07.04.2013 Views

30. Furan-Based Adhesives

30. Furan-Based Adhesives

30. Furan-Based Adhesives

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

30<br />

<strong>Furan</strong>-<strong>Based</strong> <strong>Adhesives</strong><br />

Mohamed Naceur Belgacem and Alessandro Gandini<br />

Ecole Française de Papeterie et des Industries Graphiques (INPG),<br />

St. Martin d’Hères, France<br />

I. INTRODUCTION<br />

The dwindling availability of fossil reserves constitutes a driving force towards finding<br />

alternative resources which can substitute them, totally or partially, in order to prepare<br />

chemicals and materials that are normally produced from petroleum and coal. In this<br />

context, vegetable biomass represents a very promising source since it offers a large variety<br />

of potential monomers, oligomers, and polymers, some of which can be extracted and used<br />

as such (namely, products such as terpenes, tannins, rosins, lignins, and cellulose) and<br />

others which can be suitably transformed to give monomers, solvents, surfactants, and a<br />

variety of polymeric materials (e.g., modified sugars, saponified oils, furfural and its<br />

derivatives, and cellulose acetates). We tried to show [1] that, besides its extensive use<br />

as a source of fibers for papermaking and textiles, vegetable biomass can also lead to<br />

interesting chemicals and materials. In a recent review [2], we focused on the use of furanic<br />

monomers for the preparation of polymeric materials and showed that different petroleum-based<br />

monomers (especially aromatic derivatives) could be replaced by their furanic<br />

counterparts. Thus, a variety of totally furanic, aromatic–furanic, and aliphatic–furanic<br />

polymers display properties similar to (and sometimes better than) those of currently used<br />

polymers derived from petroleum, proving that a whole area of biomass-based materials<br />

can be developed from two first-generation compounds which are readily available from a<br />

wide spectrum of renewable resources.<br />

<strong>Furan</strong>ic monomers can be obtained from hemicelluloses which are among the main<br />

constituents of vegetal biomass and are abundant in trees and agricultural residues of<br />

annual plants, such as sugarcane bagasse, oat hulls, corn husks, rice, and wheat straw.<br />

The precursors of most industrial furan derivatives are obtained directly from hemicelluloses<br />

through the acid-catalyzed hydrolysis of pentosans (e.g., xylans) followed by<br />

dehydration and cyclization of the ensuing pentoses leading to the formation of furfural<br />

(1), which is today the most important first-generation furan derivative, produced<br />

industrially at a rate of ca. 200,000 tonnes per year. This output is spread widely<br />

among numerous countries, including both industrialized and developing economies,<br />

because the process is particularly simple and the raw materials are available and plentiful<br />

virtually everywhere and are renewable often on short cycles. An additional advantage<br />

of this approach is that it calls upon a rational exploitation of agricultural wastes. Furfural<br />

Copyright © 2003 by Taylor & Francis Group, LLC


can be used as such, but is mostly (more than 80%) converted into furfuryl alcohol<br />

(2) using either liquid-phase or vapor-phase hydrogenation in the presence of copper<br />

catalysts which were found to be very selective in avoiding the hydrogenation of<br />

the heterocycle ring [3].<br />

Furfuryl alcohol finds numerous applications as monomer (see below) and has,<br />

therefore, been for decades the most important second-generation furan derivative.<br />

It is also used to prepare 2,5-bis(hydroxymethyl furan) (3) through its reaction with<br />

formaldehyde [3], namely:<br />

Compound 3 can also be prepared by the hydrogenation of 5-hydroxymethyl<br />

furfural (4) which, in turn, is obtained from hexoses following the same acid-catalyzed<br />

process described above for furfural [2].<br />

Compounds 1, 2, and 3 are among the most relevant monomers or co-monomers for<br />

furan-based adhesives, but so also are furfurylidene acetone (5) and its bis-adduct 6. The<br />

synthesis of 5 involves the base-catalyzed reaction between 1 and acetone [2] and, in the<br />

same context, the use of an excess of 1 leads to the formation of 6:<br />

This chapter is devoted to adhesives and resins prepared from totally furanic<br />

monomers or formulations in which furanic compounds are added. In this realm, only<br />

Copyright © 2003 by Taylor & Francis Group, LLC


a few furanic monomers and resins are involved, namely: 1, 2, 3, 5, and 6, as well as liquid<br />

oligomers of 2 (poly2) and3 (poly3). The properties of these monomers together with<br />

the mechanisms of their resinification and the composition of poly2 and poly3 will<br />

be briefly dealt with before discussing their use in the manufacture of resins for binders<br />

and adhesives.<br />

II. PROPERTIES OF FURANIC MONOMERS<br />

The relevant properties of furanic compounds covered in this review are summarized in<br />

Table 1.<br />

The compositions of poly2 and poly3 were studied by several groups [2,3] and shown<br />

to have mainly the following structures:<br />

Their relative abundance depends, of course, on the conditions used for their<br />

syntheses. A typical composition [3] is given in Table 2.<br />

Table 1 Properties of <strong>Furan</strong>ic Compounds Used in <strong>Adhesives</strong><br />

Compound type 1 2 3 5 6<br />

Molecular weight 96.09 98.10 128.10 136.15 214.22<br />

Boiling point ( C) 161 170 — 116 a<br />

—<br />

Melting point ( C) 39.7<br />

Density at 20 C (kg/dm 3 ) 1.16 1.13 — 1.06 b<br />

—<br />

Refractive index at 20 C 1.53 1.49 — —<br />

Viscosity at 25 C (mPa s) 1.48 4.62 — — —<br />

Surface tension (mN/m) 40 c<br />

38 d<br />

— — —<br />

a At 10 mm Hg.<br />

b at 45 C.<br />

c at 30 C.<br />

d at 25 C.<br />

Table 2 Typical Composition (w/w %) of poly2 and poly3<br />

7 8 9 10<br />

poly2 25 12 35 28<br />

poly3 — — 5 95<br />

Copyright © 2003 by Taylor & Francis Group, LLC


III. HISTORY, ADVANTAGES, AND LIMITATIONS ASSOCIATED WITH<br />

THE USE OF FURAN-BASED ADHESIVES<br />

The first synthetic thermosets used as adhesives were phenol–formaldehyde resins<br />

produced at the end of the nineteenth century, historically linked to Baekeland’s process<br />

which attained industrial status at the beginning of the twentieth century [4]. <strong>Furan</strong>ic<br />

condensates appeared much later as a result of the marketing of 2. They were first used<br />

as foundry binders by Quaker Oats in 1960. The use of furanic resins in the aerospace<br />

industry began ten years later. Although furanic resins represent a mere 1% of the total<br />

thermoset production, the high added-value of these materials amply justifies their use.<br />

In fact, furan-based adhesives and binders are fire-, solvent-, and acid- or alkali-resistant.<br />

They are known, however, to display two main drawbacks related to their sensitivity to<br />

shrinkage and oxidation.<br />

IV. RESINIFICATION MECHANISMS<br />

The acid- or heat-initiated cross-linking mechanisms of 1 were extensively studied for<br />

decades, but because of the complexity of the reactions involved and the effect of<br />

atmospheric conditions (e.g., light, oxygen, and water vapor) intermediate products<br />

were not identified until 1975. In that study, 1 was polymerized at 100–250 C in the<br />

absence of air and the following intermediates were isolated [5,6]:<br />

And for the final product, the following structure was proposed [5,6]:<br />

The polycondensation of 2 in acidic media has also been studied for a long time,<br />

but only recently was a clear-cut reaction mechanism established from a study in our<br />

Copyright © 2003 by Taylor & Francis Group, LLC


laboratory [2,7,8]. The success of this investigation stemmed from the fact that a large<br />

number of model compounds were synthesized which helped to establish the mechanisms<br />

of both cross-linking and color formation in this process. The use of mild catalysts<br />

confirmed that the first steps of the polymerization reactions occurred as follows:<br />

This initial mechanism does not explain, however, these anomalies since both<br />

macromolecular structures should give rise to colorless and thermoplastic materials.<br />

It was then shown that only several units actually condensed following this mechanism,<br />

since the average degree of polymerization (DP) never exceeded about 5, and crosslinking<br />

and color formation rapidly took place thereafter. In the mechanism of color<br />

formation, sketched in Scheme 1, we postulated that the formation of highly conjugated<br />

sequences resulted from successive hydride-ion/proton abstraction cycles [7]. This<br />

mechanism was confirmed by using different model compounds which were treated with<br />

an excess of hydride-ion (H ) abstractors (such as dioxolenium or triphenylmethyl<br />

cations) and the ensuing reactions followed by both ultraviolet (UV)–visible and<br />

1 H nuclear magnetic resonance (NMR) spectroscopies. This mechanism also explained<br />

the presence of methyl groups already observed by several authors [9–11]. The reaction<br />

of poly2 (obtained at early stages of the polycondensation) with hydride-ion abstractors<br />

was again followed by UV–visible spectroscopy and the results confirmed the proposed<br />

mechanism. Thus, the presence of conjugated sequences of different lengths was<br />

established, since the corresponding carbenium ions absorbed at different wavelengths,<br />

namely around 420, 450, 540, 600, and 800 nm.<br />

Having solved the long-standing puzzle related to color formation, we switched to the<br />

problem of the occurrence of branching and/or cross-linking reactions [2,7,8]. It was<br />

argued that these events could start either from the ‘‘irregular’’ units formed by<br />

the mechanism shown in Scheme 1, as illustrated in Scheme 2, and/or by Diels–Alder<br />

reactions between two chains, as proposed in Scheme 3. In fact, since the participation<br />

of furanic hydrogen atoms at C3 and C4 and those of methylene bridges had been clearly<br />

excluded on the basis of model reactions, it seemed reasonable to attribute the branching<br />

and cross-linking reactions to these two mechanisms. The second alternative, involving<br />

the cross-linking through Diels–Alder reactions, was recently confirmed by using<br />

2,5-dimethyl furan as a solvent for the acid-catalyzed polycondensation of 2. In this<br />

experiment, the large excess of dimethyl furan played the role of predominant diene<br />

trap for the exo-dihydrofuran dienophiles and thus prevented their coupling with<br />

the regular units of poly2 (Scheme 3). The fact that in these conditions the polymers<br />

remained soluble up to long reaction times and high yields was taken as clear evidence<br />

of the validity of Scheme 3.<br />

Copyright © 2003 by Taylor & Francis Group, LLC


Scheme 1<br />

Scheme 2<br />

Copyright © 2003 by Taylor & Francis Group, LLC


V. FURAN RESINS AS FOUNDRY BINDERS<br />

<strong>Furan</strong> resins have been extensively used as foundry binders in combination with<br />

formaldehyde, urea, phenol, and casein, for decades [12,13]. The main two monomers<br />

used in this field are 1 and 2. Table 3 summarizes their proportions in different commercial<br />

phenolic resins [12].<br />

The main advantages of furan resins are due to their excellent thermal stability, and<br />

remarkable resistance to acidic conditions, as well as to fire and corrosion. These resins<br />

Scheme 3<br />

Table 3 Proportions of 1 and 2 in Commercial Phenolic Resins<br />

Supplier<br />

Amount<br />

added<br />

(% w/w)<br />

1 2<br />

Amount<br />

retained after<br />

curing (% of the<br />

amount added)<br />

Amount<br />

added<br />

(% w/w)<br />

Amount<br />

retained after<br />

curing (% of the<br />

amount added)<br />

Bakelite 0215 Quaker Oats Co. 10 90 10 94<br />

Bakelite 0215 Quaker Oats Co. 20 87 20 86<br />

Bakelite 2417 Quaker Oats Co. 20 85 20 83<br />

Durez 7031 OxyChem 20 88 20 85<br />

Durez 8045 OxyChem — — 20 77<br />

Durez 14000 OxyChem — — 20 87<br />

Durite 278 Contenti Inc. 10 96 10 91<br />

Durite 278 Contenti Inc. 20 96 20 92<br />

Durite 3022 Contenti Inc. 10 95 10 93<br />

Durite 3022 Contenti Inc. 20 92 20 89<br />

Durite 1530 Contenti Inc. 20 93 20 92<br />

Monsanto 795 Monsanto 20 88 — —<br />

Varcum 1364 OxyChem 20 87 20 84<br />

Varcum 1192 OxyChem 20 69 20 80<br />

Copyright © 2003 by Taylor & Francis Group, LLC


have found widespread industrial applications as witnessed by the large number of<br />

both patents covering their uses and scientific publications dealing with their chemistry,<br />

structures, and properties [3,6,12,14–16].<br />

There are three techniques associated with their production, mostly covered by<br />

patent literature, namely: (i) no-bake, (ii) hot-box, and (iii) cold-box processes. The nobake<br />

technique is simple and relatively cheap. It consists in mixing the resin (based on 2)<br />

with the sand in the presence of an acidic catalyst. The reaction starts at room temperature<br />

and the curing is accelerated by the heat generated during the polycondensation reaction.<br />

The molds thus obtained are withdrawn after 10–30 min and left undisturbed for 3–6 h in<br />

order to accomplish a total curing. The hot-box technique is used in light (e.g., aluminum)<br />

and heavy (e.g., copper, bronze) metal casting [4,17]. The resins used for light metals are<br />

urea-modified furan resins, whereas those used for heavy metals contain only furan components.<br />

The hot-box process is well suited for mass production and it consists in mixing<br />

the moist sand with a liquid resin and a curing agent. The ensuing mixture is then cured at<br />

180–260 C in heated core boxes. The main limitation of this process is its extremely long<br />

bench life. The cold-box (or SO2–furan) process is based on curing the reactive resin at<br />

room temperature in a closed-air system with SO2. This gas is converted in situ into a<br />

mixture of sulfurous and sulfuric acids which catalyze the curing.<br />

VI. FURAN RESINS AS WOOD ADHESIVES<br />

Regardless of the fact that numerous investigations exist about the possibility of incorporating<br />

the furan heterocycle into wood adhesive formulations, their industrial exploitation<br />

is still modest. The first suggestion concerning the use of 1 in partial substitution of<br />

formaldehyde in phenolic resins was put forward in 1958 by Baxter and Redfern [18]<br />

who proposed that the furfural units were incorporated into the polymer skeleton following<br />

condensation reactions such as:<br />

The intermediate oligomers such as 13 were then subjected to methylolation with formaldehyde<br />

to form phenolic–furanic–formaldehyde resins, according to:<br />

Copyright © 2003 by Taylor & Francis Group, LLC


The interest in this type of process was, of course, the decrease of formaldehyde<br />

content and, therefore, its lower release during the life cycle of the resin.<br />

This approach was then extended by Pizzi’s group to other phenolic type adhesives<br />

such as phenol–resorcinol–formaldehyde networks [19]. In this work, it was shown that<br />

the addition of 1 gave cold setting resins with performances and costs comparable to those<br />

made using formaldehyde alone. Thus, the phenol–resorcinol–furfural–formaldehyde cold<br />

sets obtained appeared to have a lower bulk shrinkage compared to those prepared without<br />

1. Moreover, it was established that the presence of furfural did not slow down the<br />

curing rate of the resins.<br />

Stamm [20] studied the dimensional stabilization of different woods with 2. Thus,<br />

Douglas fir, Engelman spruce, loblolly pine, and yellow poplar woods were treated with 2<br />

in the presence of zinc chloride, citric acid, or formic acid in order to induce their acidcatalyzed<br />

polymerization. It was established that the maximum antishrinking efficiency<br />

(around 72%) could be reached with a resin level of a minimum of 40% with respect to<br />

oven dried (OD) wood. The optimal amount of each acidic catalyst was also determined.<br />

The curing time was studied for each system and it was shown that the use of 1% zinc<br />

chloride and 6 h of curing time at 120 C gave very satisfactory fracture moduli, toughness,<br />

abrasion resistance, and antishrinking behavior. The only limitation associated with the<br />

possible uses of these systems is the dark color of the final materials.<br />

Dhamaney [21] showed that the addition of furfural into cashew nut shell liquid<br />

adhesives based on phenol–formaldehyde resins, using CuCl2 or CaCO3 as a ‘‘hardener,’’<br />

gave good adhesive bonding for ordinary plywood. Johns et al. [22] prepared white fir<br />

flakeboards using an aqueous solution containing a mixture of ammonium lignosulfonate,<br />

2, and maleic acid as a binder. Before bonding, the wood surface was activated by a nitric<br />

acid treatment. It was shown that the panels thus obtained possessed a higher elasticity<br />

modulus and lower thickness swell and water absorption compared with those prepared<br />

using classical phenol–formaldehyde binders. Nevertheless, the internal bonding and the<br />

rupture modulus were higher for panels obtained using conventional resins. It was also<br />

established that best surface activation was achieved using a 1.5% aqueous solution of<br />

nitric acid (25–40%) with respect to OD wood, since it gave the optimal mechanical<br />

properties for both high and low density panels.<br />

Gupta et al. [23] prepared plywoods from Cedrus deodora and phenol–formaldehyde<br />

resins. They showed that the addition of 5% of 1 to this adhesive did not result in any<br />

appreciable improvement, but the concomitant addition of 10% of coconut shell powder<br />

gave very high failing loads and very low glue failures. Subsequently, in another context,<br />

Pizzi et al. [24] tested different aliphatic aldehydes and 1, in tannin-based adhesives, and<br />

showed that furfural could replace formaldehyde in the manufacture of adhesive resins for<br />

beam lamination. Roczniak [25] studied the thermal properties of phenol–formaldehyde–1<br />

resins, as catalyzed by dichlorohydrin of glycerol, boric acid, hexamethylenetetramine<br />

(HMTA), or p-toluene sulfonic acid. Two main conclusions were drawn from this work:<br />

(i) p-toluene sulfonic acid gave a faster resinification rate and (ii) HMTA led to the highest<br />

thermal resistant resins. Krach and Gos [26] investigated the gluing of large dimension<br />

sawn wood structures using urea–melamine–furfural as a binder. They stated that the<br />

initial wood moisture (8 to 12%) and the time of adhesive spreading (10 to 90 min) did<br />

not influence significantly the strength properties of the glued junction.<br />

Philippou et al. [27] studied the bonding of wood by graft polymerization. They<br />

produced white fir, Douglas fir, and bishop pine particleboards using 2 as well as mixtures<br />

of ammonium lignosulfonate with 2 or with formaldehyde as cross-linking agents. Before<br />

bonding, the wood surface was activated with different amounts of hydrogen peroxide<br />

Copyright © 2003 by Taylor & Francis Group, LLC


(from 0.5 to 4% with respect to OD wood). The amount of the binder was kept constant in<br />

all experiments (7% with respect to OD wood). The internal bond strength of the materials<br />

obtained was found to increase with increasing amounts of hydrogen peroxide, whereas<br />

the thickness swelling followed the inverse trend. The use of both 2 alone and its mixture<br />

with ammonium lignosulfonate showed very good bonding capability. Bishop pine gave<br />

the highest internal bonding and white fir yielded the lowest thickness swelling and water<br />

absorption when the mixture of ammonium lignosulfonate with 2 was used as a binder.<br />

The least efficient adhesive was found to be the formaldehyde–lignosulfonate system. The<br />

differences between wood species were attributed to their different contents of extractives.<br />

In another study, Philippou et al. [28] studied the effect of the composition of the bonding<br />

materials on the properties of Douglas fir particleboards. Thus, the proportion between 2<br />

and ammonium lignosulfonate was varied as follows: 10/0, 9/1, 8/2, 7/3, 6/4, 5/5, 2.5/7.5,<br />

and 0/10. In this work, the wood was also activated by hydrogen peroxide (2% w/w with<br />

respect to OD wood) and the catalysts used were ferric chloride and maleic acid.<br />

Ammonium lignosulfonate without 2 failed to develop resistance to boiling water whereas<br />

2 without ammonium lignosulfonate gave good mechanical and water resistance properties.<br />

However, the use of a mixture containing six parts of lignosulfonate and four parts of<br />

2 yielded boards with the highest internal bond strength and water resistance values.<br />

Increasing the amount of resin with respect to wood was found to produce an increase<br />

in the elasticity and rupture moduli and a decrease in water absorption and thickness<br />

swelling. The boards prepared exhibited strength and resistance to cold and boiling water<br />

comparable to those made using classical phenol–formaldehyde resins. In a third investigation,<br />

Philippou et al. [29] studied the effect of the processing parameters on the mechanical<br />

properties of particleboards made from Douglas fir wood treated with ammonium<br />

lignosulfonate and 2 as a binder in the presence of maleic acid as a catalyst. They showed<br />

that increasing the pressing temperature from 121 to 177 C or the pressing time from 4 to<br />

8 min, progressively enhanced the internal bond strength and the water resistance of the<br />

treated boards. The water resistance was found to be further improved by the addition of a<br />

small amount of wax (0.5% w/w with respect to OD wood) in the binder mixture.<br />

Leitheiser et al. [30] prepared water dilutable furan resins as binders for particleboard<br />

and showed that the resulting composites could be used for exterior applications.<br />

These resins were readily water dilutable and had low viscosities, which made their application<br />

with conventional equipment an easy process. Kelley et al. [31] prepared wood<br />

panels from Acer saccharum var. Marsh. with various binders. They first activated the<br />

surface of the wood by nitric acid and bonded the particles with tannin, 2, and a mixture of<br />

the two, with and without maleic acid. In all cases, the particleboards obtained exhibited<br />

shear strengths as high as that obtained from a control system made with a conventional<br />

phenol–formaldehyde binder. However, the acidic treatment of wood appeared to have<br />

only a slight effect on the mechanical properties of the panels bonded with the tannin–2–<br />

maleic acid system. Subramanian et al. [32] subjected Douglas fir wood flakes to a nitric<br />

acid treatment followed by a grafting reaction with 2(1-aziridinyl)ethyl methacrylate and<br />

2. They showed that the amount of carboxylic acid groups at the wood surface had<br />

increased substantially, thus enhancing its reactivity towards both reagents.<br />

Philippou and Zavarin [33] studied the interactions between lignocellulosic materials,<br />

2, and maleic acid in the presence or absence of hydrogen peroxide. They used white fir<br />

wood flour, microcrystalline cellulose, milled-wood lignin, and ammonium lignosulfonate<br />

and followed their interactions with the binder by differential scanning calorimetry (DSC)<br />

and concluded that a graft copolymerization between hydrogen peroxide activated wood,<br />

2, and ammonium lignosulfonate had occurred. Balaba and Subramanian [34] studied the<br />

Copyright © 2003 by Taylor & Francis Group, LLC


polymerization of 2 catalyzed by the surface acidity resulting from treating wood with<br />

nitric acid. They followed the polymerization by intrinsic viscosity measurements and<br />

showed that there were two reaction regimes. The first was found to obey zero order<br />

kinetics, with an activation energy of 53.4 kJ/mol, whereas the second could not be<br />

exploited because of polymer precipitation following the formation of network structures.<br />

In 1985, experiments on an industrial scale were carried out jointly at Quaker Oats<br />

Chemicals and Collins Pine Company particleboard plants [35]. In these trials 1 was<br />

used as an extender in a polymeric methylene diphenyl isocyanate (MDI) binder<br />

(1:MDI ¼ 1:3 w/w). The main conclusions which could be reached from these trials<br />

were that savings in binder levels, pressing time, and temperature and drying requirements<br />

could be obtained compared with the corresponding performances of standard phenol–<br />

formaldehyde and urea–formaldehyde systems.<br />

Nguyen and Zavarin [36] studied graft polymerization of 2 on cellulosic materials.<br />

They showed that 2 in an aqueous medium at pH 2.0 and 90 C did not copolymerize with<br />

the cellulose surface in the presence of H2O2/Fe 2þ . However, under the same conditions,<br />

poly2 was efficiently grafted onto cellulosic fibers and the amount of homopolymer of 2<br />

was negligible. In these conditions, the amount of grafted poly2 reached 68% w/w with<br />

respect to OD fibers. They also showed that working at higher temperature and with more<br />

concentrated media yielded higher grafting efficiency. Sellers [37] prepared plywoods from<br />

southern pine (major structural species) and yellow poplar (most representative decorative<br />

species) using polymeric methyl diphenyl diisocyanate adhesive in the presence of 1 as a<br />

reactive diluent in order to reduce the adhesive costs. These formaldehyde-free plywood<br />

composites did not suffer delamination after accelerated-aging tests and, although the<br />

interfacial failure did not satisfy the requirements for structural plywood, they approached<br />

or exceeded requirements for decorative applications. Schultz [38] prepared an exterior<br />

plywood resin based on 2 and paraformaldehyde. Three-ply assemblies from yellow pine<br />

were bonded at different processing conditions and showed that the curing time necessary<br />

for these systems was longer than that which was generally required for conventional<br />

gluing systems. The use of veneers with a high moisture content (9.5 instead of 5.1%)<br />

had very negative effects on the strength properties of the plywood prepared. Pizzi [39] also<br />

prepared particleboard urea–furfural–formaldehyde binders. He concluded that a partial<br />

substitution of formaldehyde with 1 led to an enhanced CH2O emission and explained this<br />

unexpected feature in terms of two competitive reactions. In fact, he showed that in the<br />

resins which contained both formaldehyde and 1, the higher stability to hydrolysis of the<br />

1–urea bonds induced the release of formaldehyde from the final product.<br />

New adhesives from furfural-based diamines and diisocyanates were prepared by<br />

Holfinger and coworkers [40,41]. They produced flakeboards alternatively bonded with<br />

phenol–formaldehyde, MDI, and 5,5 0 -ethylidene difurfuryl diisocyanate (14) adhesives<br />

and showed that the strength properties of flakeboards prepared with 14 were slightly<br />

lower than those based on MDI and higher than those prepared with phenol–formaldehyde<br />

resins. Thus, the internal bond strength values of flakeboards bonded with MDI and<br />

14 at 3% resin content, were 1.33 and 0.97 MPa, respectively [41], which are much higher<br />

than the value required by American standard ANSI/A208.1 (0.41 MPa).<br />

Copyright © 2003 by Taylor & Francis Group, LLC


Joshi and Singh [42] showed that about 30% of formaldehyde could be replaced by 1<br />

(obtained from wheat straw) in the formulation of phenol–formaldehyde adhesives. They<br />

used these phenol–1–formaldehyde resins in the preparation of plywoods from Vateria<br />

indica and Toona ciliata and obtained materials with good resistance to boiling water.<br />

These authors mentioned, however, that 1 slowed down the curing rate of the resin and<br />

recommended longer condensation times compared with conventional phenol–formaldehyde<br />

thermosets. Motawie et al. [43] prepared 1 by hydrolysis of Egyptian cotton straw and<br />

prepared different resins by the copolymerization of the in situ formed furfural with phenol,<br />

epichloridrin–phenol, or a bisphenol A-based epoxy prepolymer. The curing of these resins<br />

was investigated using phthalic or maleic anhydride at 170–185 C or using diamines at<br />

room temperature, both in the presence or absence of kaolin as an inorganic filler. Their<br />

properties appeared to be comparable to those of commercially available wood adhesives.<br />

Ellis and Paszner [44] investigated the self-bonding of various lignocellulosic materials<br />

possessing high hemicellulose content through the in situ generation of furanic<br />

derivatives by acid-catalyzed thermal conversion of some saccharidic units. They used<br />

seven different raw materials with increasing pentosan content, i.e., elm, aspen, oak,<br />

and birch woods as well as bagasse, sweetcorn cob, and feed corn cob, with pentosan<br />

contents of 18.8, 19.4, 20.2, 25.5, 27.2, 39.7, and 42.3%, respectively. The pressing temperatures,<br />

pressures, and times tested were in the ranges of 160 to 220 C, 14–20 kg/cm 2 and<br />

2–10 min, respectively. Ammonium sulfate and ammonium chloride were used as catalysts<br />

and their amounts were varied from 0 to 6% w/w with respect to the vegetable material.<br />

The bending strength of the materials obtained was directly proportional to the xylan<br />

content of the initial lignocellulosic source. The optimal amount of catalyst was found to<br />

be around 1.5% w/w based on the natural raw material and the optimal pressing time was<br />

established to be around 6 min. Increasing the wood particle size induced a drastic<br />

decrease in the bending load, whereas an increase in press plate temperature led to a<br />

substantial increase in the mechanical properties of these self-bonding composites.<br />

Gos et al. [45] glued spruce wood (Picea excelsa L.) using three different adhesives,<br />

namely: (i) a phenol–resorcinol binder, (ii) carbamide–melamine–1 resins, and (iii) a<br />

poly(vinyl acetate) glue. They tested the bending elasticity of these glued woods in the<br />

temperature range of 20 to 150 C and a minimum loss of bending strength, when the<br />

temperature increased from 20 to 150 C, was observed when phenol–resorcinol or carbamide–melamine–1<br />

resins were used. Kim et al. [46] synthesized 1-modified phenol–formaldehyde<br />

resol resins by partial substitution of formaldehyde by 1. They tested the<br />

performance of these resins using them as adhesives for oriented strandboards. They<br />

used 13 C-NMR to establish the reaction mechanism between 1 and the other resin<br />

components and isolated and identified convincingly structures 15, 16, and 17. The use<br />

of 1 with 0.25 mole per mole of phenol in phenol–formaldehyde resol resins gave boards<br />

with properties very similar to those obtained by conventional gluing.<br />

Copyright © 2003 by Taylor & Francis Group, LLC


Recently, Schneider et al. [47] fabricated particleboards using poly2–urea–formaldehyde<br />

adhesives (P2-U-F). They observed that the curing time needed for P2-U-F was<br />

double that necessary for classical urea–formaldehyde resins. They also established that<br />

P2-U-F produced boards with lower strength properties, but with higher water resistance,<br />

if classical processing conditions were used. However, at higher resin contents, P2-U-F<br />

gave boards with better mechanical properties. The following optimal conditions were<br />

derived to produce particleboards: a blending time of 10 min, a press platen temperature<br />

of 150 C, 15% of P2-U-F resin with respect to OD softwood, 1.4 min of pressing time per<br />

millimeter thickness, and a board density of 0.67 kg/dm 3 .<br />

Dao and Zavarin [48,49] prepared boards using wood powder and 2 or poly2 as<br />

binders. The wood species was white fir (Abies concolor) which was used as powder<br />

screened to 80 mesh. Compound 2, poly2, and wood were subjected to chemical activation<br />

with hydrogen peroxide/ferrous ions or nitric acid. It was established that an increase in<br />

the degree of polymerization of poly2 yielded boards with increased strength properties<br />

and that poly2 gave materials with higher strength and water resistance properties than<br />

those obtained using 2. They also showed that the addition of the activator to poly2, rather<br />

than to wood, was more efficient. Finally, they also isolated the acetone-soluble fraction of<br />

poly2 (about 73%) and used it as a binder for the same wood samples. They found that the<br />

tensile properties of the corresponding boards exceeded, by over 50%, those of composites<br />

prepared with conventional phenol–resorcinol–formaldehyde resins.<br />

Abd El Mohsen et al. [50] modified classical urea–formaldehyde resins by adding<br />

different amounts of 2 and used them as binders for beech-based plywoods. These modified<br />

resins gave materials with higher shear strength properties (100% increase) in<br />

comparison to unmodified adhesives. They also established the following optimal<br />

formulations: addition of 30, 45, and 60% of 2 to classical urea–formaldehyde resins<br />

and 3, 4.5, and 6% of p-toluene sulfonic acid as a hardener, respectively. Coppock [51]<br />

prepared durable wood adhesives from furfural-based diols, diamines, and diisocyanates.<br />

She then made plywoods or particleboards using modified urea–formaldehyde resins, with<br />

3 and 4 as binders and found that the materials thus obtained showed acceptable mechanical<br />

properties. These properties were not improved by the addition of further modifiers,<br />

such as 5,5 0 -ethylidene furfuryl amine (18). Measurements using DSC showed that 3 did<br />

not react under alkaline conditions, but readily resinified at pH values below 3.0. These<br />

materials were found to have lower formaldehyde emission compared with those made<br />

with unmodified resins. The mechanical performances of flakeboards made with 14<br />

exceeded the industrial standard requirements and were equivalent to those prepared<br />

using MDI. Finally, materials based on 14 in the presence of 3 or 18 as modifiers were<br />

obtained and found to have better performances in comparison to those prepared without<br />

these additives.<br />

Suzuki et al. [52] prepared wood-meal/plastic composites with an average thickness of<br />

4 mm using urea–2 and phenol–1 resins as binders. The molar ratio between urea and 2 was<br />

varied from 9:1 to 1:9. The amount of formaldehyde emission decreased with increasing<br />

Copyright © 2003 by Taylor & Francis Group, LLC


quantities of added 2 and the optimal ratios were found to lie between 2:1 and 1:2.<br />

Hexamethylenetetramine was added to phenol–formaldehyde resins which were formulated<br />

with a molar ratio of 1:3. The bending strengths of composites prepared using urea–2<br />

adhesives were substantially higher than those made using phenol–formaldehyde binder.<br />

More recently, Raknes [53] studied the natural aging of 14 different commercial adhesives<br />

used in plywood manufacturing. He glued spruce (Picea abies) pieces and subjected them to<br />

30 years of natural aging ! He concluded that the shear strength and the water resistance of<br />

samples bonded with ‘‘furfurylated’’ urea–formaldehyde resins (Cascorit 1250 and Dynorit<br />

L166, manufactured by Casco Wood <strong>Adhesives</strong>, Sweden) were still satisfactory.<br />

Kim et al. [54] explored the possibility of using 2 as a cobinder in conventional urea–<br />

formaldehyde adhesives. They successfully prepared water-insoluble poly2 as oil-in-water<br />

emulsions and added them to urea–formaldehyde in different proportions. The ensuing<br />

mixtures were used to produce particleboards from a mixture of southern pine and hardwoods<br />

(75/25). The resin content of these panels was 8% w/w based on OD wood particles<br />

and the catalyst used was ammonium sulfate at a level of 0.3% w/w with respect to the dry<br />

resins. The optimal quantity of added 2 was found to be in the range of 20–30% with<br />

respect to conventional urea–formaldehyde resins. These formulations gave panels with<br />

increased strength and low formaldehyde emission. Russian investigators [55–58] used 5 as<br />

a binder for fir (Abies) plywoods and showed that the properties of these materials met the<br />

Russian standard requirements if pressing time of about 10 min, pressing temperature of<br />

160 C, and a platen pressure of 1.8 MPa were used. Thus, the shear strength of the<br />

plywoods reached almost 1.5 MPa, and their water uptake did not exceed 39%. The<br />

use of clay as a filler (up to 40% w/w with respect to the binder) decreased substantially<br />

the final properties of the materials [57]. Mezhov et al. [59] also studied the furfural<br />

emission from plywoods prepared using 5 as a binder (produced in situ by reaction of 1<br />

with acetone) and showed that it was much lower than that allowed, i.e., 3–5 mg/100 g of<br />

plywood instead of 10 mg/100 g.<br />

VII. FURAN RESINS AS CEMENT ADHESIVES<br />

<strong>Furan</strong> resins have also been extensively used in formulating mortars, grouts, and ‘‘setting<br />

beds’’ for brick lining destined to be exposed to highly corrosive environments, such as<br />

concentrated acids or highly alkaline cleaning solutions [3,16,60–62]. Two techniques are<br />

used in order to realize assemblies, namely tilesetter’s and bricklayer’s methods. The first<br />

method is based on the use of quarry tiles or pavers with smooth surfaces. The second<br />

method consists in using acid-resistant brick linings. Depending on the end use, three types<br />

of bricks are used for the installation of this type of assembly, namely:<br />

(i) Red shale bricks which have the highest resistance to chemical attack. They are<br />

relatively fragile towards thermal and mechanical shocks. Typically, standard<br />

brick dimensions are 20.3 cm by 9.5 cm.<br />

(ii) Fire clay bricks which are less resistant to chemical attack, but much more<br />

stable against thermal and physical shocks. Their standard dimensions are<br />

22.8 cm by 11.4 cm.<br />

(iii) Carbon bricks which are used to withstand hydrofluoric acid, fluoride salts,<br />

and hot, strong alkaline media. They are also very resistant to thermal shocks.<br />

In 1990, 2 was also used in order to prepare low temperature ( 10 C) hardening<br />

epoxy resin mortar adhesive [63]. For this, 2 was added as a reactive diluent to classical<br />

Copyright © 2003 by Taylor & Francis Group, LLC


epoxy resin based on bisphenol A and the adhesive thus obtained was found to have<br />

good mechanical properties. These compositions are presently being produced by the<br />

Chinese Yanan Chemical plant. More recently, 2 was used in polymer compositions<br />

in building and structural repairs and showed properties similar to those obtained with<br />

epoxy resins [64].<br />

VIII. FURAN RESINS/GLASS FIBER COMPOSITES<br />

Corrosion-resistant glass fiber reinforced composites were also produced on the basis<br />

of furfuryl alcohol thermosetting resins [3,16,60]. Thus, many furan-based glass fiber<br />

reinforced materials have been available for many years, particularly for the storage<br />

of chlorinated aromatic and aliphatic hydrocarbon solvents. Amongst the commercial<br />

units available one finds: (i) very large scrubbing towers packed with Raschig rings.<br />

These containers are resistant to hot (up to about 120 C) HCl and organic chlorides;<br />

(ii) large brink mist eliminators typically working close to 85 C; (iii) acid wash surge<br />

tanks used to store waste liquids with a pH of about 2 at temperatures of 55–60 C;<br />

and (iv) dryer exhaust water driven coolers for incoming hot (230 C) acidic HCl and<br />

aromatic vapors. These few examples do not cover all the equipment constructed on the<br />

basis of furan resin reinforced by fiberglass but they show clearly the usefulness of these<br />

materials in different industrial areas. Other applications include the use of 2 as a matrix<br />

for fiberglass in the production of wrappings of pipes carrying corrosive liquids and vapors<br />

[12]. Thus, steel pipes previously coated with bitumen or coal tar pitch can be wrapped<br />

with a bonded glass fiber mat based on this type of resin. In this context, 2 is mixed<br />

with water in the presence of an emulsifying agent and an acid catalyst and the ensuing<br />

emulsion impregnates the mat. Then, the resulting composite is heated in order to<br />

remove the water and induce the acid-catalyzed polycondensation of the matrix.<br />

Amongst the composites used one can cite furfuryl alcohol resins reinforced<br />

with carbon filled woven glass fiber (commercialized under the name of Permanite,<br />

manufactured by the IKO Group, Canada). The main mechanical properties of such<br />

composites are: tensile, shear, flexural, and compressive strengths of 15, 20, 39, and<br />

41 MPa, respectively. Their average density, thermal conductivity, and coefficient of thermal<br />

expansion are 1.57 kg/dm 3 , 3.44 W/(m 2 K) and 1.8 10 5 / C. Permanite-based pipes<br />

are hard, tough, and rigid with exceptional resistance to thermal shocks. They can<br />

be used up to 140 C and should be protected against high tensional, torsional, and<br />

shear loads.<br />

The combination of furanic derivatives with formaldehyde is also used in order to<br />

produce pipes. Haveg 61, manufactured by High Performance Alloys, Inc., Tipton, IN,<br />

can be cited as an example of these resins which are usually filled with acid-digested<br />

asbestos [16]. These composites are resistant to thermal shocks and have been used<br />

continuously at high temperatures (150 C). They have very low electrical and thermal<br />

conductivities. The main mechanical properties of these composites are: ultimate tensile,<br />

shear, and compressive strengths, at 26 C, of 28, 109, and 72 MPa, respectively.<br />

Their coefficient of thermal expansion is 3.2 10 5 / C. The working and hardening<br />

times of these resinous cements depend strongly on the working temperature [12]. Thus,<br />

for example, for carbon filled 2-based resin cement, the following critical values are found:<br />

at 16, 21, and 27 C the working and hardening times are 90, 60, and 30 min and 48, 20,<br />

and 12 h, respectively [12].<br />

Copyright © 2003 by Taylor & Francis Group, LLC


IX. MISCELLANEOUS APPLICATIONS OF FURAN RESINS<br />

Azimov et al. [65] compared the performances of furan resins with those of conventional<br />

phenol–formaldehyde adhesives. They used these binders to assemble aluminumto-aluminum<br />

and glass-to-glass structures and showed that furanic resins gave<br />

much higher rupture moduli in both systems studied. Rassokha and Avramenko [66]<br />

studied similar systems, but gave more information about the furan resin used. They<br />

used 5- and6-based adhesives both in the presence and absence of zeolite-based fillers,<br />

and assembled aluminum-to-aluminum and glass-to-glass structures. They showed that<br />

the use of polyethylene-co-vinylacetate as a filler dispersant gave well dispersed suspensions<br />

and consequently the best mechanical properties of the assembly. Nikolaev<br />

et al. [67] studied the thermal stability of furan resins produced by the reaction<br />

between a furfuryl ether of glycerol and 2,4-toluene diisocyanate. They showed clearly<br />

that the incorporation of this resin into conventional adhesives improved their thermal<br />

stability. The mechanical properties of steel-to-steel assemblies based on these compositions<br />

were found to follow the same tendency as that observed for the thermal<br />

properties.<br />

Poly(hydroxymethyl furfurylidene-acetone) adhesive resins were synthesized and<br />

characterized [68–70] through the 5-formaldehyde adduct (19) and its acid-catalyzed polymerization.<br />

The catalysts used were sulfuric, phosphoric, or p-toluenesulfonic acid. The<br />

authors postulated that the first condensation products resulted from the condensation of<br />

two methylol groups of two 19 molecules (adduct 20). They also proposed a hypothetical<br />

structure of the network formed after curing (21). It seems, however, difficult to envisage<br />

the acid-catalyzed resinification of 19 without the participation of hydrogen atoms at the<br />

C5 position of the furanic ring [2].<br />

Copyright © 2003 by Taylor & Francis Group, LLC


Macro-diisocyanates based on the reaction of an excess of 2,4-toluene diisocyanate<br />

with different poly(dimethylsiloxane)diols of different lengths have been prepared<br />

by Nikolaev et al. [71]. These macro-diisocyanates were reacted with 2 in stoichiometric<br />

proportions and the resulting adduct (22) was cured with a commercial epoxy<br />

resin in the presence of what was termed ‘‘poly(ethylene)-poly(amine)’’ at room tempcerature,<br />

80, and 100 C. The mechanical and thermal properties of steel-to-steel<br />

assemblies joined by these adhesives were better than those obtained using more<br />

common binders.<br />

Bowles et al. [72] studied the copolymerization of different methacrylates with NCOethyl<br />

methacrylate to obtain dental adhesives. Furfuryl methacrylate (23) was among the<br />

monomers tested. The main objective of this investigation was to establish a correlation<br />

between the solubility parameter of the copolymers and their shear strength. It was<br />

moreover shown that the setting time of the furan-based copolymer was very short compared<br />

to that of aliphatic homologues, but its shear strength was relatively low.<br />

Dopico et al. [73] prepared 5- and6-based furan resins which, after acid-catalyzed<br />

polymerization, were subjected to epoxidation with thiokol in different proportions.<br />

In a second series of experiments, 6–7% of 2 was added to the epoxidized resins. They<br />

showed that all these resins presented a lower flexure resistance compared to unmodified<br />

totally furanic binder. Moreover, the addition of 2 was found to induce negative effects on<br />

the mechanical properties of metal-to-metal assemblies.<br />

<strong>Furan</strong> resins have also been used as binders in grinding wheels [4]. In this field,<br />

5–20% of phenolic resin in combination with 1 as a special wetting agent is added to<br />

the abrasive grains and the resulting wheels thereafter coated onto the surfaces of different<br />

substrates. Paper, cloth as well as composites based on glass fiber reinforced films, have<br />

been used as grinding wheel supports. Acid-catalyzed poly2 has also been used in the<br />

aircraft industry as a low-temperature setting adhesive to bond wood and plastic parts.<br />

This adhesive was found to be suitable for assemblies subjected to warping and other<br />

deformations at high temperatures [12].<br />

X. CONCLUSIONS AND SUMMARY<br />

From the above survey, it appears that the industrial use of furanic monomers such as<br />

furfuryl alcohol and furfural, i.e., chemicals based on renewable resources, as binders<br />

in foundry molds is highly successful. Similar furan-based resins can also be used as<br />

efficient adhesives in wood–particle composites and thus are interesting alternatives to<br />

petroleum-based counterparts. The fact that the substitution of formaldehyde by furfural<br />

has not yet met with a reasonable industrial success probably stems from the<br />

higher cost of the furan aldehyde. The increasing pressure on the reduction of formaldehyde<br />

emission and the renewable character of furfural should play in its favor in<br />

the near future.<br />

This chapter has dealt with the use of furanic derivatives as adhesives and binders. It<br />

has been shown that the main industrial applications concern the use of furfural and<br />

Copyright © 2003 by Taylor & Francis Group, LLC


furfuryl alcohol as raw materials for binders for coating different surfaces, namely:<br />

(i) The storage and the transport of hot and highly corrosive fluids such as<br />

chlorinated solvents, acids, and bases. For this purpose, the vessels (e.g.,<br />

tanks, pipes, or towers) are coated with a composite material based on filled<br />

and/or glass fiber reinforced furanic matrix.<br />

(ii) The molding of liquid metals. In this context, foundry molds are produced<br />

from furan derivatives, or in combination with phenolic resins, and are utilized<br />

because of their excellent fire resistance and thermal stability.<br />

(iii) The preparation of highly resistant cements and concretes which are employed<br />

when the object is used as a container for chemicals and/or exposed to corrosive<br />

cleaning agents.<br />

(iv) The preparation of grinding wheels in which furan resins are used to bond<br />

abrasive grains.<br />

In addition to these well known applications, different studies dealing with the use of<br />

furan derivatives as wood adhesives and other miscellaneous applications are presented<br />

and discussed.<br />

REFERENCES<br />

1. A. Gandini and M. N. Belgacem, in Polymeric Materials Encyclopedia (J. Salamone, ed.), Vol.<br />

11, CRC Press, Boca Raton, FL, 1996, pp. 8518–8549.<br />

2. A. Gandini and M. N. Belgacem, Progress Polym. Sci. 22: 1203 (1997).<br />

3. W. J. McKillip and E. Sherman, in Kirk-Othmer Encyclopedia of Chemical Technology<br />

(M. Grayson, ed.), Vol. 11, John Wiley and Sons, New York, 1981, pp. 499–527.<br />

4. A. Gardziella, L. A. Pilato, and A. Knop, Phenolic Resins, 2nd ed., Springer-Verlag, Berlin,<br />

2000, 560 p.<br />

5. N. Galego and A. Gandini, Revista CENIC, Sci. Fis. 6: 163 (1976).<br />

6. A. Gandini, Adv. Polym. Sci. 25: 47 (1977).<br />

7. M. Choura, M. N. Belgacem, and A. Gandini, Macromolecules 29: 3839 (1996).<br />

8. M. Choura, M. N. Belgacem, and A. Gandini, Macromol. Symp. 122: 263 (1997).<br />

9. J. B. Barr and S. B. Wallon, J. Appl. Polym. Sci. 15: 1079 (1971).<br />

10. E. M. Wewerka, J. Polym. Sci., Part A-1 9: 2703 (1971).<br />

11. A. H. Fawcett and W. Dadamba, Macromol. Chem. 183: 2799 (1982).<br />

12. A. P. Dunlop and F. N. Peters, The <strong>Furan</strong>s, Reinhold, New York, 1953.<br />

13. C. N. Bye, in Handbook of <strong>Adhesives</strong> (I. Skeist, ed.), Von Nostrand Reinhold, New York,<br />

1990, p. 148.<br />

14. W. J. McKillip, in <strong>Adhesives</strong> from Renewable Resources (R. W. Hemingway, A. H. Conner, and<br />

S. J. Branham, eds.), ACS Symposium Series 385, American Chemical Society, Washington<br />

DC, 1989, pp. 408–423.<br />

15. R. C. Schmitt, Polym.-Plast. Technol. Eng. 3: 121 (1974).<br />

16. O. H. Fenner and J. E. Callaham, Mater. Perform. 16: 37 (1977).<br />

17. A. Gardziella, A. Kwasniok, and L. Cobos, Modern Casting 86(3): 39 (1996).<br />

18. G. F. Baxter and D. V. Redfern, U. S. Patent 2,861,977 (1958).<br />

19. A. Pizzi, E. Orovan, and F. A. Cameron, Holz Roh Werkst. 42: 467 (1984).<br />

20. A. J. Stamm, in Wood Technology, ACS Symposium Series 43 (I. S. Goldstein, ed.), American<br />

Chemical Society, Washington DC, 1977, pp. 141–149.<br />

21. C. P. Dhamaney, Paintindia 28(10): 27 (1978).<br />

22. W. E. Johns, H. D. Layton, T. Nguyen, and J. K. Woo, Holzforschung 32: 162 (1978).<br />

23. R. C. Gupta, M. S. Rajawat, and P. P. Gupta, Holzforsch. Holzverwert. 31: 87 (1979).<br />

Copyright © 2003 by Taylor & Francis Group, LLC


24. A. Pizzi, D. du T. Rossouw, and G. M. E. Daling, Holzforsch. Holzverwert. 32: 101 (1980).<br />

25. K. Roczniak, in Phenolic Resins: Chemical Applications, Proc. Weyrhaeuser Sci. Symp.,<br />

Vol. 2, June 6–8, 1979, Tacoma, Washington, 1981, pp. 213–222.<br />

26. H. Krach and B. Gos, Ann. Warsaw Agric. Univ., Forest Wood Technol. 29: 57 (1982).<br />

27. J. L. Philippou, W. E. Johns, and T. Nguyen, Holzforschung 38: 119 (1982).<br />

28. J. L. Philippou, E. Zavarin, W. E. Johns, and T. Nguyen, Forest Prod. J. 32: 55 (1982).<br />

29. J. L. Philippou, W. E. Johns, T. Nguyen, and E. Zavarin, Forest Prod. J. 32: 27 (1982).<br />

<strong>30.</strong> R. H. Leitheiser, B. R. Bogner, F. C. Grant-Acquah, W. E. Johns, and W. Plagemann,<br />

J. Adhesion 14: 305 (1982).<br />

31. S. S. Kelley, R. A. Young, R. M. Rammon, and R. H. Gillespie, J. Wood Chem. Technol.<br />

2: 317 (1982).<br />

32. R. V. Subramanian, W. M. Balaba, and K. N. Somasekharan, J. Adhesion 14: 295 (1982).<br />

33. J. L. Philippou, and E. Zavarin, Holzforschung 38: 119 (1984).<br />

34. W. M. Balaba and R. V. Subramanian, Holzforschung 39: 143 (1985).<br />

35. J. W. Frink and H. D. Layton, Proc. Washington State Univ. Int. Particleboard/Composite<br />

Mater. Symp. 19, 1985, pp. 323–347.<br />

36. T. Nguyen and E. Zavarin, J. Wood Chem. Technol. 6: 15 (1986).<br />

37. T. Sellers, Jr., Forest Prod. J. 39: 53 (1989).<br />

38. T. P. Schultz, Holzforschung 44: 467–468 (1990).<br />

39. A. Pizzi, Holz Roh Werkst. 48: 376 (1990).<br />

40. M. S. Holfinger, (1992). Wood adhesives from furfural-based diamines and diisocyanates,<br />

Ph.D. Dissertation, University of Wisconsin, Madison (available from Univ. Microfilms Int.,<br />

Order No. DA9221915, Diss. Abstr. Int. B 1992, 53(6), 3022-3), 1992, 551 pp.<br />

41. M. S. Holfinger, A. H. Conner, L. F. Lorenz, and C. G. Hill, J. Appl. Polym. Sci. 49: 337 (1993).<br />

42. L. Joshi and S. P. Singh, J. Timber Dev. Assoc. India 39: 19 (1993).<br />

43. A. M. Motawie, E. H. Hassan, and M. M. Kamel, Pigment Resin Technol. 22: 4 (1993).<br />

44. S. Ellis and L. Paszner, Holzforschung 48(Suppl.): 82 (1994).<br />

45. B. Gos, K. Jurkowska, and M. Fichtel, Ann. Warsaw Agric. Univ., Forest Wood Technol.<br />

46: 115 (1995).<br />

46. M. G. Kim, G. Boyd, and R. Strickland, Holzforschung 48: 262 (1994).<br />

47. M. H. Schneider, Y. H. Chui, and S. B. Ganev, Forest Prod. J. 46: 79 (1996).<br />

48. L. T. Dao, Furfuryl alcohol-based adhesives in wood bonding by chemical activation, Ph.D.<br />

Dissertation, University of California, Berkeley (available from Univ. Microfilms Int., Order<br />

No. DA9028795, Diss. Abstr. Int. B 1990, 51(5), 2152), 1989, 127 pp.<br />

49. L. T. Dao and E. Zavarin, Holzforschung 50: 470 (1996).<br />

50. F. F. Abd El Mohsen, R. M. Mohsen, and Y. M. A. Ayana, Pigment Resin Technol. 25: 17–20<br />

(1996).<br />

51. K. M. P. Coppock (1996), Durable wood adhesives from furfural-based diols, diamines, and<br />

diisocyanates, Ph.D. Dissertation, University of Wisconsin, Madison (available from Univ.<br />

Microfilms Int., Order No. DA9708820, Diss. Abstr. Int., B 1997, 57(12), 7625), 1996, 317 pp.<br />

52. M. Suzuki, Y. Yamaga, T. Kitagima, S. Takehara, H. Nishiguchi, and M. Takemoto, Shinrin<br />

Kankyo Shigen Kagaku 35: 71 (1997).<br />

53. E. Raknes, Holz Roh Werkst. 55: 83 (1997).<br />

54. M. G. Kim, L. Wasson, M. Burris, Y. Wu, C. Watt, and R. Strickland, Wood Fiber Sci.<br />

30: 238 (1998).<br />

55. I. S. Mezhov, F. N. Karpunin, and S. A. Ugryumov, Derevoobrab. Prom-st. no. 2: 17 (1998).<br />

56. I. S. Mezhov, F. F. Sokolov, and S. A. Ugryumov, Derevoobrab. Prom-st. no. 3: 24 (1998).<br />

57. S. A. Ugryumov, Derevoobrab. Prom-st no. 1: 27 (2000).<br />

58. S. A. Ugryumov, Derevoobrab. Prom-st. no. 2: 28 (2000).<br />

59. I. S. Mezhov, S. A. Ugryumov, and A. I. Glushchenko, Derevoobrab. Prom-st. no. 4: 23 (1998).<br />

60. R. H. Leitheiser, M. E. Londrigan, and C. A. Rude, in Plastic Mortars, Sealants, and Caulking<br />

Compounds, ACS Symposium Series 113 (R. B. Seymour, ed.), American Chemical Society,<br />

Washington DC, 1979, pp. 7–26.<br />

Copyright © 2003 by Taylor & Francis Group, LLC


61. P. I. Kovalenko and V. B. Reznik, Proc. Second Int. Congress on Concrete, October 25–27,<br />

Austin, Texas, 1978.<br />

62. T. Sugama, E. Kukacka, and W. Horn, Cement Concrete Res. 11: 497 (1981).<br />

63. S. Mu and Z. Wang, Polymer and Concrete, Proc. Int. Congr., 6 (Yiunyuan Huang, Keru Wu,<br />

and Zhiyuan Chen, eds.), Int. Acad. Publ., Beijing, China, 1990, pp. 569–575.<br />

64. R. E. Yazev, Stroit. Mater. no. 5: 10 (1995).<br />

65. F. I. Azimov, V. M. Ershov, and L. S. Mindibaeva, Izv. Vyssh. Uchebn. Zaved., Stroit.<br />

Arkhit. no. 3: 119 (1980).<br />

66. A. N. Rassokha and V. L. Avramenko, Plast. Massy no. 5: 63 (1987).<br />

67. V. N. Nikolaev, L. V. Stolyarova, V. I. Chichiblina, L. I. Nikolaeva, and G. M. Zhelonkina,<br />

Plast. Massy no. 8: 62 (1988).<br />

68. K. Roczniak and B. Siepracka, Polimery (Warsaw) 26: 299 (1981).<br />

69. K. Roczniak and B. Siepracka, Polimery (Warsaw) 26: 363 (1981).<br />

70. K. Roczniak, Polimery (Warsaw) 26: 386 (1981).<br />

71. V. N. Nikolaev, S. M. Verkhunov, and N. I. Kol’tsov, Dokl. Akad. Nauk SSSR 319: 368 (1991).<br />

72. C. Q. Bowles, R. G. Miller, C. C. Chappelow, C. S. Pinzino, and J. Eick, J. Biomed. Mater.<br />

Res. 48: 496: (1999).<br />

73. M. Dopico, J. Rieumont, A. Gomez, D. Diaz, and J. Garcia, Revista de Plasicos Modernos<br />

78: 547 (1999).<br />

Copyright © 2003 by Taylor & Francis Group, LLC

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!