organic–inorganic hybrid materials derived from epoxy resin and polysiloxanes: synthesis and...

8
Organic–Inorganic Hybrid Materials Derived From Epoxy Resin and Polysiloxanes: Synthesis and Characterization C.F. Canto, L.A.S. de A. Prado, E. Radovanovic, I.V.P. Yoshida Instituto de Quı´mica, Universidade Estadual de Campinas, 13084-971 Campinas, Sa ˜ o Paulo, Brazil In this study, hybrid materials based on epoxy resin were prepared as transparent self-supported films by a sol–gel process. 4,4 0 -Diaminodiphenylmethane or oligo- meric epoxy resin were used as precursors, which were conveniently functionalized with trialkoxysilanes as end-groups. The effect of the introduction of poly (dimethylsiloxane) was also investigated. The hybrid films showed good thermal stability, a nondefined glass transition temperature, and a dense morphology without phase segregation. The tendency to a flat sur- face could be observed by atomic force microscopy. The hybrid films also showed good performance as coatings for glass plates, with an improved hydro- phobic character in comparison to neat epoxy resin. POLYM. ENG. SCI., 48:141–148, 2008. ª 2007 Society of Plastics Engineers INTRODUCTION Sol–gel manufacturing of organic–inorganic hybrids has gained increased interest in the last decade because of its versatility and the easy access to conditions in car- rying out the process [1]. This technique uses a mixture of organic and inorganic components at the molecular level, leading to materials with improved thermal and mechanical properties. Therefore, applications of this process can be found in different fields like membranes [2–11], organically modified electrolytes (Ormolytes) [12–15], electrochemical devices [5, 7, 16], and biomate- rials [7, 17–19] among others. One of the key areas of application of sol–gel processing is that of coatings, where many advantages over conventional methods have been claimed [20–22]. These include the ability to coat large and curved substrates in a cost-effective way, using simple deposition equipment, as well as the ability to obtain coatings with high homogeneity. With sol–gel processes it is possible to prepare materials that were inaccessible by other methods (e.g., organic–inorganic hybrid materials) [23, 24]. In this study, epoxy resin was used as the organic com- ponent in the preparation of organic–inorganic materials for metallic and glass surface coatings. Epoxy resins are one of the most important classes of thermosetting poly- mers, and are widely used in applications for high-per- formance adhesives and as matrix resins for advanced composite materials. Cured epoxy resins exhibit excellent adhesion to a variety of substrates; outstanding chemical and corrosion resistance [25–27]; excellent electrical insu- lation; high tensile, flexural, and compressive strengths; thermal stability; a wide range of curing temperatures; and also low shrinkage upon curing. However, in contrast to such desirable properties, epoxy networks are brittle and display low fracture toughness [28–30]. Improvement in the toughness of the epoxy resin can be achieved by the addition of a second component, such as a thermo- plastic or elastomer modifier. The modification of epoxy resins using polysiloxanes [31–34], alkoxysilanes [35–37], or polyhedral oligomeric silsesquioxanes [35–37] is attracting increasing attention in this field because of the low surface tension, good flexibility, nonflammability, and high resistance to thermal oxidation provided by these inorganic components. The toughness of epoxides could be improved by incorporating linear polysiloxanes [31]. The modification of DGEBA (poly(bis phenol A–co-epichlorohydrin))- based resin with PDMS and the use of epoxy-functional- ized linear polysiloxanes as modifiers for epoxides are some methods of incorporation of polysiloxanes in epox- ides through covalent bonds [31], while the stiffness can be improved by the dispersion of POSS moieties [38]. Siloxane-containing epoxy resins have also been uti- lized in formulations for paints [30] that exhibited improved thermal stability. These coatings also showed L.A.S. de A. Prado is currently at Institut fu ¨r Kunststoffe und Verbund- werkstoffe, Technische Universita ¨t Hamburg-Harburg, Denickerstraße 15, Geb. K, D-21073 Hamburg, Germany. E. Radovanovic is currently at Departamento de Quı ´mica, Universidade Estadual de Maringa ´, Avenida Colombo, 5790 CEP: 87020-900 Maringa ´, PR, Brazil. Contract grant sponsors: FAPESP, CNPq. Correspondence to: Inez Vale ´ria Pagotto Yoshida; e-mail: valeria@iqm. unicamp.br DOI 10.1002/pen.20931 Published online in Wiley InterScience (www.interscience.wiley.com). V V C 2007 Society of Plastics Engineers POLYMER ENGINEERING AND SCIENCE—-2008

Upload: cf-canto

Post on 06-Jul-2016

216 views

Category:

Documents


4 download

TRANSCRIPT

Organic–Inorganic Hybrid Materials DerivedFrom Epoxy Resin and Polysiloxanes:Synthesis and Characterization

C.F. Canto, L.A.S. de A. Prado, E. Radovanovic, I.V.P. YoshidaInstituto de Quımica, Universidade Estadual de Campinas, 13084-971 Campinas, Sao Paulo, Brazil

In this study, hybrid materials based on epoxy resinwere prepared as transparent self-supported films by asol–gel process. 4,40-Diaminodiphenylmethane or oligo-meric epoxy resin were used as precursors, whichwere conveniently functionalized with trialkoxysilanesas end-groups. The effect of the introduction of poly(dimethylsiloxane) was also investigated. The hybridfilms showed good thermal stability, a nondefinedglass transition temperature, and a dense morphologywithout phase segregation. The tendency to a flat sur-face could be observed by atomic force microscopy.The hybrid films also showed good performanceas coatings for glass plates, with an improved hydro-phobic character in comparison to neat epoxy resin.POLYM. ENG. SCI., 48:141–148, 2008. ª 2007 Society ofPlastics Engineers

INTRODUCTION

Sol–gel manufacturing of organic–inorganic hybrids

has gained increased interest in the last decade because

of its versatility and the easy access to conditions in car-

rying out the process [1]. This technique uses a mixture

of organic and inorganic components at the molecular

level, leading to materials with improved thermal and

mechanical properties. Therefore, applications of this

process can be found in different fields like membranes

[2–11], organically modified electrolytes (Ormolytes)

[12–15], electrochemical devices [5, 7, 16], and biomate-

rials [7, 17–19] among others. One of the key areas of

application of sol–gel processing is that of coatings,

where many advantages over conventional methods have

been claimed [20–22]. These include the ability to coat

large and curved substrates in a cost-effective way, using

simple deposition equipment, as well as the ability to

obtain coatings with high homogeneity. With sol–gel

processes it is possible to prepare materials that were

inaccessible by other methods (e.g., organic–inorganic

hybrid materials) [23, 24].

In this study, epoxy resin was used as the organic com-

ponent in the preparation of organic–inorganic materials

for metallic and glass surface coatings. Epoxy resins are

one of the most important classes of thermosetting poly-

mers, and are widely used in applications for high-per-

formance adhesives and as matrix resins for advanced

composite materials. Cured epoxy resins exhibit excellent

adhesion to a variety of substrates; outstanding chemical

and corrosion resistance [25–27]; excellent electrical insu-

lation; high tensile, flexural, and compressive strengths;

thermal stability; a wide range of curing temperatures;

and also low shrinkage upon curing. However, in contrast

to such desirable properties, epoxy networks are brittle

and display low fracture toughness [28–30]. Improvement

in the toughness of the epoxy resin can be achieved by

the addition of a second component, such as a thermo-

plastic or elastomer modifier. The modification of epoxy

resins using polysiloxanes [31–34], alkoxysilanes [35–37],

or polyhedral oligomeric silsesquioxanes [35–37] is

attracting increasing attention in this field because of the

low surface tension, good flexibility, nonflammability, and

high resistance to thermal oxidation provided by these

inorganic components.

The toughness of epoxides could be improved by

incorporating linear polysiloxanes [31]. The modification

of DGEBA (poly(bis phenol A–co-epichlorohydrin))-

based resin with PDMS and the use of epoxy-functional-

ized linear polysiloxanes as modifiers for epoxides are

some methods of incorporation of polysiloxanes in epox-

ides through covalent bonds [31], while the stiffness can

be improved by the dispersion of POSS moieties [38].

Siloxane-containing epoxy resins have also been uti-

lized in formulations for paints [30] that exhibited

improved thermal stability. These coatings also showed

L.A.S. de A. Prado is currently at Institut fur Kunststoffe und Verbund-

werkstoffe, Technische Universitat Hamburg-Harburg, Denickerstraße

15, Geb. K, D-21073 Hamburg, Germany.

E. Radovanovic is currently at Departamento de Quımica, Universidade

Estadual de Maringa, Avenida Colombo, 5790 CEP: 87020-900 Maringa,

PR, Brazil.

Contract grant sponsors: FAPESP, CNPq.

Correspondence to: Inez Valeria Pagotto Yoshida; e-mail: valeria@iqm.

unicamp.br

DOI 10.1002/pen.20931

Published online in Wiley InterScience (www.interscience.wiley.com).

VVC 2007 Society of Plastics Engineers

POLYMER ENGINEERING AND SCIENCE—-2008

improved stability against corrosive media (alkaline and

acidic media).

In the present study, epoxy/polysilsesquioxane hybrid

materials were prepared by the reaction between an

uncured epoxy resin and an aminofunctionalized alkoxysi-

lanes, using tetrahydrofuran (THF) as solvent, or from the

reaction of 3-glycidoxypropyltrimethoxysilane and 4,40-diaminodiphenylmethane (DAF).

EXPERIMENTAL

Materials

Commercially available 3-aminopropyltrimethoxysilane,

SA1, 2-aminoethyl-3-aminopropyltrimethoxysilane, SA2,

3-glycidoxypropyltrimethoxysilane, SA3, and poly(dime-

thylsiloxane) with ��Si(CH3)2OH end-groups (Mn � 2200

g mol21) were supplied by Dow Corning do Brasil (Horto-

landia, SP, Brazil), while the DGEBA-based epoxy resin

poly(bisphenol-A-co-epichloridrin) of low molar mass and

4,40-diaminodiphenylmethane were purchased from Dow

Chemical (Sao Paulo, Brasil) and Fluka (Sao Paulo, Brasil),

respectively, and used as received. THF from Merck (Rio

de Janeiro, Brasil) was purified and dried in accordance

with standard literature procedures [39].

Measurements

The 29Si{1H} nuclear magnetic resonance (NMR) spec-

tra of the hybrid precursors were recorded with a Varian

NMR spectrometer (model Gemini 300), in CDCl3 as sol-

vent. The NMR-frequency of the 29Si nuclei is 59.6 MHz.

Fourier transform infrared (FTIR) spectra of hybrid films

were obtained with a Bomem MB-series spectrometer,

collecting 256 scans in the range from 4000 to 400 cm21

using KBr pellets, with a resolution of 4 cm21. The ther-

mal stabilities of these films were determined by thermog-

ravimetry (TGA) on a TA 2950 thermobalance, TA

Instruments, from 25 to 10008C, at a heating rate of 208Cmin21, under argon flow (100 cm3 min21, 99.999%). The

differential scanning calorimetry analyses (DSC) were

performed on a 2910 DSC TA Instruments. About 10 mg

of samples was heated at 108C min21 under an argon

flow, from 21508C to 2008C. The morphology of the

hybrid materials was analyzed by field emission scanning

electron microscopy (FESEM) using a JEOL JSM-6340F

operated at 5 kV. For this purpose, the films were cryo-

genically fractured, and the fracture surface was sputter-

coated with a very thin gold layer in a Bal-Tec MED 020

instrument. The images of the hybrid film surface were

obtained using atomic force microscopy (AFM) on a Dis-

coverer TMX 2010 AFM Scanner, by the noncontact

mode, with a silicon cantilever nanoprobe 70-mm long.

Contact angle measurements were carried out at room

temperature (258C) and at 65% relative humidity. A drop

of water was deposited on the film surface with a micro-

pipette. A digital camera was used to obtain the drop

image. Three drops from each sample were analyzed in

order to get an average contact angle.

Preparation of the Self-Supported Hybrid Films

Organic–inorganic hybrid films were obtained by an

epoxy-amine addition reaction followed by hydrolysis/

condensation reactions of the alkoxysilane end-groups.

The inorganic components were 3-aminopropyltrimethox-

ysilane, SA1, 2-aminoethyl-3-aminopropyltrimethoxysi-

lane, SA2, or 3-glycidoxypropyltrimethoxysilane, SA3,

and the organic components were the DGEBA-based ep-

oxy resin poly(bisphenol-A-co-epichloridrin) or DAF. Themolar proportion between the NH2 and the epoxy groups

was 1:1. THF solutions containing the components, as

described in Fig. 1, were stirred at 508C for 4 h. Aliquots

of the alkoxysilane hybrid precursor solution were iso-

lated for characterization. After this step, dibutyltin dilau-

rate (Sn, 1 wt%) was added as a catalyst for the hydroly-

sis/condensation reactions. The homogeneous solutions

were transferred to Petri dishes, which after 24 h gave

rise to hybrid polysiloxane (HSPi) transparent self-sup-

ported films. These films were dried under reduced pres-

sure at 708C, for 24 h. HPPi hybrids were prepared in a

similar way with the addition of 10 wt% of PDMS to the

FIG. 1. Structures of the precursors used for the syntheses of the

epoxy/silsesquioxane hybrids.

142 POLYMER ENGINEERING AND SCIENCE—-2008 DOI 10.1002/pen

system, prior to Sn catalyst addition. Self-supported films

prepared from the DGEBA-based epoxy resin poly(bi-

sphenol-A-co-epichloridrin) and DAF were also prepared

to be used as comparison in the thermal stability study;

these were coded as resin.

Preparation of the Hybrid Films forContact-Angle Measurements

In order to coat the glass plates, the same synthetic

approach for preparing the self-supported films was used.

An aliquot of the solution containing the mixture of or-

ganic and inorganic components and the catalyst, before

being transferred to the Petri dishes, was stirred and

heated for 30 min at 508C in order to promote an increase

in the viscosity of the solution due to formation of hybrid

siloxane oligomers. The resulting homogeneous solution

was cast on the surface of 0.3-mm glass plates using a

Kcontrol Coater equipment, model 101, and a standard

bar to deposit films of 12-mm thickness. The coated glass

plates were dried in the oven for 30 min at 708C.

RESULTS AND DISCUSSION

The Molecular Structure of the Organic–InorganicHybrid Films

In this investigation, the hybrid films were obtained by

the epoxy-amine addition reaction, followed by hydrolysis

and condensation reactions of trifunctional alkoxysilane

precursors, catalyzed by dibutyltin dilaurate, which led to

a HSPi network, or those also containing PDMS, HPPi, as

illustrated in Fig. 1.

Reaction of an epoxy group with a primary amine ini-

tially produces a secondary alcohol and a secondary

amine. The secondary amine, in turn, can also react with

an epoxy group to give a tertiary amine and two second-

ary hydroxyl groups. Primary amines react approximately

twice as fast as secondary amines, as the overall reaction

rate of an amine with an epoxy resin is also influenced

by the steric hindrance. Moreover, no competitive reac-

tions were detected between a secondary hydroxyl group

in the backbone and an epoxy group to give ether,

because of an equivalent epoxy/amine stochiometry in

the system. As an example, the reaction between DAF

and AS3 produced the water-sensitive precursor P3,

which was the product of the addition of the amine

groups from DAF to the epoxide groups from AS3. The

product was analyzed by 29Si{1H} NMR, as can be seen

in Fig. 2a. Only one narrow peak at 241 ppm [40], asso-

ciated with CSi(OCH3)3 was observed for this precursor.

The FTIR spectrum of P3 precursor showed typical

absorptions related to the n(C��H) of the Si��OCH3 at

2830 cm21 and an intense and narrow band at 1100

cm21, assigned to the n(Si��O��C) bonds of this group

[41], as can be seen in Fig. 2b. In addition, other

expected absorptions related to the organic segment were

also observed. Similar results were also obtained for the

other precursors.

The conversion of the precursors into organic–inor-

ganic hybrid materials proceeded via formation of silox-

ane (Si��O��Si) bonds. This process took place via for-

mation of Si��OH groups, which were produced by the

hydrolysis of the Si��OCH3 groups by moisture. The

reactions are illustrated by the Eqs. 1–3 [42]:

BSiAOCH3 þ H2O ! BSiAOHþ CH3OH� hydrolysis:

(1)

BSiAOHþBSiAOCH3 ! BSiAOASiB

þ CH3OH� condensation: ð2Þ

BSiAOHþBSiAOH!BSiAOASiB

þ H2O� condensation: ð3Þ

Therefore, the structure of these hybrids can be

described as organic segments interconnected by polysil-

sesquioxane nodes [43]. The polysilsesquioxane nodes

constitute the inorganic component of the organic–inor-

ganic hybrid material prepared in the present investiga-

tion. The organic moieties are the oligomers separating

the silicon atoms. Therefore, these hybrids can be

regarded as belonging to the category II [4], i.e., organic–

inorganic hybrid materials in which the organic and the

inorganic components are connected through covalent

bonds.

The infrared spectra of the hybrid materials (not

shown) were characterized by a broad absorption centered

FIG. 2. P3 precursor. (a) 29Si-{1H} NMR (CDCl3); (b) FTIR spectrum.

DOI 10.1002/pen POLYMER ENGINEERING AND SCIENCE—-2008 143

at 3370 cm21 associated with the n(O��H) of the alcohol

formed (resulting from the reaction between the epoxide

and amine groups) and residual BSi��OH groups pro-

duced by the hydrolysis of the alkoxysilanes. In addition,

a shoulder at 3270 cm21 was related with n(N��H) of the

formed secondary amine. The absorption peaks at �1620

and 1510 cm21 were associated with n(C¼¼C) of the aro-

matic ring, and a broad band centered at 1100 cm21

assigned to n(Si��O��Si) and n(C��O��C) absorptions.

Apart from these absorptions, a characteristic of PDMS at

1260 cm21, which corresponds to d(C��H) of the

��Si(CH3)2 repeating units [41], could also be observed

in the spectra of the HPPi hybrids.

Thermal Behavior of the Hybrid Films

The thermal stability of the hybrid materials was inves-

tigated by TGA and the results were compared to that

obtained from the resin. Four parameters were chosen to

characterize the thermal stability of these materials: the

temperature at which the thermal decomposition process

starts, Ti; the temperature associated to the maximum

decomposition rate, TMAX; and the percentages of the

weight loss at 4008C and 9008C, which are summarized

in Table 1.

The thermal degradation of polysiloxanes and polysil-

sesquioxane materials has been extensively discussed in

the literature [43–49]. While the degradation of linear

silicones occurs via depolymerization processes due to

inter- and intramolecular rearrangements [48, 49], the

degradation of polysilsesquioxanes and other polysilox-

ane-based networks takes place via scission and redis-

tribution of Si��C and Si��O bonds, yielding, at high

decomposition temperatures, mostly silicon oxycarbide

glasses [43, 50–52].

The weight loss in hybrid materials that contain silane

and/or silicones usually starts through the loss of volatiles

retained in the hybrid on account of incomplete polycon-

densation reactions of the residual silanol in the 100–

2508C range [53, 54].

The hybrids HSP3 and HPP3, derived from DAF, pre-

sented a higher weight percent of inorganic component,

which promoted a lower percentage of weight loss at Tiand TMAX, as reported in Table 1. The resin presented

lower Ti and TMAX values compared to the hybrids

derived from P3 and also a lower amount of residue at

9008C.The degradation of the hybrids showed a maximum in

the 353–4458C range, associated with the degradation of

the organic moiety, as well as of the Si��C bonds of the

siloxane component [44, 48, 49]. The plain epoxy resins

presented a maximum rate of the weight loss at 3958C,

TABLE 1. Values of the initial degradation temperature (Ti), weight

loss percent at Ti, temperature associated to maximum mass fluxes from

solid to vapor (TMAX), weight loss percent at 4008C and percent of

residues at 9008C.

Hybrid Ti (8C)

Weight

loss %

at Ti TMAX (8C)

Weight

loss % at

4008CResidue at

9008C (%)

HSP1 163 2 371; 445 24 25

HPP1 168 0.5 353; 445 24 33

HSP2 170 2 361; 439 32 26

HPP2 165 0.5 358; 445 30 25

HSP3 193 2 405 21 37

HPP3 192 0.5 412 20 37

Resin 177 2 395 43 7

FIG. 3. DTGA curves of the hybrid films: (a) HSP1 and HPP1; HSP2

and HPP2; and (c) HSP3 and HPP3. The DTGA curve of an entirely or-

ganic resin constituted by an equimolar mixture of DGEBA and DAF is

shown for the sake of comparison.

144 POLYMER ENGINEERING AND SCIENCE—-2008 DOI 10.1002/pen

which is comparable to other epoxy resins reported else-

where. All the hybrids prepared in this work were signifi-

cantly more stable at a higher temperature, which is in

line with previous studies on modified epoxides [38, 55,

56]. Furthermore, the DTGA curves (Fig. 3) clearly dem-

onstrate that the degradation of the DGEBA-DAF resin

took place much faster than the degradation of the

hybrids. This fact is supported not only by the intensity

of the peak associated with the main degradation step of

the resin, which was around three times more intense than

the peaks for the hybrids, but also by the higher weight

loss at temperatures ranging from 4008C up to 9008C.This behavior is also in accordance with the results

reported for similar organic–inorganic hybrid materials

based on epoxides modified with POSS [36] or other

modified epoxides [55, 56].

The DSC curves of the hybrids, Fig. 4, exhibited a

very broad glass transition from approximately 2808C to

1258C, which suggests the existence of a great number of

relaxations associated with the regions with different

FIG. 4. DSC traces of the hybrid films.

FIG. 5. FESEM micrographs of the cryogenic fracture surface of HSPi and HPPi hybrid films.

DOI 10.1002/pen POLYMER ENGINEERING AND SCIENCE—-2008 145

degrees of connectivity in the polysilsesquioxane

domains, promoted by the condensation reactions between

silanol groups. This fact reflects the high structural heter-

ogeneities of these hybrid materials.

However, no glass transition could be identified associ-

ated with the PDMS phase, in the curves of the HPPi

hybrids. This is strong evidence of a homogenous distri-

bution of the PDMS chains in the tridimensional structure

of the hybrids. Furthermore, postcuring of the resins could

not be detected, suggesting that no residual ��NH groups

reacted with residual epoxide rings, which could possibly

be present.

Morphological Analysis of the Hybrid Films

The morphology of the fracture surface of the hybrid

films was observed by FESEM, as can be seen in Fig. 5.

The hybrid films did not show phase segregation or pores,

indicating that the films were homogeneous at the resolu-

tion limit of the equipment. The features of the fracture

were similar to epoxy resins reported in the literature

[57]. As the PDMS chains were introduced in hybrids of

the HPPi series, the fracture surfaces became slightly

rougher and exhibited some irregular stripes in the matrix.

Although the fractures were fairly rough for the hybrids

containing PDMS, the morphology of these hybrids still

showed the characteristic lack of features of the cryogenic

fracture of neat epoxy films [34].

The topography of the surface of the films obtained

was characterized by AFM, which revealed different fea-

tures, as can be seen in Fig. 6. Table 2 shows the average

roughness, Ra, the peak-valley height, Rt, and the average

height, Rz. Hybrids of the series HSPi showed a tendency

toward a flatter surface, in relation to those containing

PDMS. The flatness was particularly pronounced in the

HSP1 film, which was derived from epoxy resin and an

alkoxysilane with short organic chains. The HSP2 surface

was a continuous spread of pit-and-trench features,

widths, depths, and round shapes, with a maximum peak-

valley height of 365 nm. The HSP3 hybrid also presented

a regular flat surface. Hybrids of the series HPPi showed

a uniform distribution of ridge-and-valley structures in

the surface. However, HPP3 presented relatively flat

ridges and rough valleys, which suggests segregation

with some domains probably having different PDMS

contents.

Contact Angle Measurements

Table 3 shows the contact angle for the hybrid films

deposited on glass plates. When PDMS was introduced,

the surface became more hydrophobic, as indicated by the

increase in the contact angle values. This characteristic

makes this material interesting to be applied on surfaces

in which antifogging and antisoiling effects are desirable.

FIG. 6. AFM micrographs of HSPi and HPPi hybrid films. [Color fig-

ure can be viewed in the online issue, which is available at www.

interscience.wiley.com.]

TABLE 2. Topographic image (5 � 5 lm2) of the hybrids.

Hybrids Ra (nm) Rt (nm) Rz (nm)

HSP1 0.8 15 2.8

HPP1 4 66 19

HSP2 38 365 169

HPP2 50 588 196

HSP3 1.8 20 8

HPP3 5 48 20

TABLE 3. Contact angle of the HSPi and HPPi hybrid films on glass

plates.

Hybrids Contact angle (8)

HSP1 62

HPP1 71

HSP2 51

HPP2 80

HSP3 65

HPP3 79

Resin 53

Glass without coating 25

146 POLYMER ENGINEERING AND SCIENCE—-2008 DOI 10.1002/pen

CONCLUSIONS

The sol–gel process was shown to be a very effective

technique to coat substrates with hybrid films based on

epoxy/siloxane material. The hybrid films exhibited good

thermal stability and slower degradation at temperatures

higher than 4008C. XRD results suggested that the hybrid

materials were amorphous, while DSC curves indicated

that they have a very complex structure. FESEM analyses

indicated that these materials were homogeneous,

although AFM suggested a possible phase separation in

the HPP3 hybrid.

ACKNOWLEDGMENTS

The authors acknowledge Prof. Carol H. Collins (IQ-

UNICAMP) for manuscript revision.

REFERENCES

1. R. Kasenann and H. Schmidt, New J. Chem., 18, 1117

(1994).

2. S.S. Kulkarni, S.M. Tambe, A.A. Kittur, and M.Y. Karidura-

ganavar, J. Membr. Sci., 285, 420 (2006).

3. Y.N. Yang and P. Wang, Polymer, 47, 2683 (2006).

4. N.W. Deluca and Y.A. Elabd, J. Polym. Sci. Part B: Polym.Phys., 44, 2201 (2006).

5. N.M. Jose and L.A.S. de A. Prado, Quim. Nova, 28, 281

(2005).

6. T.W. Xu, J. Membr. Sci., 263, 1 (2005).

7. L.C. Klein, Y. Daiko, M. Aparicio, and F. Damay, Polymer,46, 4504 (2005).

8. C.M. Wu, T.W. Xu, M. Gong, and W.H. Yang, J. Membr.Sci., 247, 111 (2005).

9. N.M. Jose, L.A.S. de A. Prado, and I.V.P. Yoshida, J.Polym. Sci. Part B: Polym. Phys., 42, 4281 (2004).

10. R.A. Zoppi and C.G.A. Soares, Adv. Polym. Technol., 21, 2(2002).

11. J.H. Kim and Y.M. Lee, J. Membr. Sci., 193, 209 (2001).

12. A. Evans, V.D. Bermudez, M.J. Smith, and D. Ostrovskii,

Electrochim. Acta, 52, 1542 (2006).

13. F.R. Zaggout, I.M. El-Nahhal, A.E.A. Qaraman, and N. Al

Dahoudi, Mater. Lett., 60, 3463 (2006).

14. J.A. Chaker, K. Dahmouche, C.V. Santilli, S.H. Pulcinelli,

V. Briois, P. Judeinstein, and A.F. Craievich, J. Non-Cryst.Solids, 352, 3457 (2006).

15. S. Grandi, A. Magistris, P. Mustarelli, E. Quartarone, C.

Tomasi, and L. Meda, J. Non-Cryst. Solids, 352, 273

(2006).

16. D.M. Tigelaar, M.A.B Meador, J.D. Kinder, and W.R. Ben-

nett, Macromolecules, 39, 120 (2006).

17. H. Fomg, S.H. Dickens, and G.M. Flaim, Dent. Mater., 21,520 (2005).

18. Y.A. Shchipunov and T.Y. Karpenko, Langmuir, 20, 3882(2004).

19. N. Moszner and U. Salz, Prog. Polym. Sci., 26, 535 (2001).

20. A.S. Hamdy, Prog. Org. Coat., 56, 146 (2006).

21. S.W. Duo, M.S. Li, M. Zhu, and V.C. Zhou, Surf. Coat.Technol., 200, 6671 (2006).

22. N. Carmona, M.A. Villegas, and J.M.F. Navarro, Thin SolidFilms, 458, 121 (2004).

23. D.R. Uhlmann, T. Suratwala, K. Davidson, J.M. Boulton,

and G. Trowee, J. Non-Cryst. Solids, 218, 113 (1997).

24. W. Boysen, A. Frattini, N. Pellegri, and O. Sanctis, Surf.Coat. Technol., 122, 14 (1999).

25. S.J. Garcia and J. Suay, Prog. Org. Coat., 57, 319 (2006).

26. S.J. Garcia, M.T. Rodriguez, K.A. Razzaq, J.J. Carpio, and

J.J. Saura, Prog. Org. Coat., 46, 121 (2003).

27. T.T.X. Hang, T.A. Truc, T.H. Nam, V.K. Oanh, J.B. Jorcin,

and N. Pebere, Surf. Coat. Technol., 201, 7408 (2007).

28. W.R. Ashcroft, ‘‘Curing Agents for Epoxy Resins,’’ in Chem-istry and Technology of Epoxy Resins, 1st ed., E. Bryan, Ed.,Chapman & Hall, London, Chapter 2, 37 (1993).

29. S.J. Shaw, ‘‘Additives and Modifiers for Epoxy Resins,’’ in

Chemistry and Technology of Epoxy Resins, E. Bryan, Ed.,Chapman & Hall, London, Chapter 4, 126 (1993).

30. X.M. Chen and B. Ellis, ‘‘Coatings and Other Applications

of Epoxy Resins,’’ in Chemistry and Technology of EpoxyResins, E. Bryan, Ed., Chapman & Hall, London, Chapter 9,

303 (1993).

31. S.A. Kumar, Z. Denchev, and M. Alagar, Eur. Polym. J.,42, 2419 (2006).

32. S. Ahmad, A.P. Gupta, E. Sharmin, M. Alam, and S.K. Pan-

dey, Prog. Org. Coat., 54, 248 (2005).

33. S.A. Kumar and T.S.N.S. Narayanan, Prog. Org. Coat., 45,323 (2002).

34. T.-H. Ho and C.S. Wang, Eur. Polym. J, 37, 267 (2001).

35. W.-G. Ji, J.-M. Hu, L. Liu, J.-Q. Zhang, and C.-N. Cao,

Surf. Coat. Technol., 201, 4789 (2007).

36. L. Matejka, O. Dukh, and J. Kolarik, Polymer, 41, 1449

(2000).

37. S. Wu, M.T. Sears, and M.D. Soucek, Prog. Org. Coat., 36,89 (1999).

38. Y.H. Liu, S.X. Zheng, and K.M. Nie, Polymer, 46, 12016(2005).

39. D.D. Perrin, W.L.F. Armarego, and D.R. Perrin, Purificationof Laboratory Chemicals, Pergamon, Oxford, Chapter 5, 557

(1983).

40. R.B. Taylor, B. Parbhoo, and D.M. Fillmore, ‘‘Nuclear Mag-

netic Resonance Spectroscopy,’’ in The Analytical Chemistryof Silicones, A.L. Smith, Ed., Wiley, New York, Chapter 12,

347 (1991).

41. L.J. Bellamy, The Infrared Spectra of Complex Molecules,Wiley, New York, Chapter 20 (1957).

42. C.J. Brinker and G.W. Scherer, Sol–Gel Science—ThePhysics and Chemistry of Sol–Gel Processing, Academic

Press, New York, Chapter 8 (1990).

43. M.A. Schiavon, S.U.A. Redondo, S.R.O. Pina, and I.V.P.

Yoshida, J. Non-Cryst. Solids, 304, 92 (2002).

44. G. Camino, S.M. Lomakin, and M. Lazzari, Polymer, 42,2395 (2001).

45. G. Camino, S.M. Lomakin, and M. Lageard, Polymer, 43,2011 (2002).

DOI 10.1002/pen POLYMER ENGINEERING AND SCIENCE—-2008 147

46. N.M. Jose, L.A.S. de A. Prado, and I.V.P. Yoshida,

J. Polym. Sci. Part B: Polym. Phys., 45, 299 (2007).

47. L.A.S. de A. Prado, E. Radovanovic, H.O. Pastore, I.V.P.

Yoshida, and I.L. Torriani, J. Polym. Sci. Part A: Polym.

Chem., 38, 1580 (2000).

48. M.M. Werlang, I.V.P. Yoshida, and M.A. de Araujo,

J. Inorg. Organomet. Polym., 5, 75 (1995).

49. P.R. Dvornic and R.W. Lenz, ‘‘Polysiloxanes,’’ in HighTemperature Siloxane Elastomers, P.R. Dvornic and R.W.

Lenz, Eds., Huthig & Wepf, Basel, Switzerland, Chapter 2

(1990).

50. C. Moysan, R. Riedel, R. Harshe, T. Rouxel, and F. Auger-

eau, J. Eur. Ceram. Soc., 27, 397 (2007).

51. M. Scheffler, R. Bordia, N. Travitzky, and P. Greil, J. Eur.Ceram. Soc., 25, 175 (2005).

52. T. Takahashi, H. Munstedt, M. Modesti, and P. Colombo, J.Eur. Ceram. Soc., 21, 2821 (2001).

53. F.I. Hurwitz, P. Heimann, S.C. Farmer, and D.M. Hembree

Jr., J. Mater. Sci., 28, 6622 (1993).

54. A. White, S.M. Oleff, R.D. Boyer, P.A. Budinger, and J.R.

Fox, Adv. Ceram. Mater., 2, 45 (1987).

55. J. Macan, I. Brnardic, S. Orlic, H. Ivankovic, and M. Ivan-

kovic, Polym. Degrad. Stab., 91, 122 (2006).

56. E. Amerio, M. Sangermano, G. Malucelli, A. Priola, and B.

Voit, Polymer, 46, 11241 (2005).

57. S.R. Kumar and M. Alagar, High Perform. Polym., 19, 3 (2007).

148 POLYMER ENGINEERING AND SCIENCE—-2008 DOI 10.1002/pen