Sei sulla pagina 1di 10

Journal of Materials Research and Technology

w w w. j m r t . c o m . b r

REVIEW ARTICLE

Thermogravimetric Stability of Polymer Composites Reinforced with Less Common Lignocellulosic Fibers an Overview
Sergio Neves Monteiro1,*, Vernica Calado2, Rubn J. S. Rodriguez3, Frederico M. Margem3
Instituto Militar de Engenharia (IME), Materials Science Department, Rio de Janeiro, Brazil. Escola de Qumica, Universidade Federal do Rio de Janeiro, Rio de Janeiro, Brazil. 3 Universidade Estadual do Norte Fluminense, Campos dos Goytacazes, Brazil.
1 2

Manuscript received April 14, 2012; in revised form June 19, 2012

Sergio Neves Monteiro (June 24, 1943) graduated as a Metallurgical Engineer (1966) at the Universidade Federal do Rio de Janeiro (UFRJ). Received his MSc (1967) and PhD (1972) from the University of Florida, followed by a course (1975) in Energy at the Escola Superior de Guerra (Brazilian War College) and Post-doctorate (1976) at the University of Stuttgart. Joined the Metallurgy Department (1968) and was appointed (1977) Full Professor of the Post-Graduation Program in Engineering (COPPE) of the UFRJ. He was elected Head of the Department (1978), Coordinator of COPPE (1982) and Under-Rector for Research (1983). Invited as Under-Secretary of Science for the State of Rio de Janeiro (1985) and Under-Secretary of College Education for the Federal Government (1989). Retired in 1993 and joined the Universidade Estadual do Norte Fluminense (UENF). Published over 900 articles in journals and conference proceedings. Has been honored with several awards including the ASM Fellowship. Is presently top researcher (1A) of the Conselho Nacional de Desenvolvimento Cientco e Tecnolgico (Brazilian Council for Scientic and Technological Development CNPq), vice-president of the Fundao de Amparo Pesquisa do Estado do Rio de Janeiro (Superior Council of the State of Rio de Janeiro Research Foundation FAPERJ), consultant for the main Brazilian R&D agencies and member of the Editorial board of three international journals.

Environmental, economic, and technical reasons justify research efforts aiming to provide natural materials with possibility of replacing synthetic ber composites. Commonly known lignocellulosic bers, such as jute, sisal, ax, hemp, coir, cotton, wood, and bamboo have not only been investigated as reinforcement of polymeric matrices but already applied in automobile components. Less common bers, such as curaua, henequen, que, buriti, olive husk, and kapok are recently being studied as potential reinforcement owing to their reasonable mechanical properties. The relatively low thermal stability of these bers could be a limitation to their composites. The works that have been dedicated to analyze the thermogravimetric stability of polymer composites reinforced with less common lignocellulosic bers were overviewed.
*Corresponding author.

E-mail address: sergio.neves@ig.com.br (S. N. Monteiro) 2012 Brazilian Metallurgical, Materials and Mining Association. Published by Elsevier Editora Ltda. All rights reserved.

J. Mater. Res. Tecnol. 2012; 1(2):117-126

118
KEY WORDS: Natural bers; Polymer-matrix composites (PMCs); Thermogravimetric analysis.

Monteiro et al.

2012 Brazilian Metallurgical, Materials and Mining Association. Published by Elsevier Editora Ltda. All rights reserved.

1. Introduction
The rst engineered composites materials were probably reinforced with natural bers and produced in the primordium of mankind. Evidence of adobe construction blocks of sun-dried clayey mud incorporated with straw was found in the ancient Egyptian civilization. According to our current interpretation, these rst composites would be considered environmentally friendly owing to their all natural components. Even today, the rudimentary technique used by our ancestors to fabricate such composites persists in many low income regions of the world. As mentioned by Bledzki and Gassan[1], in a seminal review on natural ber composites, only at the end of the 19th century the industrial manufacture of natural bers with small content of polymeric binder was reported for airplane seats and fuel tanks. Large amounts of sheets, tubes, and pipes for electronic purpose were continuously fabricated as earlier as 1908, using natural bers, paper, or cotton incorporated into phenolic matrix[1,2]. With the rising performance of petroleum-based polymers and synthetic bers as well as the facility to combine them into more uniform and stronger composites, the industrial production of traditional natural ber composites declined. In another classic review on biobres and biocomposites, Mohanty et al.[2] indicated that the synthetic ber composites reached commodity status in the 40s with glass ber reinforcing (berglass) unsaturated polyesters. These synthetic ber composites experienced an exponential growth after the World War II to become the most successful class of engineering materials[35] with application in practically all elds of human interest, from appliances and sports to surgical prosthesis and aerospace components. At the end of the last century, increasing environmental concerning over generalized pollution caused by non-degradable materials, especially long lasting plastics, and climate changes resulting from CO2 emission, promoted a growing tendency towards the substitution of synthetic bers composites. Natural materials, particularly cellulose-rich bers, also known as lignocellulosic bers, were renewed as reinforcement of polymer composites. A considerable number of researches have, in the past few decades, been dedicated to lignocellulosic bers as engineering materials and their reinforced polymer composites for applications in substitution of synthetic ber composites. Several reviews and general articles covered this trend[1,2,615]. In these publications, advantages are emphasized and drawbacks discussed aiming at reduce the limitation, which exists in practical use. Despite the drawbacks, a growing industrial application of lignocellulosic ber composites is nowadays occurring in sectors such as building construction, packaging, sport devices, electrical parts, and vehicle components[14]. Automotive industries, initially the Europeans followed by Americans and Japanese, are adopting this type of composites in several interior and exterior parts[1620].

It was emphasized[21,22] that, as compared to berglass, the lignocellulosic ber composites are lighter, cheaper, and less abrasive in contact with processing equipments. Furthermore, berglass represents a problem to the environment with restriction to nal destination by incineration in thermo-electric plants. Glass ber particulates also represent potential health hazardous both in the initial composite processing and latter at the end-of-life degradation. By contrast, the thermal stability of a lignocellulosic ber composite is inferior to similar matrix composite reinforced with glass ber. This could be a critical restriction for conditions associated with relatively high temperatures attained during the curing of the composite or its in-service use. In fact, temperature usually causes an initial degradation of the ber organic structure and thus limits the polymer composite application.

2. Thermal Decomposition of Lignocellulosic Material


Several review articles[2326] have, since more than ve decades ago, been dedicated to the thermal decomposition of lignocellulosic materials. Beall and Eickner[23] reviewed works on thermal analysis results of wood and its cellulose, hemicelluloses, and lignin constituents. Kilzer[24] reviewed works on the thermal decomposition of cellulose. Nguyen et al.[25,26] conducted subsequent reviews on the application of differential thermal analysis (DTA), differential scanning calorimetry (DSC), and thermogravimetry (TG) to the study of lignocellulosic materials as well as modied forms of lignocelluloses. These review articles[2326] indicated the following results:

2.1 Cellulose
Cellulose decomposition takes place by major reaction involving depolymerization, thermoxidation, dehydration, and formation of glycosans, depending on the presence of oxygen or an inert atmosphere. In a non-oxidative atmosphere, dehydration occurs in a range of 210C260C and depolymerization with volatilization of levoglucosans, at about 310C. Under oxygen, thermoxidative reactions occur in the temperature range of 160C250C. The degradation of cellulose by pyrolysis has been assumed to follow a rst-order kinetic. In the specic case of wood, cellulose decomposition in air begins at 320C with a maximum weight loss at 350C. In helium it also begins at 320C, but the maximum rate is shifted to 375C. Cotton cellulose displays two endothermic DTA peaks at 100C and 367C.

2.2 Hemicellulose
Hemicellulose constituents decompose at temperatures as low as 159C175C, as in the case of acetyl galactoglucomannan. DTA of xylan in oxygen exhibits a rst exothermic
J. Mater. Res. Tecnol. 2012; 1(2):117-126

Thermogravimetry of Composites with Less Common Natural Fibers peak at 215C. Both arabinogalactan and deacetylated galactoglucomannan begin degradation at 195C, reaching a maximum at 280C.

119

2.3 Lignin
Obtained by hydrolysis of wood, it was found to decompose in three stages. At temperatures below 220C250C, condensation and splitting of the side chains takes place. Between 300C400C, active pyrolysis leads to the formation of free radicals. Above 400C, decomposition is associated with a series of degradation and condensation reactions with accumulation of aromatic products.

2.4 Wood
In its natural and unmodied form, showed, under helium, a rst endothermic peak at 107C, corresponding to water loss. Weak endothermic peaks were found at 207C and 330C, corresponding to dehydration and depolymerization of various constituents, probably including hemicellulose. A very strong endothermic peak at 367C was attributed to cellulose decomposition. An exothermic peak rises above 400C, presumably related to recombination of cellulose and lignin fragments. Under oxygen, it was reported an endothermic peak at 87C, owing to evaporation of water, and two exothermic peaks at 343C and 470C, probably due to cellulose and lignin.

3. Thermal Analysis of Common Lignocellulosic Fiber Composites


A short review on the thermal stability of polymer composites reinforced with few common lignocellulosic bers was, for the rst time, presented as one of the sections of the Nabi Saheb and Jog[6] review on natural ber polymer composites. They indicated that the thermal degradation of natural bers is a crucial aspect in the development of their composites and thus has a bearing on the curing temperature in the case of thermosets and extrusion temperature in thermoplastic matrix composites. Nabi Sahed and Jog[6] also stated that thermal stability improvement have been attempted by coating and/or grafting the bers with monomers, quoting the works of Mohanty et al.[27] and Sabaa[28]. The effect of the composite fabrication ambient was discussed by the authors[6] as a possibility of lignocellulosic ber degradation, quoting the work of Sridhar et al.[29], and indicating that the actual practice is carried out under air and that thermal degradation can lead to inferior mechanical properties. As a nal remark, Nabi Saheb and Jog[6] concluded that the thermal degradation of the lignocellulosic ber inside the polymeric composite matrix also results in production of volatiles at processing temperatures above 200C. This could result in porous composites with lower densities and inferior mechanical properties. Since this rst short review[6], numerous works have investigated the thermal stability of polymer composites reinforced with common lignocellulosic bers. The reader may nd specic data and conclusions on the following articles listed by employed reinforced ber: a) Jute Fiber Composites, in phenol formaldehyde[30], polyester[31], vinyl ester[32], polyester with acrylic acid[33], high
J. Mater. Res. Tecnol. 2012; 1(2):117-126

density polyethylene[34], polypropylene[35], and polylactic acid[36] matrices; b) Hemp Fiber Composites, in polyester[37], cashew nut shell resin[38], epoxy[39], polypropylene[40,41], and starch-base thermoplastic[42] matrices; c) Sisal Fiber Composites, in polypropylene[43,44], blend of polypropylene with high density polyethylene[43], polystyrene[45], polypropylene and maleic acid anhydride grafted styrene-ethylene-co-butylene-styrene copolymer[46,47], epoxy[48], phenolic and lignophenolic[49], and soy protein blended with gelatin[50] matrices; d) Flax Fiber Composites, in polypropylene[51], epoxy[39], and polylactic acid[52] matrices; e) Coir Fiber Composites, in copolymer of starch with ethylene vinyl alcohol[53], and polyester[54] matrices; f) Cotton Fiber Composites, in phenolic thermoset[55] matrix; g) Kenaf Fiber Composites, in epoxy[39] chitosan[56,57], and starch-based thermoplastic[40] matrices; h) Wood Fiber Composites, in polypropylene[5860], low and high density polyethylene[59,61], and polyhydroxy(butyrateco-valerate)[60] matrices; i) Pineapple Fiber Composites, in polyethylene[62], polyhydroxy(butyrate-co-valerate)[63], and polycarbonate[64] matrices; j) Bamboo Fiber Composites, in polylactic acid[65], polybutylene succinate[65], epoxy[66], and polyhydroxy(butyrate-covalerate)[67] matrices; k) Ramie Fiber Composites, in polylactic acid[36] matrix; l) Banana Fiber Composites, in polyvinyl chloride[68] matrix; m) Bagasse Fiber Composites, in polyurethane (PU)[69], and recycled high density polyethylene[61] matrices. As a general comment, one may conclude from all these works on the thermal stability of polymer matrix composites reinforced with commonly known and used lignocellulosic bers that their processing and applications are restricted to a safe temperature of 250C, or to a maximum of 367C, in case of more stable specic polypropylene and maleic anhydride grafted styrene-ethylene-co-butylene-styrene copolymer matrix reinforced with sisal ber[46]. The reader should also bear in mind that all these composites, owing to the contribution of the lignocellulosic ber, display a DTG water loss peak which is found at temperatures as low as 37C[32] and inferred at about 140C144C[48,53].

4. Thermogravimetric Stability of Less Common Lignocellulosic Fiber Composites


In principle, a complete assessment of the thermal behavior of a material would require not only temperature difference, DTA, thermogravimetric, TG/DTG, differential calorimetric, DSC, and dynamic-mechanical (DMA) thermo analyses, but also the evaluation of properties such as thermal conductivity, specic heat, and thermal diffusivity. The scope of this overview is limited to thermal stability results associated with weight loss variation with temperature obtained by thermogravimetric analysis. These results, displayed as TG thermograms as well as its derivative DTG, will be covered for relevant works on polymer composites reinforced with less known bers published in internationally recognized sources. Numerous less common lignocellulosic bers such as curaua, henequen, que, buriti, olive

120 husk, kapok, abaca, isora, piassava, artichoke, milkweed, sabei, okra, caroa, sansevieria, palmyrah, maize, phragmite, esparto, malva, paina, sponge gourd, communis, pita oja, roselle, canary, rye, mesta, barley, raphia, oat, and rape are commercially available and currently being investigated for their potential as engineering materials[14,15,70]. Only a few of these less common bers, however, have been studied for thermogravimetric stability in association with polymer composites. These few bers curaua, henequen, que, buriti, olive husk, and kapok serve as sub-titles in reviewing the following related thermogravimetric stability works on their polymer composites.

Monteiro et al. shoulder peak around 370C, which is more pronounced for higher volume fractions of curaua ber, as well as a major peak almost coincident with that of the neat polyester. Fig. 1 illustrates the TG/DTG curves of neat polyester and composites reinforced with up to 30 vol.% of continuous and aligned curaua bers, adapted from the Ferreira et al.[74] work. In this gure, an enlarged insert reveals shoulder peaks associated with curaua ber thermal degradation in the composite. By considering these overviewed works, and based on the results for the isolated curaua bers[71,72,7577] with shoulder at 268C290C and major peak at 310C365C, one should conclude that PU[71], HDPE[72], and polyester matrices[74] composites are thermally more stable than the pure curaua ber.

4.1 Curaua Fiber Composites


Moth and Arajo[71] showed TG/DTG and DTA curves for PU matrix composites reinforced with up to 20 wt.% of curaua bers. For the neat PU, the DTG curve displays a small initial peak at 260C that was assigned to the decomposition of additives in the PU. A broad shoulder peak around 360C followed by a major peak at 422C were attributed to decomposition of the rigid and soft urethane bonding, respectively. The DTA curve shows three endothermic events at 250C, 330C, and 420C that corroborate those found in the DTG peaks. The investigated composites[71] also display faint DTG peaks around 60C, due to the release of water. Two other decomposition peaks with temperatures (wt.% of curaua bers) can be seen at 358C and 418C (5 wt.%); 356C and 420C (10 wt.%); and 356C and 418C (20 wt.%). Endothermic DTA peaks are also seen at the same temperatures and conditions. These results indicate that the thermal stability of the composites is practically the same of the PU. Moreover, the amount of curaua ber in the composite causes no apparent change in the thermal stability. Arajo et al.[72] presented results from TG/DTG curves of isolated curaua ber as well as high density polyethylene (HDPE) and HDPE matrix composites reinforced with 20 wt.% of either unmodied (natural) or poly(ethyleneco-vinyl-acetate) (EVA) and maleic anhydride grafted polyethylene (PE-g-MA) compatibilized curaua bers. The DTG curve of the neat HDPE shows a peak at 478C. The DTG curves of the composites display two distinct peaks. The rst, coinciding for all composites at 349C, occurred closer to the main decomposition peak of the curaua ber, 363C. Additionally, the authors calculated a weighted mean expected DTG curve if there was no interaction among the degradative process for the composites, quoting the work of Waldman and de Paoli[73]. The second peak was found at about 468C471C for the composites. A comparison of TG curves with the calculated curve allowed the authors[72] to indicate that the composite compatibilized with PE-g-MA is less stable, while those with EVA as well as with no compatibilization are more stable. Ferreira et al.[74] showed TG/DTG curve, obtained at a heating rate of 10C/min in nitrogen, for polyester composites reinforced with up to 30 vol.% of untreated curaua bers. A small initial peak observed for all composites around 70C was attributed to moisture release. The neat polyester begins to decompose around 250C and displays a major DTG peak at about 410C related to the degradation of its molecular chains. The composites display a

4.2 Henequen Fiber Composites


Sgriccia and Hawley[39] presented thermogravimetric results on epoxy matrix composites reinforced with 15 wt.% of hemp (section 3, item b), ax (section 3, item d), kenaf (section 3, item g), and henequen bers. These composites were both oven and microwave cured. For comparison, similar glass ber composites were also investigated. Although no TG curves were shown, the authors[39]
(a)

(b)

Fig. 1 TG/DTG curves of polyester composites reinforced with different volume fraction of curaua bers. (a) Plain curves; (b) enlarged DTG curves. Adapted from Ferreira et al.[74].

J. Mater. Res. Tecnol. 2012; 1(2):117-126

Thermogravimetry of Composites with Less Common Natural Fibers presented the degradation temperature associated with 5 wt.% to 75 wt.% of weight loss. Tables 1 and 2 reproduces these results and the reader can observe that the isolated henequen ber shows degradation temperature (weight loss) of 61C (5 wt.%) up to 392C (75 wt.%). On the other hand, the microwave cured composites show a signicant increase: 355C (5 wt.%) up to 483C (75 wt.%). These values are closer (Tables 1 and 2) to those corresponding to the neat epoxy microwaved of 403C (5 wt.%) and 484C (75 wt.%). Fig. 2, adapted from Sgriccia and Hawley[39] with necessary corrections, illustrated ESEM images of microwave cured composites. The authors[39] indicated that the natural ber composites investigated, including the henequen/epoxy, both oven and microwave cured, composites would not be suitable for applications at temperatures as high as those of similar glass ber composites.

121

4.3 Fique Fiber Composites


Gan and Mondragon[78] performed thermogravimetric analysis on both polypropylene (PP) and polyoxymethylene (POM) matrices composites reinforced with 20 wt.% que bers, untreated as well as modied with maleic anhydride (MA), propionic acid (PA), glycidyl-methacrylate (G) and formaldehyde (F) or compatibilized with a copolymer of polypropylene (Fluka) and maleic anhydride (MAPP). The discussed composite results were limited to some of the ber treatments. For instance, a major decomposition peak was reported for the neat PP at 473C, while the untreated as well as PA and MAP treated ber composites display three peaks. Fig. 3, adapted from the work of Gan and Mondragon[78], shows TG/DTG curves for neat PP and PP composites reinforced with 20 wt.% of que bers as well as neat POM and POM composites reinforced with 20 wt.% of que bers. The authors[78] indicated that the rst (310C) and second (387C) peaks for the untreated ber composites, as well as the corresponding rst (344C) and second (393C) peaks for the MAPP treated ber composites, are related to those of the isolated que ber, 301C and 356C, respectively, quoting their previous work[79]. The third peak at 451C for untreated and at 476C for MAPP treated ber composites correspond to the decomposition of the PP matrix. As a general comment, Gaan and Mondragon[78] stated that the lower thermal stability of POM with respect to PP does not allow for separating the contributions for thermal degradation of que bers and POM matrix in related composites. Higher temperatures than that for the neat POM are necessary for complete degradation of its composites, possibly due to ber degradation delaying and matrix crystallinity variations.

Table 1 Degradation temperatures for glass and henequen bers associated with different levels of TG weight loss T(C) 5 Weight loss (%) wt.% Henequen Glass 61.1 586.4 T(C) 25 wt.% 301.4
[39]

T(C) 50 wt.% 337.4

T(C) 75 wt.% 391.6

Adapted from Sgriccia and Hawley

Table 2 Degradation temperatures for glass and henequen ber reinforced epoxy matrix composites associated with different levels of TG weight loss T(C) 5 wt.% 400.3 400.4 305.8 355.0 396.6 403.3 T(C) 25 wt.% N/A N/A 374.6 389.1 N/A N/A T(C) 50 wt.% 401.8 401.7 391.0 392.0 397.9 N/A T(C) 75 wt.% 538.6 481.7 455.9 483.0 489.8 484.3

4.4 Buriti Fiber Composites


Santos et al.[80] conducted thermogravimetric experiments on cardanol-formaldehyde (CFR-thermoset resin) matrix composite incorporated with 5 wt.%, 10 wt.%, and 15 wt.% of buriti bers obtained from leaf straw. Buriti bers were both untreated and subjected to NaOH alkali (mercerization) or silanization treatments. TG curves show that the
(a) (c)

Weight loss (%) Glass oven Glass microwave Henequen oven Henequen microwave Neat epoxy oven Neat epoxy microwave

Adapted from Sgriccia and Hawley[39].

(b)

(d)

Fig. 2 ESEM micrographs of microwave cured epoxy composites reinforced with: (a) henequen and (b) glass bers. Adapted from Sgriccia and Hawley[39].

Fig. 3 TG/DTG curves: (a) neat PP; (b) PP composites with 20wt.% of que bers; (c) neat POM; (d) POM composites with 20 wt.% of que bers. Adapted from Gan and Mondragon[78].

J. Mater. Res. Tecnol. 2012; 1(2):117-126

122 CFR has, apparently, an onset temperature for thermal degradation, around 320C, which is higher than those for the 10 wt.% buriti ber composites, around 240C. In fact, the authors[80] indicated that both composites present an intermediate thermal stability in relation to the isolated buriti ber and the matrix. Furthermore, the mercerized ber composite curve shows an inection at about 305C because of the thermal ber degradation and another at 400C owing to the thermal degradation of the thermoset resin. Fig. 4, reproduced from the work of Santos et al.[80], shows SEM micrographs of untreated, mercerized and silanized buriti bers as well as 10 wt.% buriti ber reinforced CFR matrix composites. As indicated by the authors[80], the mercerized ber (Fig. 4a) displays a rough aspect, probably due to the removal of low molar mass compounds, which leaves cavities at the surface. The silanized ber (Fig. 4c) presents a more regular surface. This suggests that the silanization produces a lm coating on the entire ber surface. The composite (Fig. 4d) displays a lamellar-like morphology with evidence of excellent adhesion between the matrix and mercerized ber. This good coverage of the ber by the resin may contribute to improve its thermal stability.

Monteiro et al. maleic-anhydride-polypropylene (PPMA). The pure untreated OHF begins to degrade at 210C, while the VTAS treated OHF at 201C. This reduction was attributed by the authors[81] to the elimination of hydrogen bonds that requires signicant energy. DTG peaks were also related to the OHF degradation. The rst, around 100C, was assigned to water evaporation; the second, at 260C, for untreated and 250C for treated OHF were ascribed to both hemicellulose and glycidic bonds of cellulose decomposition. The third peak, at 325C, was attributed to cellulose decomposition, while the fourth peak, at 350C, to the lignin decomposition. The authors[81] indicated that these results are in agreement with those of Mder et al.[35] and Pracella et al.[40]. As for the neat matrix, a single DTG peak was observed at 397C in association with 97% of weight loss. The authors[81] also concluded that the grafting reaction of MA onto the PP, for the OHF treatment, generates a 16C reduction in thermal stability. TG/DTG curves for the composites showed an intermediate behavior between the pure OHF and the neat PP. In fact, according to Amar et al.[81], the thermal degradation of the olive husk ber composites occurred in a three step degradation process. However, the 20 wt.% OHF composites apparently display four DTG peaks. A small rst peak at about 100C, not mentioned by authors[81], is probably due to water release from the OHF. A second, in the temperature range of 232C308C, and a third at 308C350C, shoulder peaks were attributed to the decomposition of hemicellulose and cellulose, respectively. The main DTG peaks around 400C, related by the authors to the third decomposition stage between 350C 454C, were assigned to the PP matrix and OHF lignin joint degradation. It was also indicated by the authors that the composite modied by PPMA revealed a better thermal stability than those with untreated or VTAS treated OHF.

4.5 Olive Husk Fiber Composites


Amar et al.[81] performed thermogravimetric analysis at a heating rate of 10C/min in nitrogen on polypropylene (PP) matrix composites added with 10 wt.% and 20 wt.% of olive husk our (OHF), the solid portion remaining after pressing olives, which contains signicant amounts of lignocellulosic ber. The OHF were both untreated and subjected to vinyltriacetoxysilane (VTAS) chemical treatment or grafted with

4.6 Kapok/ Cotton Fibers Hybrid Composites


Mwaikambo et al.[82] employed hybrid kapok/cotton bers wove as a fabric, which was both untreated and chemically treated by either one hour soaking in acetic anhydride at 70C (acetylation) or dipped in 2% NaOH solution for 48 hours (alkali/mercerization), reinforcing both conventional isotactic polypropylene (iPP) and anhydride grafted polypropylene resins (MAiPP) matrices composites. In addition, the authors[82] also investigated the effect of accelerated weathering of the composites by immersion in boiling water for two hours before conducting thermogravimetric analysis at a heating rate of 10C/ min in nitrogen. Table 3, reproduced from their work[82], summarizes the main parameters obtained from TG curves for all distinct composites. Although not mentioned in the abstract, the authors[82] indicated in the experimental procedure that the plain weave kapok/cotton fabric was used together with nonwoven glass mat. Along the article, no comment or discussion was given regarding the role of the glass mat on the thermogravimetric analysis. As shown in Table 3, all composites display a dehydration temperature in the range of 59C78C, which is probably related to the water release from the kapok and cotton bers, as usually reported in other lignocellulosic ber composites[3032,50,53,54,56,71,74]. As major thermal stability results, Mwaikambo et al.[82], emJ. Mater. Res. Tecnol. 2012; 1(2):117-126

Fig. 4 SEM micrographs: (a) untreated; (b) mercerized; and (c) silanized buriti bers as well as (d) composite with 10 wt.% mercerized bers, indicated by white arrows. Inserts with high magnication details. Reproduced from Santos et al.[80].

Thermogravimetry of Composites with Less Common Natural Fibers


Table 3 Thermogravimetric results of polypropylene reinforced with kapok/cotton for different alkali- or acetylated-treated fabrics conventional or maleic anhydride grafted matrices and unweathered and weathered composites. Alkali-treated kapok/ cotton-iPP composite 74.1 332.3 372.8 656.7 Unweathered kapok/ cotton-MAiPP composite 77.5 339.5 381.9 459.1 657.9 Acetylated kapok/ cotton-iPP composite 59.3 339.5 372.63 431.7 655.3 Weathered kapok/ cotton-iPP composite 71.8 318.5 380.6 423.7 658.7 Unweathered kapok/ cotton-iPP composite 76.3 316.4 370.4 428.2 656.4 Weathered kapok/ cotton-MAiPP composite 71.8 339.2 384.2 462.5 655

123

Properties Dehydration temperature (oC) Onset temperature (oC) Degratation temperature 1 (oC) Degratation temperature 2 (oC) Decomposition temperature ( C)
o

Properties Dehydration temperature (oC) Onset temperature ( C)


o

Degratation temperature 1 (oC) Degratation temperature 2 (oC) Degratation temperature 2 (oC) Reproduced from Amar B et al.[81].

phasized that the weathered kapok/cotton-iPP composite shows the highest degradation temperature followed by kapok/cotton-MAiPP composite. As for the fabric treatment, the acetylated kapok/cotton fabric-iPP composites display a higher onset temperature of degradation than both the weathered composite and mercerized kapok/ cotton fabric-iPP composites, but lower than MAiPP composites. The authors[82] stated that the maleic anhydride is likely to be the possible source for the improved thermal properties of PP as also evidenced by the rise in the degradation temperature of the composites. By contrast, mercerization of bers, which is known to result in reduced crystalline cellulose, has been found to decrease the thermal stability of the kapok/cotton fabric-iPP composites. As a general conclusion, it has also been found, with exception of mercerization kapok/cotton fabric-iPP (Table 3), that all composites exhibit two degradation temperatures. Mwaikambo et al.[82] suggested that the rst degradation temperature is a result of depolymerization of the cellulose materials. The second degradation temperature was caused by the polymeric matrices breakdown into monomers and/or decomposition of the levoglucosan.

5. Concluding Remarks
A few points are worth being discussed concerning the results presented in this overview. First, the reader should noticed the relatively small number of papers[39,71,72,74,78,8082] covering the thermogravimetric stability of polymer composites reinforced with less common lignocellulosic bers. Indeed, it was rather surprising that only such limited works have so far been dedicated to these composites in spite of the potential presented by less common lignocellulosic bers. For instance, Fiore et al.[83] reported tensile strength above 300 MPa and elastic modulus above 1.5 GPa for artichoke bers, while De Rosa et al.[84] found
J. Mater. Res. Tecnol. 2012; 1(2):117-126

tensile strength above 800 MPa and elastic modulus above 4 GPa for okra bers. These less common lignocellulosic bers certainly possess an engineering potential, of both strength and stiffness, for polymer composite reinforcement. In addition to future specic works on the mechanical properties of these less common natural ber composites, the thermal stability of such composites will also need to be investigated for practical use. A second point of the present overview is the alert on the growing demand for studies covering the several less common lignocellulosic bers with potential application as engineering materials. Consequently, the general properties and, in particular, the TG/DTG analysis of novel composites based on these bers have to be assessed. Despite the limited and fragmented information existing in the literature, this overview conveys relevant conclusion regarding the thermogravimetric stability of polymer composites reinforced with less common lignocellulosic bers. Apparently, an initial weight loss associated with a water release DTG peak below 200C is a common feature of these composites[39,71,74,81,82], although not emphasized in some works[73,78,80]. This initial peak is most probably a result of the evaporation of water from the ber surface, since the polymeric matrix contribution, if existing, should be relatively small. For practical use, however, the temperature related to the onset of thermal degradation can be considered the composite thermal stability limit. In the overviewed works, this limit was found to be in the range of 240C[80]355C[39], and attributed to the lignocellulosic ber decomposition. According to Santos et al.[80], these onset degradation temperatures are intermediate between the isolated ber and the polymer matrix. Other higher temperature DTG peaks, 422C[71]463C[81], are related to the polymer matrix macromolecular degradation or depolymerization but not as important as the onset degradation temperature to dene the composite thermogravimetric stability.

124

Monteiro et al.
23. Beall FC, Eickner HW. Thermal degradation of wood components: a review of literature. USDA Forest Service Research Paper 1970; FPL 130. 24. Kilzer FJ. Cellulose and Cellulose Derivatives. New York: WileyInterscience, 1971, p. 101722. 25. Nguyen T, Zavarin E, Barral EM. Thermal analysis of lignocellulosic materials. Part I - Unmodied materials. J Macromol Sci Rev Macromol Chem 1981; C 20:165. 26. Nguyen T, Zavarin E, Barral EM. Thermal analysis of lignocellulosic materials. Part II - Modied materials. J Macromol Sci Rev Macromol Chem 1981; C 21:160. 27. Mohanty AK, Patnaik S, Singh BC, Misra M. Graft copolymerization of acrylonitrile onto acetylated jute bers. J Appl Polym Sci 1989; 37(5):117181. 28. Sabaa MW. Thermal degradation behavior of sisal bers grafted with various vinyl monomers. Polym Degrad Stab 1991; 32(2):20917. 29. Sridhar MK, Basavarajjappa G, Kasturi SS, Balsubramanian N. Thermal stability of jute bers. Indian J Tex Res 1982; 7:8791. 30. Das S, Saha AK, Chourdhury PH, Basak RK, Mitra BC, Todd T et al. Effect of steam pre-treatment of jute ber on dimensional stability of jute composite. J Appl Polym Sci 2000; 76:165261. 31. Dash BN, Rana AK, Mishra HK, Nayak SK, Tripathy SS. Novel lowcost jute-polyester composites. III. Weathering and thermal behavior. J Appl Polym Sci 2000; 78:16719. 32. Ray D, Sarkar BK, Basak RK, Rana AK. Thermal behavior of vinyl ester resin matrix composites reinforced with alkali-treated jute bers. Appl Polym Sci 2004; 94:1239. 33. Manfredi LB, Rodriguez ES, Wladyka-Przybylak M, Vazquez A. Thermal degradation and re resistance of unsaturated polyester, modied acrylic resins and their composites with natural bres. Polym Degrad Stab 2006; 91:25561. 34. Mohanty S, Verma SK, Nayak SK. Dyamic mechanical and thermal properties of MAPE treated jute/HDPE composites. Compos Sci Technol 2006; 66:53847. 35. Doan T-T-L, Brodowsky H, Mder E. Jute bre/polypropylene composites. II. Thermal, hydrothermal and dynamic mechanical behavior. Compos Sci Technol 2007; 67:270714. 36. Yu T, Li Y, Ren J. Preparation and properties of short natural ber reinforced poly(lactic acid) composites. Trans Nonferrous Met Soc China 2009; 19:s6515. 37. Aziz SH, Ansell MP. The effect of alkalinization and bre alignment on the mechanical and thermal properties of kenaf and hemp bast bre composites: Part 1 Polyester resin matrix. Compos Sci Technol 2004; 64:121930. 38. Aziz SH, Ansell MP. The effect of alkalinization and bre alignment on the mechanical and thermal properties of kenaf and hemp best bre composites: Part 2 Cashew nut shell liquid matrix. Compos Sci Technol 2004; 64:12318. 39. Sgriccia N, Hawley MC. Thermal, morphological, and electrical characterization of microwave processed natural ber composites. Compos Sci Technol 2007; 67:198691. 40. Pracella M, Chionna D, Anguillesi I, Kulinski Z, Piorkowska E. Functionalization, compatibilization and properties of polypropylene composites with hemp bre. Compos Sci Technol 2006; 66:221830. 41. Beckermann GW, Pickering KL. Engineering and evaluation of hemp bre reinforced polypropylene composites: bre treatment and matrix modication. Composites Part A 2008; 39:97988. 42. Moriana R, Vilaplana F, Karlsson S, Ribes-Greus A. Improved thermo-mechanical properties by the addition of natural bres in starch-based sustainable biocomposites. Composites Part A 2011; 42:3040. J. Mater. Res. Tecnol. 2012; 1(2):117-126

Acknowledgements
The authors thank the support to this investigation by the Brazilian agencies: CNPq, CAPES and FAPERJ.

References
1. Bledzki AK, Gassan J. Composites reinforced with cellulosebased bers. Prog Polym Sci 1999; 4:22174. 2. Mohanty AK, Misra M, Hinrichsen G. Biober, biodegradable polymers and biocomposites: an overview. Macromol Mat Eng 2000; 276/277:124. 3. Agarwal BD, Broutman LJ. Analysis and Performance of Fiber Composites. 2.ed. New York: Wiley, 1990. 4. Ashbee KH. Fundamental Principles of Fiber Reinforced Composites. 2.ed. Lancaster, PA: Technomic, 1993. 5. Chawla KK. Composite Materials Science and Engineering. 2. ed. New York: Springer-Verlag, 1998. 6. Nabi Sahed D, Jog JP. Natural ber polymer composites: a review. Adv Polym Technol 1999; 18(4):35163. 7. Eichhorn SJ, Baillie CA, Zafeiropoulos N, Mwaikambo LY, Ansell MP, Dufresne A et al. Review Current international research into cellulosic bres and composites. J Mat Sci 2001; 36:210731. 8. Mohanty AK, Misra M, Drzal LT. Sustainable bio-composites from renewable resources: opportunities and challenges in the green materials world. J Polym Environ 2002; 10:1926. 9. Netravali AN, Chabba S. Composites get greener. Mater Today 2003; 6:229. 10. Mohanty AK, Mishra M, Drzal LT. Natural Fibers, Biopolymers and Biocomposites. Boca Raton: CRC press, 2005. 11. Satyanarayana KG, Guimares JL, Wypych F. Studies on lignocellulosic bers of Brazil. Part I: Source, production, morphology, properties and applications. Compos Part A 2007; 38:1694709. 12. Crocker J. Natural materials innovative natural composites. Mater Technol 2008; 23:1748. 13. Monteiro SN, Lopes FPD, Ferreira AS, Nascimento DCO. Natural ber polymer matrix composites: cheaper, tougher and environmentally friendly. JOM 2009; 61:1722. 14. Kalia S, Kaith BS, Kaurs I (eds.). Cellulose Fibers: Bio- and Nano- Polymer Composites. New York: Springer, 2011. 15. Monteiro SN, Lopes FPD, Barbosa AP, Bevitori AB, Silva ILA, Costa LL. Natural lignocellulosic bers as engineering materials. Metal Mater Trans A 2011; 42:296374. 16. Larbig H, Scherzer H, Dahlke B, Poltrock R. Natural ber reinforced foams based on renewable resources for automotive interior applications. J Cell Plast 1998; 34:36179. 17. Marsh G. Next step for automotive materials. Mater Today 2003; 6:3643. 18. Holbery J, Houston D. Natural-ber-reinforced polymer composites in automotive applications. JOM 2006; 58:806. 19. Zah R, Hischier R, Leo AL, Brown I. Curaua bers in automobile industry A sustainability assessment. J Clean Prod 2007; 15:103240. 20. Alves C, Ferro PMC, Silva AJ, Reis LG, Freitas M, Rodrigues LB. Ecodesign of automotive components making use of natural jute ber composites. J Clean Prod 2010; 18(4):313-27. 21. Wambua P, Ivens I, Verpoest I. Natural bers: can they replace glass and bre reinforced plastics? Compos Sci Technol 2003; 63:125964. 22. Joshi SV, Drzal LT, Mohanty AK, Arora S. Are natural ber composites environmentally superior to glass ber reinforced composites? Compos Part A 2004; 35:3716.

Thermogravimetry of Composites with Less Common Natural Fibers


43. Albano C, Gonzalez J, Ichazo M, Kaiser D. Thermal stability of blends of polyolens and sisal ber. Polym Degrad Stab 1999; 66:17990. 44. Joseph PV, Joseph K, Thomas S, Pillai CKS, Prasad VS, Groeninckx G et al. The thermal and crystallization studies of short sisal bre reinforced polypropylene composites. Composites Part A 2003; 34:25366. 45. Nair KCM, Thomas S, Groeninckx G. Thermal and dynamic mechanical analysis of polystyrene composites reinforced with short sisal bres. Compos Sci Technol 2001; 61:251929. 46. Xie XL, Fung KL, Li RKY, Tjong SC, May Y-W. Structural and mechanical behavior of polypropylene maleated styrene(ethylene-co-butylene-styrene/sisal ber composites prepared by injection molding. J Polym Sci Part B: Polym Phys 2002; 40:121422. 47. Xie XL, Li RKY, Tjong SC, May Y-W. Structural properties and mechanical behavior of injection molded composites of polypropylene and sisal ber. Polym Compos 2002; 23(3):31928. 48. Gan P , Garbizu S, Llano-Ponte R, Mondragon I. Surface modication of sisal bers: effects on the mechanical and thermal properties of their epoxy composites. Polym Compos 2005; 26:1214. 49. Paiva JMF, Frollini E, Macromol J. Unmodied and modied surface sisal bers as reinforcement of phenolic and lignophenolic matrices composites: Thermal analyses of bers and composites. Mater Eng 2006; 291:40517. 50. Kim JT, Netravali AN. Mechanical and thermal properties of sisal ber-reinforced green composites with soy protein/gelatin resins. Biobased Mater Bioenergy 2010; 4(4):33845. 51. Arbelaiz A, Fernandez B, Ramos JA, Mondragon I. Thermal and crystallization studies of short ax bre reinforced polypropylene matrix composites: effect of treatments. Thermochim Acta 2006; 440(2):11121. 52. Aydin M, Tozlu H, Kemaloglu S, Aytac A, Ozkoc G. Effects of Alkali Treatment on the Properties of Short Flax FiberPoly(Lactic Acid) Eco-Composites. J Polym Environm 2011; 19(1):117. 53. Rosa MF, Chiou B-S, Medeiros ES, Woods DF, William TG, Mattoso LHC et al. Effect of ber treatments on tensile and thermal properties of starch/ethylene vinyl alcohol copolymers/coir biocomposites. Bioresource Technol 2009; 100(21):5196202. 54. Santaf Jr. HPG, Rodriguez RJS, Monteiro SN, Castillo TE. Characterization of thermogravimetric behavior of polyester composites reinforced with coir ber. In: Characterization of Minerals, Metals and Materials Symposium TMS 2011 Annual Conference, San Diego, CA, USA, March 2011, p. 16. 55. Silva CG, Benaducci DB, Frollini E. Lyocell and cotton bers as reinforcements for a thermoset polymer. BioResources 2012; 7(1):7898. 56. Julkapli NM, Akil HM. Thermal Properties of Kenaf-Filled Chitosan Biocomposites. Polym-Plast Technol Eng 2010; 49(2):14753. 57. Akil HM, Omar MF, Mazuki AAM, Saee S, Ishak ZAM, Bakar AA. Kenaf ber reinforced composites: A review. Mater Design 2011; 32(8-9):410721. 58. Coutinho FMB, Costa THS, Carvalho DL, Gorelova MM, de Santa Maria LC. Thermal behaviour of modied wood bers. Polym Test 1998; 17(5):299310. 59. Doh GH, Lee S-Y, Kang I-A, Kong Y-T. Thermal behavior of liqueed wood polymer composites (LWPC). Compos Struct 2005; 68(1):1038. 60. Bhardwaj R, Mohanty AK, Drzal LT, Pourboghrat F, Misra M. Renewable resource-based green composites from recycled cellulose ber and poly(3-hydroxybutyrate-co-3-hydroxyvalerate) bioplastic. Biomacromolecules 2006; 7(6):204451. J. Mater. Res. Tecnol. 2012; 1(2):117-126

125

61. Lei Y, Wu Q, Yao F, Xu Y. Preparation and properties of recycled HDPE/natural ber composites. Composites Part A 2007; 38(7):166474. 62. George J, Bhagawan SS, Thomas S. Thermogravimetric and dynamic mechanical thermal analysis of pineapple bre reinforced polyethylene composites. J Thermal Anal Cal 1996; 47(4):112140. 63. Luo S, Netravali AN. Mechanical and thermal properties of environment-friendly green composites made from pineapple leaf bers and poly (hydroxybutyrate-co-valerate) resin. Polym Compos 1999; 20(3):36778. 64. Threepopnatkul P , Kaerkitcha N, Athipongarporn N. Effect of surface treatment on performance of pineapple leaf berpolycarbonate composites. Composites Part B 2009; 40(7):62832. 65. Lee S-H, Wang S. Biodegradable polymers/bamboo ber biocomposite with bio-based coupling agent. Composites Part A 2006; 37(1):8091. 66. Shih Y-F. Mechanical and thermal properties of waste water bamboo husk ber reinforced epoxy composites. Mater Sci Eng A 2007; 4456:28995. 67. Singh S, Mohanty AK, Sugie T, Takai Y, Hamada H. Renewable resource based biocomposites from natural ber and polyhydroxybutyrate-co-valerate (PHBV) bioplastic. Composites Part A 2008; 39(5):87586. 68. Zainudin ES, Sapuan SM, Abdan K, Mohamad MTM. Thermal degradation of banana pseudo-stem lled unplasticized polyvinyl chloride (UPVC) composites. Materials and Design 2009; 30(3):55762. 69. Moth CG, de Arajo CR, de Oliveira MA, Yoshida MI. Thermal decomposition kinetics of polyurethane-composites with bagasse of sugar cane. J Therm Anal Cal 2002; 67(2):30512. 70. Jawaid M, Abdul Khalil HPS. Cellulosic/synthetic bre reinforced polymer hybrid composites: A review. Carbohydrate Polym 2011; 86(1):118. 71. Moth CG, de Araujo CR. Thermal and mechanical characterization of polyuretane composites with curaua bers (in Portuguese). Polmeros: Cincia e Tecnologia 2004; 14(4):27488. 72. Arajo JR, Waldman WR, de Paoli M-A. Thermal properties of high density polyethylene composites with natural bres: Coupling agent effect. Polym Degrad Stab 2008; 93(10):17705. 73. Waldman WR, de Paoli M-A. Thermo-mechanical degradation of polypropylene, low-density polyethylene and their 1:1 blend. Polym Degrad Stab 1998; 60(23):3018. 74. Ferreira AS, Rodriguez RJS, Lopes FPD, Monteiro SN. Thermal analysis of curaua ber reinforced polyester matrix composites. In: Characterization of Minerals, Metals and Materials Symposium TMS Conference (TMS 2010), Seattle, WA, USA, March 2010, p. 17. 75. Santos PA, Spinac MAS, Fermoselli KKG, de Paoli M-A. Polyamide-6/vegetal ber composite prepared by extrusion and injection molding. Composites Part A 2007; 38(12):240411. 76. Tomczak F, Satyanarayana KG, Sydenstricker THD. Studies on lignocellulosic bers of Brazil: Part III Morphology and properties of Brazilian curaua bers. Composites Part A 2007; 38(10):222736. 77. Spinac MAS, Lambert CS, Fermoselli KKG, de Paoli M-A. Characterization of lignocellulosic curaua bres. Carbohydrate Polym 2009; 77(1):4753. 78. Gan P, Mondragon I. Thermal and degradation behavior of que ber reinforced thermoplastic matrix composites. J Therm Anal Cal 2003; 73(3):78395. 79. Gan P, Mondragon I. Surface modication of que bers. Effect on their physico-mechanical properties. Polym Compos 2002; 23(3):38394.

126
80. Santos RS, Souza AA, de Paoli M-A, Souza CML. Cardanolformaldehyde thermoset composites reinforced with buriti bers: Preparation and characterization. Composites Part A 2010; 41(9):11239. 81. Amar B, Salem K, Hocine D, Chadia I, Juan MJ. Study and characterization of composite materials based on polypropylene load with olive husk our. J Appl Polym Sci 2011; 122: 138294.

Monteiro et al.
82. Mwaikambo LY, Martuscelli E, Avella M. Kapok/cotton fabric polypropylene composites. Polym Test 2000; 19(8):90518. 83. Fiore V, Valenza A, di Bella G. Artichoke (Cynara cardunculus L.) bres as potential reinforcement of composite structures. Compos Sci Technol 2011; 71(8):113844. 84. De Rosa IM, Kenny JM, Puglia D, Santulli C, Sarasimi F. Morphological, thermal and mechanical characterization of okra (Abelmoschus esculentus) bres as potential reinforcement in polymer composites. Compos Sci Technol 2010; 70(1):11622.

J. Mater. Res. Tecnol. 2012; 1(2):117-126

Potrebbero piacerti anche