reac-
tion rates, as shown in Figs. 11 and 12. All the observations (c)
related to 1 H and 13 C NMR spectroscopy were qualitatively
Fig. 10. 29 Si NMR spectra of APS hydrolysis after reaction times of (a) 25
unchanged, but, after 48 h, some MPS was still present; i.e., min, (b) 6 h, and (c) 4 days.
the yield of CH3 OD had not reached 100%. These condi-
tions were adopted for a kinetic study followed by 29 Si NMR
spectroscopy, with 3 h acquisition time and 2 h delay be- the end of the run. The T2 units (centered at −59 ppm) were
tween each spectrum (with a 20 s d1 relaxation delay). The not detected at the beginning of the reaction, which is in
spectra obtained are given in Fig. 13. During the first three agreement with the 13 C NMR study, and only appeared at
hours, the major species was still MPS (T0R ), detected at later stages. Then after the first 3 h the broad peaks between
−45.4 ppm. The hydrolyzed moieties (T0H ) and dimeric units −65 and −70 ppm, attributed to the T3 units, appeared and
(T1 ), respectively at −40 and −51 ppm, were also present, increased with time. The observations deduced from 29 Si
but in much smaller amounts. The T1 signals then increased NMR study (Fig. 12) corroborated the 1 H, 13 C NMR results
with reaction time, and then decreased and disappeared at (Fig. 11), namely,
258 M.-C. Brochier Salon et al. / Journal of Colloid and Interface Science 289 (2005) 249–261

Table 3
T1 measurements by inversion recovery method
13 C

Product CH2α CH2β CH2γ CH3methacryl CH3 OSi CH3 ODfree


Initial MPS 2.173 2.082 1.973 5.236 11.700
Condensed 0.230 0.241 0.212 1.600
MPS
29 Si

T0H T0R T1 T2 T3
Initial MPS 20
Condensed 16.4 12.5–12.6
MPS

based NMR spectra. This shift is probably due to the differ-


ent experimental conditions versus relaxation times between
the two nuclei. By some relaxation T1 measurements, on the
Fig. 11. MPS + εTEA kinetic curves, as deduced from the 1 H and 13 C pristine silane, the final polycondensated results for 13 C and
NMR spectra. Data deduced from 1 H NMR, black forms; those obtained 29 Si were realized and the results are summarized in Table 3;
from 13 C NMR, gray forms. they show clearly that if the 13 C experimental conditions
were quantitative for the CH2 (α) signal, the chosen 29 Si ex-
perimental conditions (d1 = 20 s) were fatal to T0 species
detection below some concentration. A relaxation delay of
100 s would be much more appropriate, but in this case the
number of scans would be insufficient to obtain in 3 h acqui-
sition a well-resolved signal for T3 polycondensated units
because of the signals polydispersity.

3.2. Adsorption isotherm of silanes

The adsorption isotherm of nonprehydrolyzed MPS (0 h


of hydrolysis), prehydrolyzed MPS (24 h of hydrolysis) and
nonprehydrolyzed MPS-TEA (0 h of hydrolysis) are pre-
sented in Fig. 14, which shows that the amount of adsorbed
silane increased with increasing initial concentration and
reached a plateau situated at about 0.17 × 10−3 , 0.24 × 10−3 ,
and 0.37 × 10−3 mol of the adsorbed silane per gram of cel-
lulose, respectively. Then, with further increase in the silane
Fig. 12. MPS + εTEA kinetic curves, as deduced from the 29 Si NMR spec-
concentration, at C0 above about 0.12 mmol/l, the adsorp-
tra.
tion started to rise again. The shape of the isotherm sug-
gested that the silane adsorption proceeded through a mono-
(i) The rate of decrease in initial silane units T0R was con- layer formation, followed by multilayer mechanism [45]. At
firmed. the plateau region, the silane adsorption increased by 41%
(ii) The hydrolyzed species T0H , increased up to a maximum after MPS aging during 24 h, and became more than the
value and then disappeared. double after the addition of TEA. Based on the previous
(iii) Intermediate T1 , T2 species, and T3 units were detected. study, this enhancement could be explained by the genera-
(iv) The latency delay for the T2 and T3 beginning forma- tion of silanol groups, which promoted specific interactions
tion, as far as the maximum of the hydrolyzed unities through hydrogen bonding with the surface hydroxyl groups
T0H was reached. of cellulose. Furthermore, the growth of oligomeric struc-
(v) The continuous increase of the T3 condensed oligomeric tures from silanol self-condensation, would also enhance the
entities occurred to the detriment of the other intermedi- adsorption, since each adsorbed oligomer would imply the
ates. presence of several silane molecules. In the case of unhy-
drolyzed MPS (0 h aging), the adsorption was driven by both
Nevertheless a time shift was noticed between the curves dispersive interactions and polar ones, the latter arising be-
built up from 13 C and 29 Si NMR studies. In fact, after 48 h tween the C=O groups of the MPS and the surface hydroxyl
reaction, the initial unhydrolyzed silane was still present in groups of cellulose. This hypothesis is supported by the fact
13 C NMR, while it disappeared after only 36 h in silicone- that the C=O vibration in the FTIR spectrum of cellulose
M.-C. Brochier Salon et al. / Journal of Colloid and Interface Science 289 (2005) 249–261 259

Fig. 13. 29 Si NMR spectra of εTEA-catalyzed hydrolysis of MPS.

Fig. 15. Evolution of FTIR band corresponding to C=O function of MPS.

Fig. 14. Adsorption isotherms of MPS. the surface OH groups were engaged with Si–OH moieties
through stronger hydrogen bonds.
The adsorption isotherms of APS and TAS were con-
structed before and after 1 h of the silane hydrolysis. As
treated with MPS (0 h aging) gives two peaks at 1720 and in the case of MPS the adsorption isotherms displayed a
1710 cm−1 , corresponding to free and OH-bound C=O, re- plateau region followed by a second buildup of the adsorp-
spectively (Fig. 15). On the other hand, in the presence of tion above a given concentration (Fig. 16). The plateau val-
TEA, the C=O band was dominated by free species, because ues were at about 0.91 and 0.6 mmol/g for APS and TAS,
260 M.-C. Brochier Salon et al. / Journal of Colloid and Interface Science 289 (2005) 249–261

ing the self-condensation of the ensuing silanol groups. This


is particularly interesting, since the grafting efficiency (both
in terms of product configuration at the surface and quantity
of adsorbed molecules) can be modulated, depending on the
envisaged extent of the modification.
Work is in progress to examine the solid state 29 Si NMR
of these systems and their nanomorphology by AFM.

Acknowledgment

The authors thank OSI-WITCO Chemicals for the gift of


the silane coupling agents used in this work.

References
Fig. 16. Adsorption isotherms of APS and TAS.
[1] E.P. Plueddeman, Silane Coupling Agents, second ed., Plenum, New
York, 1991.
respectively. These relatively high values, compared with
[2] K. Weaver, J.O. Stoffer, E.D. Day, Polym. Compos. 16 (1995) 161.
those for MPS or MPS-TEA, could be associated, on the [3] A. Sabata, W.J. Van Ooij, R.J. Koch, J. Adhesion Sci. Technol. 11
one hand with the faster hydrolysis, which leads to a higher (1993) 1153.
proportion of silanol groups and oligomeric structure, and [4] E. Papirer, H. Balard, J. Adhesion Sci. Technol. 4 (8) (1990) 653.
on the other hand, with the presence of the amino func- [5] C.W. Chu, D.P. Kirby, P.D. Murphy, J. Adhesion Sci. Technol. 7 (5)
(1993) 417.
tions, which favored the interaction between the coupling
[6] M.A. Rodriguez, M.J. Liso, J. Mater. Sci. 34 (1999) 3867.
agents and the cellulose substrate, thanks to their ability [7] H. Ishida, J.L. Koenig, J. Colloid Interface Sci. 64 (3) (1978) 565.
to establish of hydrogen bonds. Even though TAS contains [8] S. Navroj, S.R. Culler, J.L. Koenig, H. Ishida, J. Colloid Interface
three amino functions, its plateau adsorption level was lower Sci. 97 (2) (1984) 309.
than that of APS. This apparent anomaly was attributed to [9] E.P. Plueddeman, J. Paint Technol. 40 (1968) 516.
[10] E.P. Plueddeman, Applied Polymer Symposium, No. 19, Wiley, New
the extremely high reactivity of TAS toward hydrolysis and
York, 1972, pp. 25–90.
condensation reactions, which generated a high proportion [11] H. Ishida, J.L. Koenig, Appl. Spectrosc. 32 (1978) 469.
of three-dimensional network (T structure), which was less [12] M.W. Daniels, L.F. Francis, J. Colloid Interface Sci. 205 (1998) 191.
readily adsorbed, compared with linear structures. More- [13] N. Nishiyama, K. Horie, T. Asakura, J. Colloid Interface Sci. 129
over, these three-dimensional networks tended to precipitate (1989) 113.
[14] K.C. Vrancken, L.D. Coster, P.V.D. Voort, P.J. Grobet, E.J. Vansant,
as colloidal gel-like particles with further condensation, and J. Colloid Interface Sci. 170 (1995) 71.
thus become unavailable for adsorption. [15] K.C. Vrancken, P.V.D. Voort, I.G. D’Hamers, E.J. Vansant, G.P. Gro-
It is interesting to note that, even though the cellulose sur- bet, J. Chem. Soc. Faraday Trans. 88 (1992) 3197.
face is less reactive towards silanes than inorganic oxides, [16] F. Beari, M. Brand, P. Jenkner, R. Lehnert, H.J. Metternich,
the adsorption efficiency was relatively high once the silane J. Monkiewicz, H.W. Siesler, J. Organomet. Chem. 625 (2001) 208.
[17] J.K. Crandall, C. Morel-Fourrier, J. Organomet. Chem. 489 (1995) 5.
coupling agents had undergone (partial) hydrolysis. Thus, [18] H.J. Kang, W. Meesiri, F.D. Blum, Mater. Sci. Eng. A 126 (1990) 773.
the fraction of the adsorbed silane at the beginning of the [19] F.D. Osterholz, E.R. Pohl, J. Adhesion Sci. Technol. 6 (1992) 127.
plateau, with respect to its initial concentration, was about [20] F.D. Blum, W. Meesiri, H.J. Kang, J.E. Gambogi, J. Adhesion Sci.
14, 40, 67, and 44% for MPS (0 h aging), MPS-TEA, APS, Technol. 5 (1991) 479.
and TAS, respectively. These results emphasize the poten- [21] M.W. Daniels, J. Sefcik, L.F. Francis, A.V. McCormick, J. Colloid In-
terface Sci. 219 (1999) 351.
tiality of using trialkoxysilane coupling agent as a relatively [22] N. Nishiyama, N. Horie, T. Asakura, J. Colloid Interface Sci. 129
simple method to modify the surface of cellulose substrates. (1989) 113.
[23] S. Savard, L.P. Blanchard, J. Léonard, R.E. Prud’homme, Polym.
Composites 5 (1984) 11.
4. Conclusions [24] C.H. Chiang, H. Ishida, J.L. Koenig, J. Colloid Interface Sci. 74 (1980)
396.
[25] P.T.H. Shih, J.L. Koenig, Mater. Sci. Eng. 20 (1975) 137.
The study of the rates and mechanisms of (i) the hydroly- [26] N. Nishiyama, T. Asakura, N. Horie, J. Colloid Interface Sci. 124
sis of three alkoxysilanes, (ii) the self-condensation of their (1988) 14.
hydrolyzed products, and (iii) the interaction of these species [27] M. Abdelmouleh, S. Boufi, A. Ben Salah, M.N. Belgacem, A. Gandini,
with the surface of cellulose fibers opens the way to the ratio- Langmuir 18 (2002) 3203.
[28] M. Abdelmouleh, S. Boufi, M.N. Belgacem, A.P. Duarte, A. Ben
nal use of silane coupling agents to functionalize the surface Salah, A. Gandini, Int. J. Adhesion Adhesives 24 (2004) 43–54.
of organic materials. Thus, in some cases, optimal conditions [29] M. Abdelmouleh, S. Boufi, M.N. Belgacem, A. Dufresne, A. Gandini,
were established favoring the hydrolysis reaction and limit- J. Appl. Polym. Sci., in press.
M.-C. Brochier Salon et al. / Journal of Colloid and Interface Science 289 (2005) 249–261 261

[30] M. Sain, Appl. Surf. Sci. 158 (2000) 92–103. [38] M.P.J. Peeters, W.J.J. Wakelkamp, A.P.M. Kentgens, J. Non-Cryst.
[31] A.V. Krasnolobodtsev, S.N. Smirnov, Langmuir 18 (2002) 3181. Solids 77 (1995) 189.
[32] L. Britcher, D. Kehoe, J. Matisons, G. Swincer, Macromolecules 28 [39] F. Babonneau, K. Thorne, J.D. Mackenzie, Chem. Mater. 1 (5) (1989)
(1995) 3110. 554.
[33] J. Brus, J. Dybal, P. Sysel, R. Hobzová, Macromolecules 35 (2002) [40] T.M. Alam, R.A. Assink, D.A. Loy, Chem. Mater. 8 (1996) 2366.
1253–1261. [41] R.H. Glasser, G.L. Wilkes, Polym. Bull. 19 (1988) 51.
[34] E.A. Williams, Recent Advances in Silicon-29 NMR Spectroscopy, [42] P.M. Henrichs, V.A. Nicely, Macromolecules 23 (1990) 3193–3194.
Annual Reports on NMR Spectroscopy, vol. 15, London, 1983. [43] H. Sun, S.J. Mumby, J.R. Maple, A.T. Hagler, J. Am. Chem. Soc. 116
[35] J.P. Blitz, R.S.S. Murthy, D.E. Leyden, J. Am. Chem. Soc. 109 (1987) (1994) 2978–2987.
7141. [44] Y.G. Hsu, I.L. Chiang, J.F. Lo, J. Appl. Polym. Sci. 78 (2000) 1179–
[36] G.R. Bogart, D.E. Leyden, J. Colloid Interface Sci. 167 (1994) 18. 1190.
[37] S.K. Young, W.L. Jarrett, K.A. Mauritz, Polymer 43 (2002) 2311– [45] J. Lyklema, Fundamentals of Interface and Colloid Science: I. Funda-
2320. mentals, Academic Press, San Diego, 1991.

You might also like