Sei sulla pagina 1di 70

47

Adhesives in the Wood Industry


Manfred Dunky
Dynea Austria GmbH, Krems, Austria

I. INTRODUCTION Progress in research and development within the wood-based industry and within the adhesive industry has shown many successes during the past few decades. Notwithstanding this, the industrial requirements of the wood industry still induce technical improvement in the adhesives and their application in this area. What drives this technical development is the search for cheaper, faster-curing, and more complex adhesives. The rst two requirements are caused by the heightened competition within the wood industry and eorts to minimize costs at a certain level of product quality and performance. The requirement more complex stands for new and specialized products and process. Adhesives play a central role within wood-based panels production. The quality of bonding and hence the properties of the wood-based panels are determined mainly by the type and quality of the adhesives. Development in wood-based panels, therefore, is always linked to development in adhesives and resins. Both the wood-based panels industry and the adhesive industry shown a high commitment to and great capability towards innovation. The best evidence for this is the considerable diversity of types of adhesives used for the production of wood-based panels. Well known basic chemicals have been used for a long time for the production of adhesives and their resins, the most important ones being formaldehyde, urea, melamine, phenol, resorcinol, and isocyanate. The greater part of the adhesive resins and adhesives currently used for wood-based panels is produced with these few raw materials. The how to cook the resins and the how to formulate the adhesive therefore become more and more complicated and sophisticated and are key factors to meet todays requirements of the wood-based panels industry. The quality of bonding and hence the properties and performance of the wood-based panels and beams are determined by three main parameters: the wood, especially the wood surface, including the interface between the wood surface and the bondline the applied adhesive the working conditions and process parameters.
Copyright 2003 by Taylor & Francis Group, LLC

Good quality bonding and adequate properties of the wood-based panels can be attained only if each of these three parameters contributes to the necessary extent to the bonding and production process. In this chapter are then covered the types of adhesives used in the wood industry and their characteristics. The inuences on their performance of the adhesives physicochemical characteristics, of their application parameters, of the wood itself, and of the wood composite process parameters are also described. In wood adhesives the application parameters other than the characteristics of the adhesive itself account for around 50% of performance.

II.

TYPES OF WOOD ADHESIVES

In the wood-based panels industry a great variety of adhesives are currently is use. Condensation resins based on formaldehyde represent the biggest volume within the wood adhesives eld. They are prepared by the reaction of formaldehyde with various chemicals such as urea, melamine, phenol, resorcinol, or combinations thereof. At delivery these adhesive resins are mainly liquid and consist of linear or branched oligomers and polymers in aqueous solution or dispersion. During hardening and gelling they convert to three-dimensionally crosslinked and, therefore, insoluble and nonmeltable networks. The hardening conditions used can be acidic (for aminoplastic resins), highly alkaline (for phenolic resins), or neutral to lightly alkaline (for resorcinol resins). Isocyanates [especially polymeric 4,40 -diphenyl methane diisocyanate (PMDI)] are another important chemical group used for various applications in the wood industry, especially for water resistant bonds. In Table 1 are reported the main wood adhesives in use today with their main applications.

III.

OVERVIEW ON REQUIREMENTS CONCERNING WOOD ADHESIVES

Table 2 summarizes the general parameters of importance for wood adhesives. Research and development in adhesives and resins are mainly driven by the requirements of the bonding and production processes and by the intended properties of the wood-based panels. These requirements are summarized in Table 3. The necessity to achieve shorter press times is omnipresent within the woodworking industry, to keep production costs low. An increased production rate gives the chance to reduce production costs. This is only valid when the market is able to absorb such a high level of production. Shorter press times within a given production line and for certain types of wood-based panels can be achieved by, among others: highly reactive adhesive resins possessing rapid gelling and hardening and steep increase in bonding strength even at a low degree of chemical curing highly reactive adhesive glue mixes obtained by the addition of accelerators, special hardeners, crosslinkers, and others the optimization of the pressing process, e.g., by increasing the eect of the steam shock by (i) increased press temperatures, (ii) a more marked dierence in the moisture content between the surface and the core layer of the panel before hot pressing, or (iii) an additional steam injection step. constancy of as many parameters of the production process as possible.
Copyright 2003 by Taylor & Francis Group, LLC

Table 1

Fields of Application for Various Wood Adhesives V20 x xb x x x x x x x V100 V313 FP MDF x x x x x x x x x x x xc x x x x x x PLW x x x x HLB x x x x MH x x ven. xa furn. xa

Adhesive type UF MUF MF/MUF MUPF PF/PUF RF PMDI PVAc old nat.adhesives nat.adhesives inorg.adhesives activation

UF, ureaformaldehyde resin; MUF, melamine fortied UF resin; MF/MUF, melamine and melamineurea resins (MF resins are only used mixed/coreacted with UF resins; MUPF, melamineureaphenolformaldehyde resin; PF/PUF, phenol and phenolureaformaldehyde resin; (P)RF, resorcinol(phenol)formaldehyde resin; PMDI, polymeric methylenediisocyanate; PVAc, polyvinylacetate adhesive; old nat.adhesives, old (historic) natural adhesives (e.g., starch, glutin, casein adhesives); nat.adhesives, natural adhesives (e.g., tannins, lignins, carbohydrates); inorg.adhesives, inorganic adhesives (e.g., cement, gypsum); activation: activation constituents of wood to function as adhesives (i.e., lignin). V20, particleboard according to DIN 68761 (parts 1 and 4, FPY, FPO), DIN 68763 (V20) and EN 312-2 to 4 and 312-6; V100, particleboard according to DIN 68763 and EN 312-5 and 312-7, option 2 (internal bond after boil test according to EN 1087-1); V313, particleboard according to EN 312-5 and 312-7, option 1 (cycle test according to EN 321); FP, hardboard (wet process) according to EN 622-2; MDF, medium density berboard according to EN 622-5; PLW, plywood according to EN 636 with various resistance against inuence of moisture and water; HLB, laminated beams; MH, solid wood panels according to OeNORM B 3021 to B 3023 (prEN 12775, prEN 13353 part 1 to 3, prEN 13017-1 and 2, prEN 13354); ven., veneering and covering with foils; furn., production of furniture. a Partly powder resins. b Boards with reduced thickness swelling, e.g., for laminate ooring. c Special production method.

Table 2 General Requirements for Wood Adhesives Composition, solids content, viscosity, purity Color and smell Sucient storage stability for given transport and storage conditions Easy application Low transport and application risks Proper gluing quality Climate resistance Hardening characteristic: reactivity, hardening, crosslinking Compatibility for additives Cold tack behavior Ecological behavior: Life cycle analysis (LCA), waste water, disposal, etc. Emission of monomers, Volatile organic compounds (VOC), formaldehyde during production of the wood-based panels and during their use

Copyright 2003 by Taylor & Francis Group, LLC

Table 3

Actual Requirements in the Production and in the Development of Wood Adhesives

Shorter press times, shorter cycle times Better hygroscopic behavior of boards (e.g., lower thickness swelling, higher resistance against the inuence of humidity and water, better outdoor performance) Cheaper raw materials and alternative products Modication of the wood surface Life cycle assessment, energy and raw material balances, recycling and reuse Reduction of emissions during the production and the use of wood-based panels

Cheaper raw materials are another way to reduce production costs. This includes, for example, the minimization of the melamine content in a MUF resin, to produce boards with reduced thickness swelling or increased resistance against the inuence of water and high humidity of the surrounding air. Impeding factors (often temporary) can be the shortage of raw materials for the adhesives, as was the case with methanol and melamine during the 1990s. Life cycle analysis and recycling of bonded wood boards also concerns the adhesive resins used, since adhesives and resins are one of the major raw materials in the production of wood-based panels. This includes, for example, the impact of the adhesives on various environmental issues such as waste water and euent management, noxious gas emission during panel production and from the nished boards, or the reuse of panels to burn for energy generation. Furthermore, for certain recycling processes the type of resin has also a crucial inuence on their feasibility and eciency. Gas emission from wood-based panels during their production can be caused by chemicals inherent to wood itself, such as terpenes or free acids, as well as by volatile compounds and residual monomers coming from the adhesive. The emission of formaldehyde especially is a matter of concern, but so are possible emissions and discharges of free phenols or other materials. The formaldehyde emission noted only after panel manufacture and adhesive resin hardening is due, on the one hand, to the residual, unreacted formaldehyde present in ureaformaldehyde (UF)-bonded boards, or as gas trapped in the wood or dissolved in the moisture still present in the panel. On the other hand, in aminoplastic resins the hydrolysis of weakly bonded formaldehyde from Nmethylol groups, acetals, and hemiacetals as well as in more severe cases of hydrolysis (e.g., at high relative humidity) from methylene ether bridges, increases again the content of emittable formaldehyde after resin hardening. In contrast to phenolic resins, a permanent reservoir of potentially emittable formaldehyde is the consequence of the presence of these weakly bonded structures. This explains the continuous, yet low, release of formaldehyde from UF-bonded wood-based panels even over long periods. However, the level of emission depends on the environmental conditions, a fact which may be described by the resin hydrolysis rate which indicates if this formaldehyde reservoir will or will not lead to unpleasantly high emission values [14]. The higher this hydrolysis rate is, the higher is the potential reservoir of formaldehyde which contributes to subsequent formaldehyde emission. The problem of formaldehyde emission after adhesive hardening in panel manufacture can fortunately be regarded today as solved, due to clear and stringent emission regulations in many European and other countries and to successful long term R&D investement by the chemical industry and the wood working industry. The so-called E1-emission class regulations shown in Table 4 for dierent panel products describe the level of formaldehyde emission which is low enough to prevent
Copyright 2003 by Taylor & Francis Group, LLC

Table 4 Actual Regulations Concerning Formaldehyde Emission from Wood-Based Panels According to the German Regulation of Prohibition of Chemicals (formerly Regulation of Hazardous Substances) for E1 Emission Class (the Lowest Emission Types panels) (a) Maximum steady state concentration in a climate chamber: 0.1 ppm (prEN 717-1; 1995) (b) Laboratory test methods (based on experimental correlation experiences): Particleboard: 6.5 mg/100 g dry board as perforator value (EN 120; 1992) MDF: 7.0 mg/100 g dry board as perforator value (EN 120; 1992) Plywood: 2.5 mg/h-m2 with gas analysis method (EN 717-2) Particleboard and MDF: correction of the perforator value to 6.5% board moisture content

any danger, irritation, or inammation of the mucous membranes in the eyes, nose, and mouth. However, it is important that not only the boards themselves, but also veneering and carpenters adhesives, lacquers, varnishes, and other sources of formaldehyde be controlled, since they also might contribute to a close environment formaldehyde steady-state concentration [14].

IV. AMINOPLASTIC ADHESIVE RESINS (UREA RESINS, MELAMINE RESINS) The various aminoplastic resins are the most important class of adhesives in the woodbased panels industry, especially for the production of particleboards and medium density breboard (MDF), and partly also for oriented strandboard (OSB), plywood, blockboards, and some other types of wood panels. They are also used in the furniture industry as well as in carpenters shops. Aminoplastic adhesive resins are formed by the reaction of urea and/or melamine with formaldehyde. Based on the raw materials that are used various types of resins can be prepared, namely: UF MF MUF mUF MF UF MUPF, PMUF ureaformaldehyde resin melamineformaldehyde resin melamineureaformaldehyde cocondensation resin melamine fortied UF resins mixture of an MF and a UF resin melamineureaphenolformaldehyde cocondensation resin.

The most important parameters for the aminoplastic resins are: (a) The type of monomers used. (b) The relative molar ratio of the various monomers in the resin: F/U molar ratio of formaldedhyde to urea F/M molar ratio of formaldehyde to melamine F/(NH2)2 molar ratio of formaldehyde to amide or amine groups, whereby urea counts for two NH2 groups, and melamine for three NH2 groups. (c) The purity of the dierent raw materials, e.g., the level of residual methanol or formic acid in formaldehyde, biuret in urea, or ammeline and ammelide in melamine.
Copyright 2003 by Taylor & Francis Group, LLC

(d)

The reaction procedures used, e.g. the the the the the pH variation sequence temperature variation sequence types and amount of alkaline and acidic catalysts sequence of addition of the dierent raw materials duration of the dierent reaction steps in the cooking procedures.

The production of aminoplastic adhesive resins is usually a multistep procedure where both alkaline and acidic steps occur. Aminoplastic resins can be prepared in a variety of dierent types for all the dierent needs in wood bonding. This can be achieved by just using the three main monomers mentioned above and varying the preparation procedure. A. UF Resins

Ureaformaldehyde resins [19] are based on a series of consecutive reactions of urea and formaldehyde. Using dierent conditions of reaction and preparation a practically endless variety of condensed UF chemical structures is possible. UF resins are thermosetting resins and consist of linear or branched oligomers and polymers always admixed with some amounts of monomers. The presence of some unreacted urea is often helpful to achieve specic eects, e.g., a better storage stability of the resin. The presence of free formaldehyde has, however, both positive and negative eects. On the one hand, it is necessary to induce the subsequent hardening reaction while, on the other hand, it causes a certain level of formaldehyde emission during the hot press, resin hardening cycle. Even in the hardened state, low levels of residual formaldehyde can lead to the displeasing odor of formaldehyde emission from the boards while in service. This fact has changed signicantly the composition and formulation of UF resins during the past 20 years. After hardening, UF resins consist of insoluble, three-dimensional networks which cannot be melted or thermoformed again. In their application stage UF resins are used as water solutions or dispersions or even in the form of still soluble spray dried powders. These, however, in most cases have to be redissolved and redispersed in water for application. Despite the fact that UF resins consist of only the two main components, namely urea and formaldehyde, a broad variety of possible reactions and resin structures can be achieved. The basic characteristics of UF resins can be ascribed at a molecular level to: their high reactivity their waterborne state, which renders these resins ideal for use in the woodworking industry the reversibility of their aminomethylene bridge, which also explains the low resistance of UF resins to water and moisture attack, especially at higher temperatures; this is also one of the reasons for the hydrolysis leading to subsequent formaldehyde emission. The reaction of urea and formaldehyde is basically a two-step process, usually consisting of an alkaline methylolation (hydroxymethylation) step and an acid condensation step. The methylolation reaction, which usually is performed at a high molar ratio (F/U 1.8 to 2.5), is the addition of up to three (four in theory) molecules of bifunctional formaldehyde to one molecule of urea to give methylolureas; the types and the proportions
Copyright 2003 by Taylor & Francis Group, LLC

of the formed methylol groups depend on the molar ratio F/U. Each methylolation step has its own rate constant ki, with dierent values for the forward and the backward reactions. The formation of these methylol groups mostly depends on the molar ratio F/U. The higher the molar ratio used, the higher the molecular weight the methylolated species formed tends to be. The UF resin itself is formed in the acid condensation step, where still the same high molar ratios as in the alkaline methylolation step is used (F/ U 1.8 to 2.5): the methylol groups, urea and the free formaldehyde react with linear and partly branched molecules with medium and even higher molar masses, forming the polydisperse molar mass distribution pattern characteristic of UF resins. Molar ratios lower than approximately 1.8 during this acid condensation step tend to cause resin precipitation. The nal UF resin has a low F/U molar ratio obtained by the addition of the so-called second urea, which might also be added in several steps [8,9]. The second urea process step needs particular care. It is important for the production of resins with good performance, especially at the very low molar ratios usually in use now in the production of particleboards and MDFs. This last step also includes the distillation of the resin solution to usually 66% resin solids content, which is performed by vacuum distillation in the reactor itself or in a thin layer evaporator. Industrial manufacturing procedures usually are proprietary and are described in depth in the literature only in rare cases [711]. The type of bonding between the urea molecules depends on the conditions used: low temperatures and slightly acid pHs favor the formation of methylene ether bridges (CH2 OCH2) and higher temperatures and lower pHs lead preferentially to the formation of more stable methylene bridges (CH2). Ether bridges can be rearranged to methylene bridges by splitting o formaldehyde. One ether bridge needs two formaldehyde molecules and additionally it is not as stable as a methylene bridge, hence it is highly recommended to follow procedures that minimize the formation of such ether groups in UF resins. In the literature other types of resin preparation procedures are also described. Some of these yield uron structures in high proportion [1215] or triazinone rings in the resins [1517]. The latter are formed by the reaction of ammonia or an amine, respectively, with urea and an excess of formaldehyde under alkaline conditions. These resins are used, e.g., to enhance the wet strength of paper. The following chemical species are present in UF resins: free formaldehyde, which is in steady state with the remaining methylol groups and the post-added urea monomeric methylol groups, which have been formed mainly by the reaction of the post-added urea with the high content of free formaldehyde at the still high molar ratio of the acid condensation step oligomeric methylol groups, which have not reacted further in the acid condensation reaction or which have been formed by the above-mentioned reaction of postadded urea molecules with higher molar masses, which constitute the real polymer portion of the resin. The condensation reaction as well as the increase in the molar mass can also be monitored by gel permeation chromatography (GPC) [18,19]. At longer acid condensation steps, molecules with higher molar mass form and the GPC peaks shift to lower elution volumes. Because of the necessity to limit the subsequent formaldehyde emission, the molar ratio F/U has been decreased constantly over the years [20]. The main dierences between
Copyright 2003 by Taylor & Francis Group, LLC

the UF resins with high and low formaldehyde content are the reactivity of the resin due to the dierent contents of free formaldehyde and the degree of crosslinking in the cured network. The main challenge has been to reduce the content of formaldehyde in the UF resins and to achieve this without any major changes in the performance of the resins. In theory this is not possible, because formaldehyde is the reactive partner in the reaction of urea and formaldehyde during the condensation reaction as well as curing. Decreasing the molar ratio F/U means lowering the degree of branching and crosslinking in the hardened network, which unavoidably leads to a lower cohesive bonding strength. The degree of crosslinking is directly related to the molar ratio of the two components. The UF resin formulators have revolutionized UF resin chemistry in the past 30 years. For example, in a straight UF resin for wood particleboard the above mentioned molar ratio F/U was approximately 1.6 at the end of the 1970s. It is now 1.021.08, but the requirements for the boards (e.g., internal bond strength or percent thickness swelling in water) as given in the quality standards are still unaltered. Also the reactivity of the resin during hardening, besides the degree of crosslinking of the cured resins, depends on the availability of free formaldehyde in the system. It has, however, to be considered that it is neither the content of free formaldehyde itself nor the molar ratio which should be taken as the decisive and only criterion for the classication of a resin concerning its subsequent level of formaldehyde emission. In reality the composition of the glue mix as well as the various process parameters during board production also determine the level of formaldehyde emission. Depending on the type of board and the process of application, it is sometimes recommended to use a UF resin with a low molar ration F/U (e.g., F/U 1.03), hence presenting a low content of free formaldehyde; while sometimes the use of a resin with higher molar ratio (e.g., F/U 1.10) to which a formaldehyde catcher has been added in the glue mix will give better results. Which of these two possible ways is the better one in practice can only be decided by trial and error in each case. The higher the molar ratio F/U, the higher is the content of free formaldehyde in the resin. Assuming stable conditions in the resins, which means that, e.g., post-added urea has had enough time to react with the resin, the content of free formaldehyde is very similar even for dierent manufacturing procedures. The content of formaldehyde in a straight UF resin is approximately 0.1% at F/U 1.1 and 1% at F/U 1.8 [1921]. It also decreases with time due to aging reactions where this formaldehyde reacts further. Table 5 summarizes the various inuences of the molar ratio F/U on various properties of woodbased panels. Table 6 summerizes the inuence of the molar rations F/U and F/(NH2)2,

Table 5 Inuence of the Molar Ratio on Various Properties of UF-Bonded Wood-Based Panels Decreasing the molar ratio leads to a decrease of the formaldehyde emission during the production of the wood-based panels the subsequent formaldehyde emission the mechanical properties the degree of hardening the thickness swelling and the water absorption the susceptibility of hydrolysis

an increase of

Copyright 2003 by Taylor & Francis Group, LLC

Table 6 Molar Ratios F/U and F/(NH2)2, Respectively, of Pure and Melamine Fortied UF Resins Currently in Use in the Wood-Based Panels Industry F/U or F/(NH2)2 molar ratio 1.55 to 1.85 Resin type Classical plywood UF resin, also cold setting; use is only possible with special hardeners and additives, e.g., melamine containing glue mixes for an enhanced water resistance UF plywood resin; use for interior boards without special requirements concerning water resistance; to produce panels with low subsequent formaldehyde emission, the addition of formaldehyde catchers is necessary Plywood or furniture resin with low content of formaldehyde; also without addition of catchers, products with a low subsequent formaldehyde emission can be produced E1 particleboard and E1 MDF resins; especially in MDF production further addition of catchers is necessary. Modication or fortication with melamine can be done MDF resins and special glue resins for boards with a very low formaldehyde emission; in most cases modied or fortied with melamine

1.30 to 1.60

1.20 to 1.30

1.00 to 1.10

below 1.00

respectively, of pure and melamine fortied UF resins currently in use in the wood-based panels industry. The molar mass distribution of UF resins is determined by the degree of condensation and by the addition of urea (and sometimes also other components) after the condensation step; this again shifts the resin mass distribution towards lower average molar masses. For this reason the molar mass distribution is much broader than for other polymers: it starts at the low molar mass monomers (the molecular weight of formaldehyde is 30, for urea it is 60) and goes up to more polymerized structures. It is not clearly known, however, what are really the highest molar masses in a UF resin. Molar masses of up to 500,000, determined by light scattering, have been reported [18,22]. The conditions of molecular level shear within the chromatographic columns [23] should guarantee that all physically bonded clusters, caused by the interaction of the polar groups present in the resins and which might simulate too high a molar mass, are separated and that these high numbers between 100,000 and 500,000, measured using low angle laser light scattering (LALLS) coupled to GPC, really do describe the macromolecular structure of a UF resin in the right manner. A second important argument for this statement is the fact that up to such a high molar mass the on-line calibration curve determined in the GPCLALLS run is stable and more or less linear. It does not show any sudden transition as would be the case of a too sharp increase in apparent molar mass if molecular clustering occurred again after the material has passed through the column. The molar mass distribution (and the degree of condensation) is one of the most important characteristics of the resin and it determines several properties of the resin. Consequence of highly condensed resin structures (high molar masses) are: the viscosity at a given solids content increases [19,24] the owing ability is reduced
Copyright 2003 by Taylor & Francis Group, LLC

the the the the the

wetting behavior of a wood surface becomes worse [24] penetration into the wood surface is reduced [25,26] distribution of the resin on the furnish (particles, bers) worsens water dilutability of the resin becomes lower portion of the resin that remains soluble in water decreases [22]

Diluting the resin with a surplus of water causes precipitation of parts of the resin. These parts preferably contain the higher molar mass molecules of the resin and their relative proportion increases at higher degrees of condensation [22]. Information on correlations between the molar mass distribution (degree of condensation) and mechanical and hygroscopic properties of the boards produced, however, is rather rare and often equivocal [7,19,2729]. The inuence of the degree of condensation is mostly felt during the application and the hardening reaction (wetting behavior and penetration into the wood surface which depend on the degree of condensation). At higher temperatures, during the curing hot press cycle, the viscosity of the resin drops, before the onset of hardening again leads to an increase of viscosity. With this temporary lowering of the viscosity the adhesive wetting behavior improves signicantly, but its substrate penetration behavior also changes. The reactivity of an aminoplastic resin seems to be independent of its viscosity (degrees of condensation), at parity of molar ratio. Ferg [30] mentioned that the bonding strength increased with the degree of condensation of the applied UF resin. The higher molar masses (higher viscosity resin fractions) give a more stable glue line and determine the cohesive properties of the hardened resin [7]. Also Rice [29] and Narkarai and Wantanabe [28] reported that the resistance of a bondline against water attack and redrying increased with the viscosity of the resin. The reason again might be that resins with an advanced degree of condensation remain to a greater extent in the glue line, avoiding resin overabsorption by the substrate and hence avoiding starving of the bondline. Rice [29] found an increase of the thickness of the glue line with an increased viscosity of the resin, obviously due to its lower penetration into the wood substrate. However, it must be taken into consideration that the strength and stability of a glue line decrease with increased glue-line thickness [31]. According to the ndings of Sodhi [32] the bonding strength decreases the longer is the waiting time before application of the glue mix. Once the hardening reaction has started and, therefore, the average molar mass has started to increase, the worse the resin wetting behavior and its penetration in the wood surface appears to be. 1. Cold Tack Properties of UF Resins

Cold tack means that the particle mat has attained some strength already after the prepress at ambient temperature, without any hardening reaction having occurred. This green strength is necessary for better handling of the particle mat during transfer on the production line. This can well be the case in multiopening presses, in special forming presses, or in plywood mills, where the glued veneer layers are prepressed to t into the openings of the presses. At least a low level of cold tack is also necessary to avoid blowing out and loss of the ne wood particles from the surface when panels enter a continuous press at high belt speeds. On the other hand, cold tack can lead to agglomeration of ne wood particles and bers in the forming station. Cold tack is generated during the dry out of glue line, and reaches a maximum after a certain period of time. After this point the cold tack decreases again, when the glue line starts to dry out. Both the intensity of the cold tack as well as the optimum length of time
Copyright 2003 by Taylor & Francis Group, LLC

in which it develops after application of the adhesive can be adjusted by the degree of condensation of the resin as well as by using special resin preparation procedures [3335]. Also various additives can increase the cold tack of the adhesive resins, e.g., some thermoplastic polymers such as poly(vinyl alcohol). 2. Isocyanate (PMDI) as Accelerator and Fortier for UF Resins Polymeric methylenediisocyanate (PMDI) can be used as an accelerator and as a special crosslinker for UF resins. UF resins and PMDI can be sprayed separately without prior mixing onto the particles [36,37] or for improved performance the two resins can be premixed and then applied [8,38,39]. In the usual mixing procedure PMDI is pumped under high pressure into the UF resin [40,41]. Usually 0.5 to 1.0% PMDI based on dry particles is used, whereas at the same time the UF gluing factor might be reduced slightly. The specic press time is said to be reduced by up to 1 s/mm. Addition of PMDI to UF resins with a very low molar ratio was also recommended to achieve low formaldehyde emission. The poor properties of the UF resin due to its very low molar ratio can then be improved by the addition of PMDI [4245]. B. Improvement of the Hygroscopic Behavior of Boards by Melamine Fortied UF Resins (MUF, MUPF and PMUF Resins)

The resin used has a crucial inuence on the properties of wood-based panels. Depending on the requirements, dierent resin types are selected for use. Whereas UF resins are mainly used for interior boards (for use in dry conditions, e.g., in furniture manufacturing), a higher water resistance can be achieved by incoroporating melamine and also some phenol into the resin (melamine fortied UF resins, MUF, MUPF, PMUF). The level of melamine addition and especially the resin manufacturing sequence used in relation to how melamine is incorporated in the resin can be very dierent. The dierent types of these resins which exist today are given in Table 7. The dierent resistances of these resins against hydrolysis are based on their dierences at the molecular level. The methylene bridge linking the nitrogens of amido groups can be split rather easily by water attack in UF resins. The same is not so easy in the case of M(U)F resins, mainly due to the much lower water solubility of melamine itself which is a consequence of the water repellency

Table 7 Molar Ratios F/(NH2)2 of MUF/MUPF Resins Currently in Use in the Wood-Based Panels Industry F/(NH2)2 molar ratio 1.20 to 1.35 0.98 to 1.15 Resin type Resins for water resistant plywood, in the case of the addition of a formaldehyde catcher E1 particleboard resin and E1 MDF resin for water resistant boards (PB: EN 312-5 and 312-7; MDF: EN 622-5). For particleboards according to option 1 (V313 cycle test) MUF resins can be used; for boards according to option 2 (V100 2 h boiling test, tested wet) MUPF or MUF with a special approval is necessary. In this case, especially for the MDF production, formaldehyde catchers are added Special resins for boards with very low formaldehyde emission during board service [81,82]

( 1.00

Copyright 2003 by Taylor & Francis Group, LLC

characteristic of the triazine ring of melamine. The equivalent methylene bridge is instead very stable to hydrolytic attack in phenolic resins. The melamine fortied products, however, are much more expensive due to the much higher price of melamine compared to urea. Therefore, the content of melamine in these resins is as high as strictly necessary but always as low as possible. A MUF resin, at parity of all other conditions, yields a lower pH drop after addition of the hardener than a UF resin [46]. This lower drop of the pH due to the buer capacity of the triazine ring of melamine, however, also causes a decrease of the hardening rate of the resin and, therefore, a lengthening of its gel time [1], hence a lengthening of the hot press time is necessary. This is also seen in the shifts of the exothermic dierential scanning calorimetry (DSC) peak of hardening which are observed in thermal experiments [47]. The deterioration of a bondline and hence its durability under conditions of weathering is determined essentially by: The failure of the resin (low hydrolysis resistance, degradation of the hardened resin causing loss of bonding strength). The failure of the interface between the resin and the wood surface (replacement of physical bondings between resin and reactive wood surface sites by water or other nonresin chemicals). The adhesion of UF resins to cellulose is sensitive to water not only due to the already mentioned lability to hydrolysis of the methylene bridge and of its partial reversibility, but also because theoretical calculations have shown that on most cellulose sites the average adhesion of water to cellulose is stronger than that of UF oligomers [8,48]. Thus, water can displace hardened UF resins from the surface of a wood joint. The inverse eect is valid for PF resins [8,49]. The breaking of bondings due to mechanical forces and stresses: water causes swelling and, therefore, movement of the structural components of the wood-based panels (cyclic stresses due to swelling and shrinking, including stress rupture). The durability of a glue line can be enhanced by the incorporation of hydrophobic chains into the hardened network. This was done by introducing urea-capped di- and trifunctional amines containing aliphatic chains into the resin structure or by using the hydrochloride salts of some of these amines as a curing agent [5054]. By this approach some exibility is introduced into the hardened network, which should decrease internal stresses. In UF resins the aminomethylene link is susceptible to hydrolysis and, therefore, it is unstable at higher relative humidity, especially at elevated temperatures [55,56]. Water also causes degradation of the UF resin with greater devastating eect the higher is the temperature of the water in which the boards are immersed. This dierent behavior of boards at dierent temperatures also is the basis for standard tests on which is based the classication of bondlines, resins, and bonded wood products. These classes include the lowest requirements (interior use) for the normal production of UF-bonded boards up to water and weather resistant boards (V100 boiling test, V313 cycle test, water and boil proof (WBP), and others) according to various national and international standard specications. Hardened UF resins can also be hydrolyzed by moisture or water, due to the relative weakness of the bond between the nitrogen of the urea and the carbon of the methylene bridge, and this is especially so at higher temperatures. During this reaction the methylene bridge is eliminated as formaldehyde [57,58]. The amount of liberated formaldehyde can be taken under certain circumstances as a measure of the resistance of the resin against
Copyright 2003 by Taylor & Francis Group, LLC

hydrolysis. The main parameters inuencing the rate and extent of the hydrolysis are temperature, pH, and degree of hardening of the resin [59]. The acid which has induced the hardening of the resin can also and especially induce such a hydrolysis and hence loss of bonding strength. Another approach to increase the resistance of UF resins against hydrolysis is therefore, based on the fact that the resin acid hardening causes acid residues in the glue line. Myers [60] pointed out that in the case of such an acid hardening system the decrease in the durability of adhesive bonds could be initiated both by the hydrolysis of the wood cell wall polymers adjacent to the glue line as well as in the case of UF-bonded products by acidcatalyzed resin degradation. A neutral pH glue line, therefore, should show a distinctly higher hydrolysis resistance. The amount of hardener (acids, acidic substances, latent hardeners) therefore should always be adjusted to the desired hardening conditions (press temperature, press time, and other parameters) and never follow the more the better. Thus, too high an addition of hardener can cause brittleness of the cured resin and a very high acid residue in the glue line. However, glue-line neutralization must not take place as long as the hardening reaction is ongoing, otherwise this would delay or even prevent curing. This aspect is quite a challenge which in practice has not yet really been solved. Higuchi and Sakata [61] found that a complete removal of acidic substances by soaking plywood test specimens in an aqueous sodium bicarbonate solution resulted in considerable increase in water resistance of UF glue lines. Another attempt was made by these authors [62,63] using glass powder as an acid scavenger, which reacts only slowly with the remaining acid of the glue line and, therefore, does not interfere with acid hardening of the resin. Dutkiewicz [64] obtained some good results in the neutralization of the inherent acidity of a hardened UF-bonded glue line by the addition of polymers containing amino or amido groups. All these solutions, however, are not used as yet in broader industrial applications. Laminate oorings require a very low, long term (24 h) thickness swelling of the MDF/high density berboard (HDF) or particleboard cores of which they are composed. Requirements usually are a maximum value of 8 or 10%, sometimes a maximum value of 6% or even lower, all gures based on the original thickness of the board. Such low percentage thickness swelling results cannot usually be obtained by just using straight UF resins, whereas the incorporation of melamine in the resin is a suitable way to achieve the desired results. Other possibilities could be a pretreatment of the particles or the bers (e.g., acetylation) or a special posttreatment of the board. The necessary melamine content in the resin depends on various parameters, e.g., the type of wood furnish, the pressing parameters (pressure prole, density prole), and on resin consumption which can vary between a few percent up to more than 30%, based on liquid adhesive resin. Due to the considerable cost of melamine itself the content of melamine must always be only as high as necessary but as low as possible. Other important parameters are the resin manufacturing procedure, which considerably inuences the thickness swelling of the boards even at the same adhesive solids content and at the same content of melamine. Melamine fortied UF resins and MUF resins can be manufactured in a variety of ways, for example: (i) By cocondensation of melamine, urea, and formaldehyde in a multistep reaction [6569]. In this regard a comprehensive study of the various reaction types was done by Mercer and Pizzi [70]. They especially compared the sequence of the additions of melamine and urea. By mixing of an MF resin with a UF resin according to the desired composition of the resin [7173].

(ii)

Copyright 2003 by Taylor & Francis Group, LLC

(iii)

(iv)

By addition of melamine in various forms (pure melamine, MF/MUF powder resin) to a UF resin during the application of the glue mix. In the case of the addition of pure melamine, a UF resin of a higher molar ratio must be used, otherwise there is not enough formaldehyde available to react with the melamine in order to incorporate it into the resin. Melamine also can be added in the form of melamine salts such as acetates, formates, or oxalates [7478], which decompose in the aqueous resin mix only at higher temperatures and enable some savings of melamine for the same degree of water resistance compared to original MUF resins. Additionally they act as a hardener. Some of the reasons why melamine salts yield a saving in melamine content have also been identied [74].

The higher the content of melamine, the higher is the stability of the hardened resin towards the inuence of humidity and water (hydrolysis resistance) [79,80]. Resins containing melamine can be characterized by the molar ratio F/(NH2)2 (Table 7) or by the triple molar ratio F:U:M. The mass portion of melamine in the resin can be described based on (i) the liquid resin, (ii) the resin solids content, or (iii) the sum of urea and melamine in the resin. One of the most interesting tasks is to clarify if there is a real cocondensation within MUF resins or if two independent networks are formed, which only penetrate each other. The application of MUF resins is very similar to the UF resins, with the dierence that the level of hardener addition is usually much higher. MUPF resins are mainly used for the production of so-called V100 exterior grade boards according to DIN 68763 and EN 312-5 and 312-7, option 2. They contain small amounts of phenol. Production procedures are described in patents and in the literature [8387] and a coreaction has been demonstrated here, although often not contributing to resin eectiveness [83,84,88,89]. PMF/PMUF resins, in which the amount of phenol is much higher than in MUPF resins, usually contain only little or no urea at all. The analysis of the molecular structure of these resins has shown that either there is no cocondensation between the phenol and the melamine, but that there exist two distinct networks [9093], or that cocondensation can indeed occur [88]. The reason for this is the dierent reactivities of the phenol methylols and the melamine methylols, depending under which pH conditions the reaction is carried out.

C. Reactivity and Hardening Reactions During the curing process a three-dimensional network is built up. This leads to an insoluble resin which is no longer thermoformable. The hardening reaction is the continuation of the acid condensation process during resin production. The acid hardening conditions can be adjusted (i) by the addition of a hardener (usually ammonium salts such as ammonium sulfate or ammonium nitrate) or (ii) by the direct addition of acids (maleic acid, formic acid, phosphoric acid, and others) or of acidic substances, which dissociate in water (e.g., aluminum sulfate). Ammonium chloride has not been in use in the particleboard and MDF industry for several years because of the generation of hydrochloric acid during combustion of wood-based panels causing corrosion problems and because of the suspected formation of dioxins [94]. Ammonium sulfate reacts with the free formaldehyde in the resin to generate sulfuric acid, which decreases the pH; this low pH and hence the acid conditions enable the
Copyright 2003 by Taylor & Francis Group, LLC

condensation reaction to restart and nally the gelling and hardening of the resin takes place. The pH decrease takes place with a rate depending on the relative amounts of available free formaldehyde and hardener and is greatly accelerated by heat [46,61]. UF resins dier from other formaldehyde resins (e.g., MF, MUF, and PF) due to their high reactivity, and hence the short hot-press times which are achievable. Hot press times shorter than 4 s/mm board thickness are possible in the production of particleboards with modern, long continuous press lines. This requires highly reactive UF resins, an adequate amount of hardener, as high a press temperature as possible, and a marked dierence in moisture content of the glued wood particles in the surface and the core of the mat before hot pressing. This moisture gradient induces the so-called steam shock eect even without the additional steam injection often used in North American plants. The optimal moisture content of the glued particles is 6 to 7% in the core and 11 to 13% in the surface. The lower the moisture content in the core, the higher the surface moisture content can be. However, a critical total moisture content in the mat must not be exceeded as this might cause problems with steam ventilation and even steam blisters in the panel. For this it is necessary to have low moisture content of the glued core particles and it is necessary to be thrifty with any extra addition of water in the mat core. The lower the resin solids content on the wood, the lower is the amount of water applied to the wood furnish and hence the lower is the moisture content of the glued core particles. For the surface layers, on the other hand, additional water is necessary in the glue mix to increase the moisture content of the glued particles. This additional water, however, cannot be replaced by a higher moisture content of the dried particles themselves before blending, because this water must be available quickly for a strong steam shock eect. This would not be the case if the water would still be present in the wood furnish as the internal wood cell wall moisture content. The mechanism of the hardening reaction of a MUPF/PMUF resin is not really clear. MUF resins harden is the acid range, whereas phenolic resins have their minimum of reactivity under these conditions. There is then the possibility that the phenolic portion of the resin might not really be incorporated into the aminoplastic portion of the resin during hardening. Dierent opinions and confusing reports have been advanced as regards PMF resin hardening. During the hardening of PMF resins either no cocondensation occurs [95] and in the hardened state two independent interpenetrating networks exist, or some cocondensation is reported to occur [88]. Only in model reactions between phenolmethylols and melamine have indications for a cocondensation via methylene bridges between the phenolic nucleus and the amino group of the melamine been found by 1H nuclear magnetic resonance (NMR). In order to increase the capacity of a production line, especially by shortening the panel hot press times, adhesive resins with a reactivity as high as possible should be used. This includes two parameters: a short gel time and a rapid and instantaneous bond strength development, even at a low degree of chemical curing. The reactivity of a resin at a certain molar ratio F/U or F/(NH2)2 is determined mainly by its preparation procedure and the quality of the raw materials used. Figure 1 shows the comparison of two straight low formaldehyde emission (E1) UF resins with the same molar ratio, but prepared according to dierent manufacturing procedures. The dierences between the two resins are clearly evident by their dierent rates of strength increase obtained in the so-called ABES (Automatic Bonding Evaluation System) test [96]. Resin A shows a distinctly quicker increase in bond strength than resin B, a fact which also has been veried in the industrial scale production of boards.
Copyright 2003 by Taylor & Francis Group, LLC

Figure 1 Comparison of two UF resins with the same molar ratio F/U, but with dierent reactivities, due to dierent preparation procedures, tested by means of the Automatic Bonding Evaluation System (ABES) according to Humphrey [96,97]. UF-resin A, UF resin with F/U 1.08 and special preparation procedure for higher reactivity; UF-resin B, traditional UF resin with F/U 1.08. Table 8 Acceleration of Aminoplastic Resins by Addition of an Accelerator [98] Standard glue mix (parts by weight) Component liquid UF resin (F/U 1.05) accelerator hardener solution (ammonium sulfate 20%) formaldehyde catcher (urea) Property calculated molar ratio F/U of the glue mix gelation time at 100 C (s) 100 10 1.05 44 Glue mix with accelerator (parts by weight) 100 2.5 10 2 1.05 36

1. Glue Mixes with Enhanced Reactivity Table 8 describes an example of the use of an accelerator which distinctly increases the gelling rate of a core layer glue mix, hence enabling a signicant shortening of the necessary press time. The quick reaction of the accelerator with the hardener salt generates the acid for the acid-induced hardening reaction of the resin. The accelerator is mixed with the resin just prior to use. Since it does not contain any hardener or acid, there is no limiting pot life of this premix. To compensate for the additional formaldehyde, small amounts of formaldehyde catchers are recommended for addition to the glue mix. 2. Highly Reactive Adhesive Resins in Plywood, Parquet Flooring, and Door Production Plywood, parquet ooring, and doors are usually produced using aminoplastic adhesives. The press time necessary for these applications depends on the press temperature, the total thickness of the wood layers which have to be heated through, and the reactivity of the resin glue mix. Traditional adhesive resin systems need rather long press times due to their
Copyright 2003 by Taylor & Francis Group, LLC

low reactivity, causing a low capacity of the production line. In these systems the hardener is premixed in greater proportion with the adhesive resin. The limitation of these is the too short pot life obtained after hardener addition, causing early gelling of the glue mix in the storage vessel. Using smaller glue mixes increases the chances of improving the resin reactivity. With this presupposition very reactive hardeners can be used. They decrease the gel time of the glue mix and hence the necessary press times. Special aminoplastic resin systems with distinctly higher reactivity have therefore been developed to fulll the requirements of each customer in terms of saving time, energy, and costs. The higher the reactivity, the higher is the capacity of the production line, or the lower is the necessary press temperature at a given press time. Lower temperatures are benecial for the quality of the wood itself as well as for saving energy costs. It has been shown that such very reactive hardeners perform favorably also when in liquid form. This enables the use of various acids or acidic substances in the formulations of these hardeners. Such reactive adhesive systems usually consist of two liquid components, one being a high viscosity resin and the other a high viscosity liquid hardener. The hardener contains some inorganic llers or organic thickeners. The mixing of these two components is performed just prior to the application of the resin mix to the roll coater. The liquidliquid two-component mixer is installed preferably above the roll coater, in order to reduce considerably the amount of each batch of prepared glue mix. If a long stop of the production occurs the lost amount of the ready-to-use glue mix is rather small. Another advantage of this system is that both components can be pumped directly from the storage vessels to the mixer, without the use of any powder. The disadvantage of these two-component systems is the xed ratio between the extender and the hardener. If the amount of extender should be changed, the amount of the hardener itself is also changed and hence the glue mix reactivity and pot life are changed too. Because of the well known marked inuence of the temperature on the pot life, cooling of the whole system is necessary using a chilled water cooler. The raw adhesive resin should have a temperature not higher than 15 C prior to use, which is especially important in summer due to the higher room temperature. Cooling of the adhesive resin can be performed in a small vessel with cooling coils, which is installed between the storage tank and the mixer. Additionally the roll coater itself also needs cooled cylinders in order to stabilize the temperature at approximately 15 C. Figure 2 shows the scheme of a liquid liquid two-component mixing station.

1. 2. 3. 4. 5. 6. 7. 8. 9. 10. 11.

glue resin tank hardener tank twin pumps filter unit flow sensor mixer switchbox distributor level sensor roll coater ventilation valve

Figure 2 Scheme of a liquidliquid two-component glue mixing station.


Copyright 2003 by Taylor & Francis Group, LLC

D. Correlations Between the Composition of Aminoplastic Resins and the Properties of the Wood-Based Panels Not much work has been done up to now concerning the prediction of bond strengths and other board properties based on the results of the analysis of the adhesive resin in its liquid state. What has been investigated and derived up to now are correlation equations that correlate the chemical structures in various UF resins having dierent molar ratios F/U and dierent types of preparations with the achievable internal bond strengths of the boards as well as the formaldehyde emission measured after resin hardening. The basic aim of such experiments is the prediction of the properties of the woodbased panels, hence of the adhesive resin in its hardened state, based on the composition and the properties of the liquid resins used before their hardening. For this purpose various structural components were determined by means of NMR spectroscopy and the ratios of the amounts of the various structural components were calculated, for example: (i) for UF resins: free urea related to total urea methylene bridges with crosslinking related to total sum of methylene bridges sum of methylene bridges in relation to sum of methylols for MF resins: unreacted melamine to monosubstituted melamine unreacted melamine to total melamine number of methylene bridges in relation to the number of methylol groups degree of branching: number of branching sites at methylene bridges in relation to total number of methylene bridges for MUF resins: sum of unreacted melamine and urea to sum of substituted melamine and urea number of methylene bridges in relation to number of methylol groups or to the sum of methylene bridges and methylol groups

(ii)

(iii)

These ratios then are correlated to various properties of the wood-based panels, e.g., internal bond strength or subsequent formaldehyde emission. Various papers in the literature describe examples of such correlations and present workable predictive equations. For UF resins: Ferg [30], Ferg et al. [99,100]; for MF resins: Mercer and Pizzi [101]; for MUF resins: Mercer and Pizzi [102], Panamgama and Pizzi [103]. For certain boards, some good correlations exist. Even these equations, however, cannot predict all properties for all types of UF resins. This is because it must be assumed that a general correlation for various resins and various panels cannot exist. Other correlation equations might have to be used sometime. However, the types of equations that have already been published describe how a universal equation for this task might look. Only the coecient need to be changed from case to case. These results are of some importance, because they show that at least for a certain combination of resin type and board type, correlations do indeed exist. It will be the task for chemists and technologists to evaluate in further detail all possible parameters as well as their inuence on the performance of the resins and the wood-based panels. It can also be assumed that the various parameters mentioned above will also be decisive for other combinations, even if the numerical values of the coecients within individual equations might dier. The range of the molar ratio under investigation in the papers mentioned above was rather broad.
Copyright 2003 by Taylor & Francis Group, LLC

Table 9 UF Resin Glue Mixes for the Production of Particleboard and MDF, Parts by Weight Components/resin mixes PBUF resina,c MDFUF resind MUF resine Water Hardener solutionf Urea solutiong PB-CLa 100 8 up to 5 PB-FLa 100 1020 2 up to 5 PB-CLb 100 15 up to 5 PB-FLb 100 1020 6 up to 5 MDFa 100 3080 2 15

PB, particleboard; CL, core layer; FL, face layer. a For use in dry conditions. b For use in moist conditions. c UF resin with molar F/U 1.03 to 1.08. d UF resin with molar F/U % 0.98 to 1.02. e MUF resin with molar F/(NH2)2 % 1.03 to 1.08. f Ammonium sulfate solution (20%). g Urea solution (40%).

It would not appear to be possible to use these equations for predictions within narrow ranges of molar ratio, e.g., the usual range of an E1 UF resin with approximate F/U 1.03 to 1.10. A method showing how dierent resin preparation procedures, for equal molar ratio resins, can be included in these correlation equations also needs to be developed. E. Glue Resin Mixes

Table 9 summarizes some resin glue mixes for dierent applications in the production of particleboard and MDF. Table 10 summarizes various resin glue mixes for dierent applications in the production of plywood, parquet ooring, and furniture.

V. PHENOLIC RESINS Phenolic resins [phenolformaldehyde (PF) resins] show complete resistance to hydrolysis of the CC bond between the aromatic nucleus and the methylene bridge and, therefore, are used for water and weather resistant glue lines and boards such as water and weather proof particleboards, OSB, MDF, or plywood for use under exterior weather conditions. Another advantage of phenolic resins is the very low formaldehyde emission in service, after hardening, also due to the stability of the methylene bridges between aromatic nuclei. The disadvantages of phenolic resins are the distinctly longer press times necessary for hardening when compared to UF resins, the dark color of the glue line and of the board surface as well as a higher equilibrium moisture content of the boards due to the hygroscopicity of the high alkali content of the board. The preparation procedure of a phenolic resin is a multistage process, characterized by the time, sequence, and amount (in the case of several steps) of the additions of phenol, formaldehyde, and alkali as the most important raw materials. Similarly to all other formaldehyde condensation resins two main reactions predominate: Methylolation: there is no special preference for ortho or para substitution, preference which, however, could be achieved using special catalysts [104106].
Copyright 2003 by Taylor & Francis Group, LLC

Table 10 UF Resin Glue Mixes for the Production of Plywood, Parquet Flooring, and Furniture, Parts by Weight Components/resin mixes UF resin UF resinb UF resinc Extenderd Water Hardener solutione Hardener solutionf Powder hardenerg Powder hardenerh Liquid hardeneri
a

A 100 20 10

B 100 40 1020 20

C 100 10 3

D 100 25

E 100 1020

Glue mix A: standard glue mix. Glue mix B: containing higher proportion of llers than in A. Glue mix C: high solids content, gives an enhanced water resistance to the glue line. Glue mix D: two-component glue mix: liquid resin ready-to-use hardener in powder form, no addition of other components necessary. Glue mix E: twocomponent glue mix: high viscosity liquid resin high viscosity liquid hardener. a UF resin with molar F/U % 1.3. b UF resin with molar F/U % 1.5 to 1.6. c High viscosity UF resin with molar F/U % 1.3 to 1.4. d Extender: rye- or wheat-our. In some cases some inorganic llers are also included. e For example, ammonium sulfate solution (20%). f For example, ammonium sulfateurea solution (20%/20%). g For example, ammonium sulfate in powder form. h Ready-to-use powder hardener, containing powdered hardener, formaldehyde catcher, extenders, and other additives. i High viscosity lled hardener, containing inorganic llers or organic thickeners, hardener, and sometimes some formaldehyde catcher and other additives.

Methylolation is strongly exothermic and includes the risk of an uncontrolled reaction [107]. Condensation methylene and methylene ether linkages are formed; the latter do not exist at high alkaline conditions. During this stage chains are formed, still carrying free methylol groups. The reaction is stopped just by cooling down the preparation reactor thus preventing resin gelling. Phenolic resins contain oligomeric and polymeric chains as well as monomeric methylolphenols, free formaldehyde, and unreacted phenol. The contents of both monomers have to be minimized by the proper preparation procedure. Various preparation procedures are described in the literature and in patents [108117]. Special PF resins consisting of a two-phase system of a highly condensed and insoluble PF resin and a lower condensation standard PF resin have also been prepared [118] and used industrially. Another two-phase resin consists of a highly condensed PF resin still in an aqueous solution and a PF dispersion [119]. The purpose of such special resins is the gluing of panel products of higher moisture content wood, where the danger of overpenetration of the resin into the wood surface would cause a starved glue line and other serious problems. The properties of the resins are determined mainly by the F/P molar ratio, the concentration of phenol and formaldehyde in the resin, the type and amount of the preparation catalyst (in most cases alkaline), and the reaction conditions. The reaction
Copyright 2003 by Taylor & Francis Group, LLC

itself is performed in an aqueous system without addition of organic solvents. The higher the F/P molar ratio, the higher is the reactivity of the resin hence the higher its hardening rate [120], the degree of branching, and the three-dimensional crosslinking. At lower F/P molar ratios linear molecules are formed preferably. Chow et al. [121] found an increase in the bonding strength of plywood with increasing F/P molar ratio; however, the bonding strength remained constant for molar ratios higher than 1.4. This value is still distinctly lower than the common industrial molar ratios of PF resins for wood adhesives. Usually sodium hydroxide is used as a catalyst, in an amount up to one mole per mole phenol (molar ratio NaOH/P), which corresponds to a portion of alkali in the liquid resin of approximately 10% by weight. The pH of a phenolic resin is in the range of 10 to 13. The preponderant part of the alkali is free NaOH, and a smaller part is present as sodium phenate. The alkali is necessary to keep the resin water soluble via the phenate ion formation in order to achieve a degree of condensation as high as possible at a viscosity that still can be used in practice. Additionally the alkali content signicantly lowers the viscosity of the reaction mixture. Thus, the higher the alkali content of the resin, the higher is its possible degree of condensation, hence the greater is the reactivity of the resin and the higher its hardening rate and, therefore, the shorter is the necessary press time. High alkali contents have also some disadvantages. The equilibrium moisture content in humid climates increases with the alkaline content as do some hygroscopic-dependent properties (longitudinal stability, thickness swelling, water absorption), and some mechanical properties (creep behavior) become worse. The alkali content also causes a cleavage of the acetyl groups of the hemicelluloses. This leads to an enhanced emission of acetic acid compared to UF-bonded boards. The higher the alkali content, the higher is the emission of acetic acid. In European Norms EN 312-5 and 312-7 the content of alkali is limited to 2.0% for the whole board and 1.7% for the face layer, both gures being based on oven-dried mass of the board. Besides NaOH, other basic catalysts can be and are used, such as Ba(OH)2, LiOH, Na2CO3, ammonia, or hexamine. However, with some notable exceptions, these are not used in practice. The type of catalyst signicantly determines the properties of the resins [122124]. Replacing alkali in PF-bonded boards could give some advantages. Ammonia being a gas evaporates during the hot press process and does not therefore contribute to the alkali content and the hygroscopicity of the boards. It is important to hold a fairly high pH as long as possible during hot pressing in order to guarantee a high reactivity and hence a short press time [125,126]. The condensation process of PF resins can be followed by monitoring the increase in viscosity and by gel permeation chromatography (GPC) to measure the molar mass distribution. Chromatograms have been obtained by Duval et al. [122], Ellis and Steiner [127], Gobec et al. [128], Kim et al. [129], and Nieh and Sellers [130]. The penetration behavior strongly depends on the molar masses present in the resin: the higher the molar masses (approximately equivalent to the viscosity of the resin at the same solid content), the worse is the wettability and the lower is the penetration into the wood surface [131,132]. The lower molar masses are responsible for the good wettability, however, too low a molar mass can cause overpenetration and hence starved glue lines. Contact angles of phenolic resins on wood increase strongly with the viscosity of the resin, which increases with the molar masses [133]. The higher molar masses remain at the wood surface and form the glue line, but they will not anchor as well in the wood surface. Depending on the porosity of the wood surface, a certain portion with higher molar masses must be present to avoid an overpenetration into the wood, causing a starved glue line; this means a certain ratio between low and high molar masses is necessary
Copyright 2003 by Taylor & Francis Group, LLC

Table 11

Properties of PF Adhesive Resins Particleboard CL Particleboard FL ca. 45 34 23 300500 ca. 1.18 AW100-Plywood 4648 710 69 500800 ca. 1.23

Solids content (%) Total alkali (%) Free alkali (%) Viscosity (mPa s) Density (g/ml)

4648 79 68 300700 ca. 1.23

CL, core layer; FL, face layer; AW100-plywood according to DIN 68705.

[127,130,134141]. Gollob and co-workers [109,142] found a decrease in the wood failure with increased molar mass averages of PF resins. The penetration behavior of resins into the wood surface also is inuenced by various other parameters, such as wood species, amount of glue spread, press temperature, and pressure and hardening time. The temperature of the wood surface and of the glue line and hence the viscosity of the resin (which itself also depends on the degree of advancement of the resin at the time of measuring) inuences the penetration behavior of the resin [143]. Table 11 summarizes the properties of various PF resins. The contents of free monomers (formaldehyde, phenol) depend on the type of the resin and the preparation procedure. Usual values are <0.3 mass% for the free formaldehyde and <0.1 mass% for free phenol. The storage stability of liquid PF resins ranges from a few weeks up to several months, depending on the degree of condensation, the content of alkali, and the viscosity. An important parameter for the length of the possible storage time is the viscosity of the resin, with regards to both the proper application onto the wood surface during blending as well as the danger that the resin might gel in its storage tank. The lower the alkali content, the lower the storage stability. The aging behavior can also be monitored using GPC [144]. The molecular characterization of PF adhesive resins is done in similar way to that of all the other condensation resins by determining: the molar ratios of the main components: F/P/NaOH; F/P; NaOH/P the composition of the resins, based on liquid form of delivery the degree of condensation and molar mass distribution, molar mass averages the content of reactive sites and functional groups and their distribution in the resin, type of bridges between the aromatic rings of the phenol molecule, branching sites and others. Due to the hydrolysis-resistant CC bonding between the aromatic ring and the methylene group it is not possible to determine the molar ratio in the usual chemical way. This is only possible by 13C-NMR. The F/P molar ratio of PF resins is usually between 1.8 and 2.5, depending on the type of resin. The higher the molar ratio, the higher its reactivity as well as its storage stability. However, the hardened resin is more brittle due to its higher level of crosslinking. A. Reactivity and Hardening Reactions

Phenolformaldehyde core layer resins usually have the highest molar masses and hence show a high reactivity and quick gelation. They contain higher amounts of alkali than face
Copyright 2003 by Taylor & Francis Group, LLC

layer resins in order to keep the resin soluble even at higher degrees of condensation. The higher the degree of condensation during the production process (the higher the viscosity), the shorter the gel time [145]. The upper limit of the degree of condensation of the resin during its production process is given by (i) the viscosity of the resin (the resin must be able to be pumped, and a certain storage stability as well as a proper distribution of the resin on the particles during blending is required) and (ii) the ow behavior of the resin under heat, guaranteeing wetting of the unglued matching second wood surface and a sucient penetration into the wood surface. Too high a moisture content of the glued particles limits the level to which it is possible to dilute the resin and its solids content. The hardening of a phenolic resin can be seen as the transformation of molecules of dierent sizes via chains lengthening, branching, and crosslinking to a three-dimensional network with theoretically an endlessly high molar mass. The hardening rate depends on various parameters, such as molar mass, molecular structure of the resin, the portions of various structural elements as well as possible catalysts and additives. Alkaline PF resins contain free reactive methylol groups in sucient number and can harden even without any further addition of formaldehyde, a formaldehyde source, or catalysts. The hardening reaction is initiated by heat only. The methylol groups react to form methylene and methylene ether bridges. Under high temperatures methylene ether bridges can rearrange to methylene bridges. The lowest possible temperature for a suciently fast gel rate is approximately 100 C. In some cases to improve this, potash in the form of a 50 mass% solution is added in the core layer resin mix in an amount of about 3 to 5% potash solid based on resin solids content. Pizzi and Stephanou investigated the dependence of the gel time on the pH of an alkaline PF resin [146]. Surprisingly they found an increase in the gel time in the region of very high pH values (above 10). Standard commercial PF adhesive resins with a content of NaOH of 5 to 10 mass% have exactly such pHs. A decrease of the pH in order to accelerate the hardening process is not possible, because a spontaneous precipitation would occur with such standard PF resins. A change in pH of the resin, however, might occur when the resin comes into contact with a wood surface. Wood is generally acidic in character, and especially with rather acidic wood species, the pH of the resin could signicantly drop when in contact with the wood surface [147]. Lu and Pizzi [148] showed that lignocellulosic substrates had a distinct inuence on the hardening behavior of PF resins, whereby the activation energy of the hardening process was much lower than for the resin alone [149] and the hardening rate much faster [149]. The reason is a catalytic activation of the PF condensation by carbohydrates such as crystalline and amorphous cellulose and hemicellulose. Covalent bondings between the PF resin and the wood, especially lignin, play only a minor role, however. The gelling process can be monitored via DSC, ABES, or dynamic mechanical analysis (DMA). The chemical hardening can be followed also by solid state NMR, looking (i) at the increase of the amount of methylene bridges based on the amount of aromatic rings [123,150,151], (ii) at the portion of 2, 4, 6-three-substituted phenols [151], or (iii) at the ratio between methylol groups and methylene bridges [152,153]. This degree of hardening, however, is not equal to the degree of hardening as monitored by DSC. Plotting one of these degrees of chemical hardening versus the degree of mechanical hardening, as measured, e.g., via ABES or DMA, reveals the hardening pattern of a resin [151,154,155]. An acid- rather than alkali-induced gelling reaction of PF adhesive resins can cause severe deterioration of the wood substrate at the interface and, therefore, its use has lost its signicance in the application of PF resins to bond wood. Pizzi et al. [156] describe,
Copyright 2003 by Taylor & Francis Group, LLC

however, an eective procedure for the self-neutralization of acid hardened PF glue lines. The system is based on a mixture of a complex formed by morpholine and a weak acid in the presence of para-toluene sulfonic acid. The complex decomposes with heat and reforms on cooling to a complex in which the weak acid has been exchanged with the weak base, yeilding an almost neutral glue line. The system prevents, to a considerable extent, the acid deterioration of the wood substrate. Several other attempts based instead on incorporating the acid chemically into the resin or xing the hardeners physically in the glue line have failed [157]. The acceleration of the hardening reaction is possible by using as high a degree of condensation as possible. Another approach is the addition of accelerating esters [146,158], among which, for example, is propylene carbonate [158,159]. The mechanism of this acceleration is not yet completely clear; it might be due to the hydrogen carbonate ion after hydrolysis of propylene carbonate [160] although this has been shown to be unlikely [146,159] or due to the formation of hydroxybenzy alcohols and temporary aromatic carbonyl groups in the reaction of the propylene carbonate with the aromatic ring of the phenol as in the KolbeSchmidt reaction of CO2 with phenol to give salicylic acid [146]. The higher the addition of esters such as propylene carbonate, the shorter the gel time of the PF resin [146]. Other accelerators for PF resins are potash (potassium carbonate), sodium carbonate [95,161], guanidines, or sodium and potassium hydrogen carbonate. Also chemicals inherent to wood might have an accelerating inuence on the hardening reactivity of PF resins [161]. Since phenolic resins for wood bonding harden only thermally, postcuring during hot stacking is very important. In contrast to UF-bonded boards, PF-bonded boards should be stacked as hot as possible to guarantee a maximum postcuring eect. The strength of the panel improves during hot stacking due to continuous slow curing of the PF resin. On the other hand, very high temperatures during stacking might cause partial deterioration of the wood, seen as discoloration.

B.

Modication of Phenolic Resins

1. Post-Addition of Urea The addition of urea to a phenolic resin causes several eects: decrease of the content of free formaldehyde decrease of the viscosity of the adhesive resin acceleration of the hardening reaction via the possible higher degree of condensation of the resin reduction of the costs of the resin. The urea can be added to the nished PF resin or during its manufacture. The distinct decrease of viscosity observed when urea is added to the nished PF resin is caused by the cleavage of hydrogen bonds [162] and by the dilution eect. There is obviously no cocondensation of this postadded urea with the phenolic resin. Urea reacts only with the free formaldehyde of the resin to form methylols which, however, do not react further due to the high pH [163]. Only at high temperatures did Scopelitis and Pizzi [164] suppose some phenolurea cocondensation occurs, but in their case the phenol used was the much more reactive resorcinol. The higher the amount of postadded urea, the worse the properties of the boards. A reason for this might be ureas diluting eect on the PF resin. Surprisingly Oldoerp and
Copyright 2003 by Taylor & Francis Group, LLC

Marutzky [165] found enhanced board properties at higher degrees of addition of uncondensed urea. Since, however, in these experiments the postadded urea could be extracted completely from the boards, no signicant cocondensation between the urea and the phenolic resin could have occurred. Using such PUF resins, the adhesive solids content should be calculated based only on the PF resin solids content in the PUF resin. 2. Cocondensation Between Phenol and Urea A real concondensation between phenol and urea can be performed in three ways: Reaction of methylolphenols with urea [166169]. Acidic reaction of ureaformaldehyde concentrate (UFC) with phenol followed by an alkaline reaction [170,171]. Reaction exclusively under alkaline conditions of urea and phenol in competition with each other leading to reaction of methylol ureas with phenol and PF oligomers, and to reactions of methylolphenols with each other, as well as reactions of methylolphenols with urea [117,172] The kinetics of the cocondensation of monomethylolphenol model compounds and urea under alkaline conditions is reported by Pizzi et al. [173] and Yoshida et al. [174]. The same kinetics but under acidic conditions is described by Tomita and Hse [166,170,175]. The interaction of esters and copolymerized urea on the fast advancement and hardening acceleration of low condensation alkaline PUF resins is reported by Zhao et al. [117,172]. 3. Addition of Tannins The purposes of the addition of tannins are to accelerate the hardening reaction [110,116,176178] and to replace phenol or a part of the PF resin [179185]. 4. Addition of Lignins Lignins can be added to phenolic resins (i) as an extender, e.g., in order to increase the cold tack or to reduce costs, or (ii) to achieve a chemical modication of the resin, whereby the lignin is chemically incorporated into the phenolic resin [186190]. The idea behind this is based on the chemical similarity between the phenolic resin and lignin or between phenol and the phenylpropane unit of the lignin. The lignin can be added at the beginning, during the cooking procedure, or at the end of the condensation reaction. It is not clear whether the lignin is really always incorporated into the phenolic resin or not. In practice lignin is used at present in just a few North American mills, only as a neutral ller/extender in adhesive resins. 5. Addition of Isocyanates Isocyanates [polymeric MDI (PMDI)] as a fortier for phenolic resins have only been used in the past in rare cases. Deppe and Ernst [41] reported a precuring reaction between the isocyanate and the phenolic resin, even if both components had been applied separately to the particles. Hse et al. [36] also found good results with an isocyanate and a PF resin added separately to wood particles. Pizzi and Walton [191] reported on the reactions and their mechanisms of PF resins premixed in the glue mix with nonemulsiable water-based diisocyanate adhesives for exterior plywood. Pizzi et al. [192] reported on the industrial applications of such systems (PF PMDI sometimes tannin accelerator; UF PMDI)
Copyright 2003 by Taylor & Francis Group, LLC

[192194], on the marked curing acceleration of the PF resin by the isocyanate, and on the excellent results which were obtained industrially with these systems. Very recently this approach, due to its excellent performance, lower cost, and ease of preparation has received intense attention, and several other studies on the subject have been recently published [195198].

C. Correlation Between the Composition of a Resin and the Properties of Wood-Based Panels Bonded with the Resin Similar to the investigations described above for the aminoplastic resins, NMR results of the liquid phenolic resins can be correlated with certain board properties [199]. For this purpose various structural components are determined by means of 13C-NMR and the ratios of the amounts of the various structural components are calculated, e.g.: methylol groups to methylene bridges ratio of free ortho and para sites in relation to all possible reaction sites methylol groups in relation to all possible reaction sites methylene bridges in relation to all possible reaction sites ether bridges in relation to all reaction sites These ratios are then correlated to various properties of the wood-based panels, e.g., internal bond strength after boiling or after boiling redrying, and subsequent formaldehyde emission. Also for phenolic resins it is not clear whether universally valid correlation equations will exist or if they will dier for dierent types of resins and boards, although the correlations obtained appear to have a lower coecient of variability than for the aminoplastic resins. Nevertheless, it can be safely assumed that the various parameters used with at least be the same in most cases, even if the actual numerical values of the coecients within the individual equations might dier.

D. Adhesive Resin Glue Mixes Table 12 summarizes the various glue mixes for dierent applications.
Table 12 Examples of PF Glue Mixes for Particleboard, Oriented Strandboard, and Plywood Parts by Weight Components PF-resin A PF-resin B PF-resin C PF-powder resin Water Potash 50% Extender Particleboard core layer 100 6 Particleboard face layer 100 OSB 100 6 OSB 100 AW100-Plywood 100 6 1015

PF-resin A, medium alkali content (810%); PF-resin B, low alkali content (35%); PF-resin C, medium alkali content (68%); PF-powder resin, no addition of water, no dissolving of the powder before blending the strands; Extender: e.g., coconut shell our.

Copyright 2003 by Taylor & Francis Group, LLC

VI. ISOCYANATES A. Adhesives for Wood-Based Panels (Polyisocyanates)

Adhesives based on isocyanates (especially PMDI) have been used for more than 25 years in the wood-based panel industry [40,41,200203], but still have a relatively low consumption volume compared to systems based on UF, MUF, or PF resins. The main application is the production of waterproof panels, but there is also the production of panels from raw materials that are dicult to glue, such as straw, bagasse, rice shells, or sugar cane bagasse. PMDI can be used as an adhesive for wood-based products such as exterior particleboard, exterior OSB, laminated strand lumber (LSL), MDF, or other specially engineered composites. During hot pressing the viscosity of PMDI is lowered, allowing it to ow across and penetrate below the surface, locking in the wood subsurface as has been shown by Roll [204]. The low wetting angle of PMDI compared to water-based condensation resins allows a rapid penetration into the wood surface; however, this also might result in starved bondlines [205]. PMDI is produced during the manufacturing of monomeric MDI. The PMDI produced industrially by phosgenation of di-, tri-, and higher amines contains a mixture of the three dierent MDI isomers, triisocyanates, and dierent polyisocyanates, and thus the structure and the molar mass depend on the number of phenyl groups. This distribution inuences to a great extent the reactivity, but also the usual characteristics such as viscosity, ow, and wetting behavior as well as the penetration behavior into the wood surface. The structure and the molar mass depend on the number of aromatic rings [206]. For PMDI the distribution of the three monomeric isomers has a great inuence on the quality, because the reactivities of the various isomers (4,40 -, 2,40 -, and 2,20 -MDI) dier signicantly [207]. The greater the portion of the 2,20 - and 2,40 -isomers, the lower the reactivity. This can lead to dierent bonding strengths as well as to residual low reactive isomers in the wood-based panels produced. In the monomeric form (MDI) the functionality is 2 and the NCO content is 33.5%, while PMDI has an average functionality of 2.7 with an NCO content of approximately 30.5%. The HCl content is usually below 200 ppm. PMDI is cheaper than pure MDI and has a lower melting point (liquid at room temperature) due to the increased asymmetry. It is less prone to dimerization and, as a consequence, it is more stable during storage than pure MDI. PMDI is used whenever the color of the nished adhesive is not of concern [208]. The excellent application properties of PMDI and of the wood-based panels produced with it are based on the special properties of PMDI, especially the excellent wetting behavior of a wood surface when compared to waterborne polycondensation resins. Due to this fact surfaces with poor wetting behavior such as straw can also be bonded. According to Larimer [209] the wetting angles for PMDI on various surfaces are much lower than for UF resins. Additionally, these resins show a good penetration behavior into the wood surface, which seems to be determined by the small molar mass of PMDI when compared with polycondensation resins. Marcinko et al. [210] found in their measurements, using solid state 13C-NMR, DSC, uorescence microscopy, and DMA, that PMDI could penetrate 510 times further into wood than PF resins. PMDI not only penetrates the macroscopic hollows of the wood substance, but even penetrates the polymer structure of the wood. This enables good mechanical anchoring. The good wetting and penetration behavior of PMDI can sometimes cause starved glue lines. Due to PMDIs high reactivity and its low molar mass, a special interfacial layer between the wood surface and the adhesive appears to form. If hardening is quicker than the thermodynamically induced desorption during the hardening reaction, then a polyurea/biuret network might form
Copyright 2003 by Taylor & Francis Group, LLC

interpenetrating the wood constituents network. Covalent bondings as well as secondary forces can help to avoid desorption reactions during hardening. Johns [211] showed that isocyanates spread easily on a wood surface; 2 to 3% of isocyanate was enough to form a lm completely covering the wood strands, which is not possible even with 6% of a phenolic resin. The good mobility of MDI is based on several parameters [211]: MDI contains no water; it cannot lose its mobility during adsorption on the wood surface its low surface tension (ca. 50 mN/m) compared to water (76 mN/m) its low viscosity. The impossibility of diluting PMDI with water was solved by the introduction of emulsied PMDI, often called EMDI, which allows an even distribution of the adhesive during the gluing process. EMDI is a product of the reaction of PMDI with polyglycols. EMDI is manufactured under high pressure and dispersed in water. The isocyanate group in PMDI is characterized by high reactivity towards all substances which contain active hydrogens. The main hardening reaction proceeds via reaction with water to the nal amide group, while at the same time CO2 is split o. The water necessary to induce the hardening reaction is applied together with the PMDI (spraying together with the PMDI or spraying of an aqueous dispersion of PMDI in water) or is present in wood in sucient amount. The amine group then reacts further with another isocyanate group to form a polyurea structure: RNCO H2 O ! RNH2 CO2 RNH2 OCN R0 ! RNHCONHR0 The reaction of an isocyanate group with a hydroxyl group leads to the so-called urethane bonding: RNCO HOR0 ! RNHCOOR0 Such a reaction can theoretically also occur between an isocyanate group and an OH group of cellulose or lignin to form a covalent bond. These bonds are usually of greater durability than purely physical bonds. If one could manage to force such a reaction to occur in an industrially useful short curing time, when the reaction of the isocyanate groups of the PMDI with water is suppressed, the probability of the formation of such covalent bonds and with this the quality of the bonding should increase, leading to higher bond strengths and especially a higher resistance against the inuence of humidity. If another isocyanate group reacts with an amide hydrogen within the polyurea structure formed, a branching point is formed (biuret group): R00 NCO RNHCONHR0 ! RNCONHR0 j CONHR00 During the hardening of PMDI Frazier et al. [212] have found the formation of urethanes, polyureas, biurets, and triurets/polyurets. The proportions of the various compounds depend on the working and hardening conditions. The forming of the network is especially inuenced by the ratio between the isocyanate and water. The formation of a urethane seems to be possible for low molar mass isocyanates, as e.g., the usual industrial PMDIs, under slightly alkaline conditions. It can also be assumed that the forming of a urethane
Copyright 2003 by Taylor & Francis Group, LLC

especially occurs by reaction with lignin. This bond, however, seems not to be stable at higher temperatures (120 C) for longer times. Hydrophobic polyols should be able to repel and eliminate water from the wood surface and, therefore to fortify the reaction of the isocyanate group with the hydroxyl groups of the wood surface [213]. Umemura et al. [214] and other workers [27] compared the reaction of isocyanate with water and small amounts of polyols using DMA. The bonding strength and the thermal stability increased by adding dipropylene glycol with molar mass in the 4001000 range. Usually no hardeners are added during the production of wood-based panels (particle board, MDF, OSB, engineered wood products) using PMDI as adhesive. With special additives a distinct acceleration of the hardening reaction and hence shorter press times or lower press temperatures can be achieved [209]. This fact is especially interesting for coldsetting systems as well as for the production of particleboards. Possible catalysts are tertiary amines (e.g., triethanol amine, triethylamine, N,N-dimethylcyclohexylamine) and metal catalysts, based on organic compounds of tin, lead, cobalt, and mercury [208,215218]. Compared to other adhesives, PMDI possesses various advantages, but also some disadvantages (Table 13). For the production of plywood the addition of extenders is recommended [221226] or the mixing with other resins [191,192] as alone PMDI cannot be used for plywood. In the production of OSB (especially for the two types OSB/3 and OSB/4 according to European Norm EN 300) often PMDI is used in the core layer. B. Polyurethane Adhesives

Polyurethane adhesives are formed by the reaction of various types of isocyanates with polyols. The polar urethane group enables bonding to various surfaces. Depending on the raw materials used, glue lines with either rubberlike behavior or elastic-to-brittle hard behavior can be achieved. The end groups determine the type of the adhesive, whether it is a reactively or a physically hardening adhesive.
Table 13 Advantages and Disadvantages of PMDI Compared to Other Adhesives, Especially UF Resins Advantages Higher storage stability Formaldehyde-free gluing, despite the fact that formaldehyde is used in the production of MDI/PMDI Higher reactivity Higher bonding strength Higher tolerance against humidity Lower consumption of adhesive Higher price, but this is compensated by the low adhesive consumptions and sometimes shorter press times Adhesion to all other surfaces, e.g., also press platens. This imposes the use of (i) special internal or external release agents [219], (ii) special types of PMDI [220] or (iii) the use in the board surface of adhesives other than isocyanates The necessary use of special emulsiers (EMDI) or special dosing and gluing systems Greater worker protection requirements due to the toxicity and the low but nevertheless existing vapor pressure of monomeric MDI, which need special precautions during use

Disadvantages

Copyright 2003 by Taylor & Francis Group, LLC

One-component isocyanate adhesive systems consist of chains with isocyanate groups on chain ends or on branching sites. These isocyanate groups can react with the moisture content of the surfaces to be bonded, and a hardened system forms from this addition reaction. Thus, at least one of the two surfaces must contain the amount of water necessary for hardening. Due to the high viscosity of these adhesives, dilution with organic solvents or higher temperatures are necessary. Additionally, the adhesive may contain various other components, such as owing agents, llers, antioxidants, bactericides, or dyes. The bondline reaches the necessary green strength within a few hours and hardens over a few days. During the reaction of the isocyanate group with the moisture content of the wood, CO2 is formed, which causes some foaming of the bondline. The bondlines themselves are more or less resistant against humidity and water. The two-component systems consist of (i) a polyol or polyamine and (ii) an isocyanate. The hardening starts with the mixing of the two components. Due to the low viscosities of the two components they can be used without addition of solvents. The weight ratio between the two components determines the properties of the bondline. Linear polyols and low amounts of isocyanates give exible bondlines, whereas branched polyols and high amounts of isocyanates lead to hard and brittle bondlines. The pot life of the two-component systems is determined by the reactivity of the two components, the temperature, and the addition of catalysts, and can vary between 0.5 and 24 h. At room temperature hardening occurs within 3 to 20 h. VII. WOOD ADHESIVES BASED ON NATURAL RESOURCES

Bio-based adhesive resins have been under investigation for a long time; however, extensive industrial application, at least in Europe, has not yet occurred. The use and application of adhesives based on natural and renewable resources is often thought of by the industry as well as the general public as a new approach that requires novel technologies and methods to implement. Despite the increasing trend toward the use of synthetic adhesives, processes based on the chemical modication of natural products oer opportunities for producing a new generation of high performance, high quality products. The distinct advantages in the utilization of natural materials, e.g., lower toxicity, biodegradability, and availability, need to be paralleled by more ecient and lower cost methods of production. Factors such as regional and species variation have to be considered in selecting the optimum feedstock for a particular process; additionally cost-eective manufacturing techniques have to be developed that will enable these materials to capture a wider percentage of the world market. Manufacturers need to have condence that a continual uninterrupted supply of raw material can be sustained throughout the life cycle of a product. It is of equal importance that the feedstock should not be restricted by geographical and climatic conditions or that yield should not dramatically vary when harvested in dierent locations and at dierent times of the year. The key to an increased usage of natural products by industry is in the control of the above variables so that the end performance by the industry remains consistent [227]. A. Tannins

Tannins are polyhydroxyphenols of vegetable origin, which are soluble in water, alcohols, and acetone and can coagulate proteins. They are obtained by extraction from wood, bark, leaves, and fruits. Other components of the extraction solutions are sugars, pectins
Copyright 2003 by Taylor & Francis Group, LLC

and other polymeric carbohydrates, amino acids, and other substances. The nontannin substances can reduce wood failure and decrease water resistance of glued bonds [176]. The polymeric carbohydrates especially increase the viscosity of the extracts. The basic structures of condensed or polyavonoid tannins are [176] in the A-ring: resorcinol, phloroglucinol in the B-ring: pyrogallol, catechol, and more rarely phenol. Depending on the chemical structure of the A-rings two main types can be distinguished: resorcinol type: in mimosa/wattle, quebracho, Douglas r, spruce tannin extracts phloroglucinol type (pine type): most pine species, e.g., Pinus radiata, Pinus patula, Pinus elliotti, Pinus taeda, Pinus pinaster, Pinus halepensis, Douglas r, and Pinus echinata. Pinus brutia and Pinus ponderosa are mixed types with predominant resorcinol character. The disadvantage of the phloroglucinol type is the distinct lower yield during extraction as well as the much higher reactivity of the A-ring towards formaldehyde, which if uncontrolled can cause extremely short pot lives of the glue mix. The disadvantages of these polyphenols are the high viscosity of the solutions in the range of the concentrations of industrial application, due to the polymeric carbohydrates and high molecular weight tannins [228,229], and in some cases their short pot life. The maximum usable concentration of tannin solutions is approximately 40% by mass, except for mimosa where it can be as high as 50%. By selectively removing the polymeric carbohydrates the viscosity can be decreased and with this the possible concentration can be increased. Such purication steps using an ultracentrifuge [229233] and an acid precipitation followed by ltration or centrifugation have been described [185,234]. However, they have not yet been introduced in industrial practice; they are only available at laboratory scale. A further possibility is the optimization of the conditions during the extraction in order to minimize the content of nontannins in the extract. The viscosity of tannin solutions usually increases at higher pHs [185,235,236], but for some tannin types no clear dependence of the viscosity on the pH is shown. The viscosity of an extract increases with the solids content, especially if carbohydrates are present from the extraction step. There are several ways to decrease the viscosity of tannin extracts: Dilution (lower solids content): this leads to increased moisture content of the glued particles (which is not necessarily a disadvantage, since tannins need high moisture contents of the glued particles to guarantee proper ow during pressing) as well as to a decreased content of active adhesive [176]. Degradation of the high molecular carbohydrates, e.g., by NaOH [237,238]. Addition of hydrogen bond breakers, e.g., urea [228,239,240]. Modication of the extract by sulte or bisulte [241]: this modication of the extracts will especially decrease the sometimes high viscosity to achieve a better performance, but also a longer pot life and a better crosslinking will be achieved; however, it can give poor results if too high a level of sulte is used. Modication by treatment with acetic anhydride or maleic acid anhydride as well as NaOH to decrease the viscosity [228,242244]. Tannins are used mostly in the southern hemisphere [176]; applications in Europe are only for niche products with special properties. Depending on the resin content applied to wood, tannins can be used for interior or exterior boards. The necessary crosslinking
Copyright 2003 by Taylor & Francis Group, LLC

is often done by addition of formaldehyde. This, however, can lead to some formaldehyde emission, but this is low due to the phenolic nature of the tannin. Sometimes crosslinking is performed by the addition of isocyanate. Hardening by tannin autocondensation without any aldehyde addition is also possible [245247]. Tannins from mimosa, quebracho, and pine (Pinus radiata) are actually used on an industrial scale for wood gluing. The extraction itself is only performed industrially in the southern hemisphere. The tannins are produced by water extraction of the wood or of the bark. Suitable solvents are water, alcohols [248], or acetone. Some of the parameters which inuence tannin extraction are: temperature [240,248252] addition of various chemicals, e.g., NaOH or sodium carbonate [185,233,234, 248,249,253261], sodium sulte or bisulte [241,248,250,262], and sulte/ bisulte with sodium carbonate with or without urea [240] duration of the extraction [240,258,260] concentration of the extraction solution: ratio of the amount of dry bark to the amount of extraction solvent [260] properties of the raw material: wood species, age, time span between harvesting and extraction, storage conditions, particle sizes [176]. Usually concentrated solutions or spray dried powders are sold [176]. A purication step usually is not done at industrial scale level [176]. As tannins contain many phenolic type subunits one may be tempted to think that they will exhibit a similar reactivity potential to that of phenol and, therefore, procedures used in standard PF production can be transferred to those containing tannin. This, however, is not the case; the real situation is that tannin is far more reactive than unsubstituted phenol due to the resorcinol and phloroglucinol rings present in the tannin structure [263,264]. This increase in hydroxyl substitution on the two aromatic rings imparts an increase in reactivity to formaldehyde 10 to 50 times greater as compared to simple phenol. This whilst initially sounding promising creates additional problems with respect to producing an industrially applicable resin, due to limited pot lives of the ready-to-use formulations [227], although these problems have been solved and solved well even at industrial level [263,264]. Besides tannin autopolymerization, crosslinking usually is achieved via methylene or other bridges in a polycondensation reaction with formaldehyde or isocyanates. Tannins react with formaldehyde similarly to phenol, whereby the nucleophilic sites of the A-ring are more reactive than those of the B-ring. Formaldehyde reacts with a tannin in an exothermic reaction forming methylene bridges, especially between the reactive sites of the tannin A-rings. The reactive sites of the B-ring need a pH of at least 10 [265,266] to react. However, at such a high pH the reactivity of the A-ring becomes so high that no useful pot lives of the glue mix are obtained any more. Due to their size and shape the tannin molecules become immobile already at rather low degrees of condensation, so that formation of further methylene bridges is impeded or hindered, causing a low degree of hardening (crosslinking) [266]. The higher the molar mass of the tannin, the earlier this eect occurs. At neutral pH a rapid reaction of formaldehyde with the sites 6 and 8 on the A-ring takes place. This leads to the advantage that no (high) alkaline pH as for the phenolic resins is necessary to achieve rapid gelling and that a neutral glue line is obtained. A minor disadvantage is the necessary exact adjustment of the pH, because the gelation time varies strongly with the pH [266,267].
Copyright 2003 by Taylor & Francis Group, LLC

From a purely technological point of view the gel time may not be reduced below a certain limit. Decisive factors are the pot life, the viscosity of the tannin solution, and the rate of the steam escaping from the mat and the board during hot pressing. One possible way is the separate addition of the crosslinker, e.g., by dosing paraformaldehyde via a small screw conveyor directly to the particles in the blender. Also a liquid crosslinker, e.g., a urea formaldehyde concentrate (UFC), can be mixed with the tannin solution in a static mixer just prior to the blender. The higher viscosity of the tannin solution at higher pHs, even without addition of the crosslinker, can be overcome by warming to 3035 C or by adding water. A higher moisture content of the glued particles is no disadvantage in tannin adhesives, on the contrary it helps to guarantee a proper ow of the tannin during hot pressing. Possible crosslinkers are formaldehyde as aqueous solution [268], paraformaldehyde [263,265,267,269], UFC [270,271], UF resins [272], aqueous formaldehyde solution emulsied in an oil [273], dimethylolurea [274] or urea and phenol methylols with longer chains to overcome steric hindrance. Tannins can also be hardened by addition of hexamethylenetetramine (hexamine) [275], whereby these boards show a very low formaldehyde emission [269,275281]. The autocatalytic hardening of tannins without any addition of formaldehyde or other aldehyde as crosslinker is possible, if alkaline SiO2 is present as a catalyst at high pH or just as a consequence of the catalysis of the reaction induced by a lignocellulosic surface [282]. 1. Application of Tannins as Adhesives The main parameter for the application of tannins as adhesives for wood-based panels is the content of reactive polyphenols and the reactivity of these components towards formaldehyde. Tannins can be used as adhesives alone (with a formaldehyde component as crosslinker) or in combination with aminoplastic or phenolic resins. These resins can react chemically with the tannin component in a polycondensation reaction, form only two interpenetrating networks, or both. The simplest adhesive mix formulation consists of the tannin solution and powdered paraformaldehyde as crosslinker [283]. The addition of paraformaldehyde can cause in the short term a relatively high level of formaldehyde emission. Glue mixes using paraformaldehyde for the production of particleboards with low formaldehyde emission are described and used industrially [284]. In the literature a large number of papers describe the combinations of tannins with synthetic resins (Table 14).

B.

Lignins

Lignins are large three-dimensional polymers produced by all vascular terrestrial plants; they are second only to cellulose in natural abundance and are essentially the natural glue that holds plant bers together. Lignins are phenolic materials. They are primarily obtained as a byproduct in wood pulping processes with estimates exceeding 75 million tonnes per annum. Therefore great interest exists for possible applications. Lignins of very dierent chemical composition and possible applications in the wood-based panels industry (adhesives, additive for part replacement of adhesives, raw material for synthetic resins) have been described in a large number of papers and patents. Research into lignin-based adhesives dates back more than 100 years with many separate examples of resins involving lignin being cited. In reality, existing applications are very rare. No industrial use as a pure adhesive for wood is currently known despite the fact that considerable research has been directed toward producing wood adhesives from lignins. By themselves lignins oer no advantages in terms of chemical reactivity, product quality, or
Copyright 2003 by Taylor & Francis Group, LLC

Table 14

Combinations of Tannins and Synthetic Resins Description UF resins with tannin molecules as end groups Addition of formaldehyde-rich UF resins Tannin UF resin mix Reaction of UFmethylols, resorcinol, and tannin UF resins with resorcinol end groups tannin Reaction of UFmethylol tannin followed by addition of resorcinol References 272,285 228,272,274 272,286,287 288 288 288

Combination 1. Tannins aminoplastic reins (a) UF resins

(b) UFresorcinoltannin

(c) Pine tannin MF/MUF resins 2. Tannins phenolic resins (a) Cocondensation of tannins with phenol and formaldehyde (b) Tannins as hardening accelerator for alkali hardening PF resins (c) Low molar mass polymethylolphenols (PMP)

289

Replacement of various amounts of phenol of tannin extract Addition of 1020%

179,180,249,253, 254,290295 184,248,266,271, 296 228,266,274,297, 298

(d) mixes of tannins and PF resins; replacement of phenolic resins by tannins

The crosslinking molecules are of greater size than formaldehyde and can, therefore, bridge better the gaps between the reactive crosslinking sites Since the reaction of tannins with PFresols and formaldehyde is accelerated strongly in the alkaline region, the pot life is reduced signicantly, so PF resins with low content of alkali are used

228,234,266,274, 287,289,297, 299301

3. Tannins with enhanced resorcinol content (a) Opening of the heterocyclic ring of the tannin for warm-setting resorcinoltannin resins

Reduction of the necessary resorcinol 302304 addition by modication with sulte (forming of resorcinol end groups from the tannin molecule by sulte-induced cleavage of the heterocyclic ring). No free methylol groups present, addition of formaldehyde as crosslinker is necessary (continued)

Copyright 2003 by Taylor & Francis Group, LLC

Table 14

Continued Description Replacement of resorcinol in a traditional PRF Forming of resorcinol by intermolecular rearrangement of the tannins Replacement of resorcinol in a traditional PRF PRF/tannin and TRF/TRF separate-application fast-set glulam and ngerjointing adhesives Distinct amelioration of the properties, partial reaction of the isocyanate group with the OH groups of the tannins; for a sucient hardening of the tannin the addition of a formaldehyde component seems to be necessary 305 285,302,306,308 References

Combination (b) Reaction of tannin with resorcinol (c) Cold-setting tanninresorcinol resins (TRF)

302,307 309311

(d) Cold-setting, honeymoon, separate application structural exterior adhesives (with and without resorcinol) 4. Tannins isocyanate (PMDI) Isocyanate as crosslinker for polyavonoid tannins

263,264,266,284, 289,312316

color when compared to conventional wood adhesives. The greatest disadvantages of lignins in their application as adhesives are (i) their low reactivity and, therefore, slow hardening compared to phenol due to the lower number of reactive sites in the molecule, causing increased press times, and (ii) the concern over the chemical variation of the feedstock. The chemical structure of lignin is very complex with the added diculty that unlike tannin the individual molecules are not xed to any particular structure, therefore no true generic molecule exists for lignin from softwood, hardwood, or cereals. Lignosulfonates can be added to synthetic glue resins as extenders (by partial replacement of resin). The partial replacement of phenol during the cooking procedure of PF resins has no real industrial importance. 1. Use of Lignins as Adhesives Without Adding Other Synthetic Resins The application of lignins as adhesives is, in principle, possible. Initial attempts needed very long press times due to the low reactivity of lignin (Pedersen process) [317,318]. This process was a condensation under strong acidic conditions, which led to considerable corrosion problems in the plant [318]. The particles are sprayed with spent sulte liquor (pH 3 to 4) and pressed at 180 C. After this step the boards are tempered in an autoclave under pressure at 170200 C, whereby the sulte liquor becomes insoluble by splitting o water and SO2. Shen [319321], Shen and Fung [323] as well as Shen et al. [322,324] modied this process by spraying the particles with spent sulte liquor containing sulfuric acid and pressing them at temperatures well above 210 C. Nimz [317,325] describes the crosslinking of lignin after an oxidation of the phenolic ring in the lignin molecule using H2O2 in the presence of a catalyst, especially SO2 [326]. This leads to the formation of phenoxy radicals and with this to radical coupling (but not to a condensation reaction), whereby inter- and intramolecular CC bonds are formed.
Copyright 2003 by Taylor & Francis Group, LLC

This reaction does not necessarily need heat or acidic conditions, but is accelerated by higher temperatures (maximum 70 C) as well as lower pHs. In this way the disadvantages of the processes mentioned above (high press temperatures, long press times, use of strong acids) can be avoided [317,326]. An oxidative activation of the lignin also can be achieved by biochemical means, e.g., by adding enzymes (phenoloxidase laccase) to the spent sulte liquor, whereby a polymerization via a radical mechanism is initiated. The enzymes are obtained from nutrient solutions of white fungi [327]. The two-component adhesive is prepared by mixing the lignin with the enzyme solution (after ltration of the mycelium). At the beginning of the press cycle the enzyme still works, since it is stable up to a temperature of 65 C. If this temperature is surpassed, the enzyme is deactivated. At such time, however, the number of quinone methides is already high enough to initiate a crosslinking reaction [86,87,327334]. This system, however, is not capable of keeping up with the demands of modern day panel hot press times. To achieve viability this problem was solved by the addition of a smaller than usual amount of isocyanate adhesives. The use of an adhesive thus denies this system any advantage [335]; the enzymatic approach alone only achieves results and pressing times comparable to those of nonenzyme treated nonglued hardboard, a long-existing process and product.

VIII. THERMOPLASTIC WOOD ADHESIVES A. Hot Melts

Hot melts are 100% solid thermoplastic compounds, which are compounded and applied in the molten state at elevated temperature, the resultant properties being obtained by cooling. Due to the quick cooling, bonds can be established in a very short time. Also a hot melt can be melted again, when already in the glue line. The advantages of hot melts are: 100% solid, contain no organic solvents; no water or solvent to be evaporated; low requirements concerning working and environmental safety easy to use, short set time allows high speed operation (up to 100 m/min) rapid bond strength increase high bond strength eective bonding even of dicult-to-bond surfaces: polyethylene (PE), polypropylene (PP), varnishes, and others combination of exibility and toughness adhesion to a wide variety of substrates even without a primer high variability in formulation (color, viscosity during application, temperature, and others) practically unlimited storage life, easy storage no time limitations in application, hence no pot life problems no pollution of machinery and adherends, because of exactly metered application good temperature control during application, easy to use in automated production systems The disadvantages of hot melts are: cold ow: hot melts creep under mechanical load, even far below the melting temperature; bonds can open slowly, this eect being accelerated by higher temperatures
Copyright 2003 by Taylor & Francis Group, LLC

low heat resistance at elevated temperatures due to thermoplastic behavior; loss of bond strength sensitivity of certain substrates to the required application temperature degradation at elevated temperature (color, viscosity) 1. Composition of Hot Melts Polymer. The polymer determines the properties of the hot melt. Variations are possible in molar mass distribution and in chemical composition (copolymers). The polymer is the main component and backbone of the hot melt adhesive blend as it gives strength, cohesion, and mechanical properties (lmability, exibility). Ethylene vinyl acetate (EVA) is the most used type (approximately 80%). It can be varied in viscosity (melt index) and content of acetate within a broad range of values, and once hardened it presents a predominantly amorphous structure. The vinyl acetate groups impart good adhesion ability towards many materials. The low heat stability, however, limits its areas of application. With increasing content of the vinyl acetate comonomer the adhesion ability, the wetting behavior, and the exibility increase, but also the setting time and the price. Heat resistance and cohesion properties become worse. The higher the average molar mass of the polymer, the worse its wetting behavior, but the better the cohesion properties, the heat resistance and temperature resistance, and the higher the melting viscosity at a given temperature. Ethyleneacrylic acid ester copolymers show high heat resistance and high elasticity at low temperatures. Amorphous poly-a-olens (APAOs) are also used as the basic polymer and their main component monomer is propylene. They present better heat resistance than EVA. APAO shows good adhesion properties to nonpolar surfaces, good exibility, and a high resistance to temperature and moisture. Polyamides give the fastest setting speed, good cohesion and very high heat resistance. They are oil and solvent resistant. Due to the narrow melting region (sharp transition between the elastic and plastic areas) a short setting time during cooling is allowed. Depending on their type the melting temperature is between 105 and 190 C. Advantages are the low melt viscosity, high bond strength, and a high green tack. Disadvantages are the high price and the susceptibility for carbonization at high temperatures in the presence of oxygen. Thermoplastic polyurethanes have no reactive isocyanate groups and cannot crosslink. Thermoplastic, linear, and saturated polyesters give, depending on their chemical composition, hard or elastic and tacky bondlines. They have relatively high melt viscosities, and the bondlines are resistant against moisture, water, and ultraviolet (UV) radiation. Tackiers. Tackiers usually are hydrocarbon resins (aliphatic C5, aromatic C9) or natural resins (polyterpenes, rosin and rosin derivatives, tall oil rosin ester). They improve hot tack, wetting characteristics and open time, and enhance adhesion. The content of tackiers in a hot melt can be in the region of 1025% of total material. Other Components. Waxes increase the resistance against water and moisture (hydrophobization) and improve ow and lubricate during application. Inorganic llers (CaCO3 and/or BaSO4) improve cohesion (small particle size) and adhesion, decrease sagging, and improve the price of the product. Pigments are also used, often in the case of white colored hot melts, the most common pigment being TiO2. Plasticizers decrease the viscosity and the heat resistance; they ameliorate the wetting behavior and the exibility of the bondline; however, cold ow can occur. Stabilizers improve the thermooxidative behavior of the hot melt (heat and aging stability).
Copyright 2003 by Taylor & Francis Group, LLC

2. Curing Hot Melts Curing hot melts are easily meltable polyurethane prepolymers (polyaddition of polyvalent alcohols and isocyanate) with reactive isocyanate end groups (N C O), which react with the moisture content of the wood under hardening. This leads to the formation of a crosslinked polyurethane network. Therefore, as thermoplasticity is no longer present, they cannot melt and are insoluble and show good mechanical and chemical resistance. During application a two-step bonding process takes place, the two steps running in parallel, but at dierent rates: (i) (ii) quick physical solidication due to cooling: high green strength for further rapid processing slower chemical hardening by crosslinking: the reaction of the free isocyanate groups is initiated by the moisture content of the surrounding air and of the adherend.

The advantages of the curing hot melts are: Higher resistance against heat, moisture and steam, good aging and long term stability. Higher mechanical bond strength. Lower application temperatures: lower molar masses and lower softening and melting temperatures. Processing of heat susceptible adherends, e.g., PVC foils, is possible, for example at a processing temperature of 70 C. The heat resistance of the bondline is up to 120 C. Good aging resistance The disadvantages of curing hot melts are: They contain monomeric isocyanate, which is toxic, and thus working safety must always be taken into account. They have stricter requirements concerning packaging and application, namely preventing the access of water during storage and application is necessary. They are expensive. Two component curing hot melts consist, for example, of (i) polyamide epoxy, or (ii) a polyol component isocyanate. After the mixing of the two components, they possess only limited pot life.

B.

Poly(Vinyl Acetate) Adhesives

Poly(vinyl acetate) (PVAc) adhesives are another important type of thermoplastic adhesive, especially in furniture manufacturing and carpentry. They form the bondline in a physical process by losing their water content to the two wooden adherends. PVAc adhesives are ready to use, have a short setting time, and give exible and invisible joints. They are easy to clean and show long storage life. Limitations are their thermoplasticity and their creep behavior. Due to the manifold variations available (homo- or copolymerization products, unmodied or modied, with or without plasticizers) PVAc adhesives show a great variety of processing and bonding properties. The various formulations dier in viscosity, drying speed, color of the bondline, exibility or brittleness, hardness or smoothness, and other characteristics. The bonding priniciple of PVAc adhesives is based on the removal of the water by penetration into the wood substrate or by evaporation to the surrounding air. The forming
Copyright 2003 by Taylor & Francis Group, LLC

of the bondline also requires the application of proper pressure. The nal bond strength is reached after migration of the residual water away from the bondline. The minimum temperature of lm formation (or white point) is 418 C, depending on the type of the adhesive and the addition of plasticizers. This temperature is determined mainly by the glass transition temperature Tg of the polymer used which for PVAc is approximately 28 C. Parameters inuencing the drying time are the type of the adhesive, the type of wood surface, the wood substrate absorption behavior, the wood moisture content, relative humidity and temperature of the surrounding air, the amount of adhesive applied, and the temperature of the adhesive and the wood surfaces. Depending on the formulations, various grades of water resistance can be achieved. For the two-component PVAc adhesives, crosslinking and hence a thermosetting behavior is obtained by addition of hardening resins (e.g., based on formaldehyde), complex forming salts [based on chromium (Cr III), e.g., chromium nitrate, or aluminum (Al III), e.g., aluminum nitrate] or isocyanate. The bondlines are then resistant against high temperatures and the inuence of water. The addition of comonomers during polymerization enables a higher exibility to be obtained compared to PVAc homopolymers. This causes also a lower glass transition temperature and a lower minimum lm formation temperature. Possible comonomers are acrylic acid esters (butylacrylate, 2-ethylhexylacrylate), dialkylfumarates, ethylene, and others. Plasticizers soften the lm and increase both adhesion and setting rate. The most common are phthalates, adipates, and benzoates. The amount added can be in a broad range of 1050% by weight. They aect swelling and softening of the PVAc particles and hence ensure the lm-forming capabilities at room temperature, the tack of the still wet and of the dried bondline, and a better water and moisture resistance of the bondline. Disadvantages are the lower resistance of the bondline against heat, possible migration of the plasticizers, and an enhanced cold ow. Fillers (calcium carbonate, calcium sulfate, aluminum oxide, bentonites, wood our) increase the solid content of the dispersion, and they are added up to 50%, based on PVAc. The purpose of their addition is the reduction of the penetration depth, a thixotropic behavior of the adhesive, gap lling properties, and the reduction of the adhesive costs. Disadvantages can be the increase of the white point and possibly the more marked tool wear rate due to greater hardness of the adhesive. Other components in PVAc formulations are defoamers, stabilizers, ller dispersants, preservatives, thickeners (hydroxyethylcellulose, carboxymethylcellulose), poly(vinyl alcohols), starch, wetting agents, tackiers, solvents (alcohols, ketones, esters), ame retardants, and others. The PVAc bond strength decreases at higher temperatures due to the thermoplastic behavior of the adhesive itself. The higher the average molar mass of the polymer, the smaller this temperature-dependent loss of strength. Under long term load, PVAc bondlines are susceptible to cold ow, especially when plasticizers are included in the formulation. Both eects limit the heat resistance of a PVAc bondline and generally the long term strength under load at higher temperatures (>40 C) as well.

IX.

INFLUENCE OF THE ADHESIVE ON THE BONDING PROCESS AND THE PROPERTIES OF WOOD PRODUCTS

For the production of wood-based panels various adhesives are in use, including aminoplastic resins (UF, MUF, MUPF), phenolic resins (PF), and isocyanate (PMDI).
Copyright 2003 by Taylor & Francis Group, LLC

Table 15 Evaluation of the Three Adhesive Types UF, PF, and PMDI with Regard to Various Parameters Property Price Necessary hardening temperature Susceptibility to wood species Eciency Manipulation Resistance against boiling water
Source: Ref. 336.

UF low low high low easy no

PF medium high low medium to high easy high

PMDI high low low high dicult high

The proper choice of the adhesive depends on the required properties of the wood-based panels, on the working conditions during production as well as on the cost of the adhesive system. This does not only include the net price of the adhesive but also the overall cost of the gluing system including glue spread, capacity of the line (necessary press time), and other parameters (Table 15). Environmental aspects can also have a signicant inuence on the choice of the adhesive system. A. Viscosity

The viscosity of a glue mix is determined by the viscosity of the resin (mainly depending on the degree of condensation and the resin solids content) and the composition of the glue mix. If the viscosity or the degree of condensation of the resin is too low, a large portion of the resin might penetrate into the wood, causing a starved glue line. In such a case no true glue line can be formed and hence no bonding strength can be obtained. Conversely, at a too high viscosity there might be a lack of proper wetting by the adhesive of the wood surface opposite to that surface where the adhesive was applied, consequently with no or very low penetration into the word surface and hence no mechanical interlocking of the adhesive into the substrate. Poor bond strength will also be obtained in such a case. Besides the viscosity of the adhesive resin itself, the viscosity of the glue mix also plays an important role in the nal result. A higher dilution of the resin gives a higher volume to be spread and with this a better distribution of the resin on the particles or bers, and thus better bonding strength [337]. This also saves on costs. B. Flow Behavior

The owability of a resin depends on its viscosity and the solids content as well as the changes in the viscosity at elevated temperatures in the hardening glue line. A low owability causes poor penetration of the resin into the wood surface and low bonding strengths. A too high owability, on the other hand, leads to overpenetration of the resin into the wood and hence to starved glue lines. Flowability and hardening act against one another during the hot press curing process. C. Surface Tension and Wetting Behavior Aqueous adhesive resins behave similarly to water regarding surface tension and wetting behavior. For UF resins the wetting behavior strongly depends on their molecular
Copyright 2003 by Taylor & Francis Group, LLC

composition [24]. The higher the F/U molar ratio the lower the surface tension, which also can be decreased by adding a detergent [24] (a practice well known in other wood adhesives too, such as resorcinol cold sets [317]) or a few percent of a PVAc adhesive [24]. The proper wetting of the wood surface is a precondition for achieving high adhesion strength between the resin and the wood surface. D. Reactivity The objective of the development of adhesive resins is to achieve as high reactivities as possible, while maintaining within acceptable limits other properties such as the storage stability of the resin or the pot life of the glue mix. The reactivities of the resin and of the glue mix are determined by various parameters: type of resin composition and preparation procedure type and amount of hardeners additives which might accelerate or retard the hardening process hardening temperature (press temperature, temperature in the glue line, temperature in the core layer) properties of the wood surfaces. E. Liquid and Powdered Resins

In the production of particleboards and MDF only liquid resins are used. In OSB production in Europe liquid resins are more often used, while rather in North America powder resins are used. The advantages and disadvantages of liquid and powder resins are summarized in Table 16. F. Combination of Various Adhesives

For the purpose of obtaining special gluing eects and results, combinations of adhesives or resins might be used, for example: addition of PVAc to UF resins in order to obtain better wetting of the wood surface [24] and a more elastic glue line [338] UF/MUF PMDI (as accelerator, crosslinker and/or fortier) [8,9,192,193] combination of adhesives in particleboard or OSB production: e.g., core layer of PMDI and face layer of MU(P)F resin or PF resin production of an MUF resin by mixing a UF and an MF resin or a UF resin with an MF powder resin.

X. ANALYSIS OF WOOD ADHESIVES BASED ON FORMALDEHYDE CONDENSATION RESINS There has been considerable progress in the characterization of formaldehyde condensation resins in the past two decades. It is now possible to analyze the polydisperse nature of the resins as well as the individual structural elements in the resins, even semi quantitatively. The curing reaction can also be monitored by means of adequate methods. The main topics of analysis are: curing reaction and building up of bonding strength; evaluaCopyright 2003 by Taylor & Francis Group, LLC

Table 16 Type

Advantages and Disadvantages of Liquid and Powder Resins Advantages Low costs No dust-related problems Disadvantages Short storage stability OSB: higher resin load on wood needed because of the poorer adhesive distribution Higher price due to costs for spray drying and packaging Dust-related problems

Liquid resin

Powder resin

Lower resin load on wood and better resin distribution on OSB strands Lower contamination of OSB resin application blenders Longer resin storage stability Quicker gelling as no evaporation of water is necessary

tion and monitoring of the degree of condensation and the molar mass distribution; analysis of the chemical composition of the resins and of their structural components. The characterization of formaldehyde condensation resins was for several decades only possible with basic chemical methods, including elemental analysis [339]. The application of modern spectroscopic and chromatographic methods started as late as the 1970s. One of the reasons for this delay certainly is the fact that condensation resins themselves are still systems that might change during their preparation for analysis or during the analysis itself. Furthermore, the resins polar character as well as their relatively low solubility often render their analysis problematic. Notwithstanding this, the chemical and structural composition of condensation resins is today well known. The validity of each analytical method (Table 17) can be compared and correlated with the information derived from the resins technological behavior and from the properties of the wood panels bonded using these resins.
Table 17 Overview of Various Analysis Methods for Formaldehyde Condensation Adhesive Resins Chemical tests: purity of raw materials content of free formaldehyde during resin preparation and in the nished resins content of formaldehyde in dierent forms in the resins (total formaldehyde, methylol groups) content of urea and melamine content of free and total alkali determination of various molar ratios: F/U; F/(NH2)2; F/P; F/P/NaOH Physical analysis: spectroscopic methods: IR, 1H-NMR, 13C-NMR, 15N-NMR thermal analysis methods for monitoring gelling and hardening processes: DTA, DSC, TMA, DMTA, ABES Physicochemical methods: determination of the molar mass distribution and the average molar masses of the resins [GPC/SEC, GPC-LALLS, vapor pressure osmometry (VPO), light scattering, intrinsic viscosity] chromatographic methods (HPLC, TLC) for the determination of low molar mass species and residual monomers in the resins
Copyright 2003 by Taylor & Francis Group, LLC

Table 18 Property

Basic Technological Tests Test method description Drying the sample for 2 h at 120 C; results can be inuenced by the test parameters Using a rotation viscometer or Ford cup (DIN cup) Direct measurement using pH electrodes Simplied method to determine the resins gel time Gel time at 100 C or at 70 C Pot life at 20 C or at 30 C B-time for PF resins at 100 to 140 C*

Solids content Viscosity pH Gel time and pot life

*Chapter 26, page 556.

A.

Laboratory Test Results

The properties of a resin which can be determined by simple test methods are shown in Table 18. The solids content of a resin usually is determined by the so-called dish method at 120 C for 2 h [dierent times and sometimes lower temperatures (105 C) are often used as several variations of this method exist]. Even if it is a rather simple test, some deviations in the results might occur because not only does all water present as solvent in the liquid adhesive resin evaporate, but also a further condensation reaction with further water elimination takes place. Both liberate condensation water and this additional water is evaporated as well. The more severe the conditions during drying, the lower the solids content measured. Also some details of the test, such as the type of oven, the number of dishes in the oven at the same time, or recirculation of air or not, can inuence the results of the test. The refractive index can be used as a quick method for the determination of solids content, however, the correlation between these two characteristic resin values is sometimes rather poor and not the same for all resins. The density is only important when using volumetric adhesive dosing systems, but not as a quality parameter of the adhesive. One of the most important characteristics is the reactivity of the adhesive resin. With some methods also the start and the end point of the gelling process, the duration of its time span, the behavior of the resin during the test as well as the shape and strength of the gelled plug obtained are essential features of the gel time test. Gelling can occur within one or two seconds (as is usual for UFs) or gelation can span ten or more seconds (as is usual for melamine fortied resins). A long gel time can indicate a slow generation of cohesive bonding strength in the actual application of the resin. The behavior of the resin in the test tube (e.g., foaming) and also the consistency and strength of the gelled plug can be evaluated. The temperature used for the gel time test should always be adjusted to the temperature of application of the resin. If the maximum temperature in a glue line during pressing is not higher than 70 C, then the gel test should be performed at such a temperature and not at water boiling point. This is recommended in order to better interpret the behavior of the resin or the resin glue mix under its conditions of industrial application. B. Chemical Composition of Adhesive Resins

The various components and raw materials of the resins can be determined using dierent chemical methods (Table 19). The content of total formaldehyde is accessible by hydrolysis of an aminoplastic resin; this process, however, is not possible for PF resins. Urea can be determined in the easiest way from the resin nitrogen content. However, other possible
Copyright 2003 by Taylor & Francis Group, LLC

Table 19 Component

Parameters to Be Determined in Adhesive Resins Analysis urea melamine phenol formaldehyde (total formaldehyde, methylol groups) alkali (free alkali, total alkali, ash) free formaldehyde unreacted urea free phenol F/U for a straight UF resin F/M for an MF resin F/(NH2)2 for an MUF resin F/P or F/P/NaOH for a PF resin

Analysis of residual monomers

Molar ratios

sources of nitrogen have to be taken into account. Melamine is measured via a UV method after hydrolysis in dilute hydrochloric acid. The content of phenol and of the total formaldehyde in PF resins can only be determined by NMR. Residual monomers such as free formaldehyde, unreacted urea, and residual phenol or methanol as a residual product of formaldehyde production can be determined by various methods, e.g., free phenol via high performance liquid chromatography (HPLC). C. Structural Components Using dierent spectroscopic methods such as infrared (IR), 1H-NMR, 13C-NMR, or 15 N-NMR, analysis of the adhesive structural compounds enables a deep insight into the structural composition of resins. These results are the basis for correlations of resin structural composition with their molar composition, their preparation procedure, and the properties of the panels produced and hence to development and production of tailormade resins. Extensive information is available on the basic nature of resins and on the content of the various structural elements, including, e.g., data concerning the type of bridges between the monomers or the degree of branching. D. Molecular Weight Distribution and Molar Mass Averages The molecular weight distribution (MWD) can be determined by means of GPC [or size exclusion chromatography, (SEC)]. This method divides the molecules according to their hydrodynamic volume, which is proportional to their molar mass. The most important consideration in the chromatography of formaldehyde condensation resins is the poor solubility of the resins in most solvents usually used in GPC and hence the proper choice of the solvent and the mobile phase. This choice inuences the solubility of the resin, the behavior of the chromatographic columns, and the eectiveness of detection. For lower molar mass PF resins, tetrahydrofuran (THF) is a suitable solvent [128], while for higher molar mass phenolics and for MF resins, dimethylformamide (DMF) can be recommended, sometimes modied e.g., by addition of small amounts of ammonium formate or other salts such as LiCl [128,340]. UF resins are only soluble in DMF (with some undissolved higher molar mass portions) and dimethylsulfoxide (DMSO). DMSO shortens the lifetime of the chromatographic columns and causes problems with high pressures because of its higher viscosity in comparison to other organic solvents and low refractive index increments [341]. The high reactivity of the functional groups of the resins
Copyright 2003 by Taylor & Francis Group, LLC

additionally requires the use of the correct solvent and mobile phase, especially concerning sample preparation, in order to obtain a satisfactory reproducibility of the results. Another problem with GPC of condensation resins is the calibration of the columns. Because in the oligomeric and polymeric regions of the resins no compounds with a special and singular molar mass and a clear molecular structure are available, similar or chemically related substances have to be used as calibration standards. However, dierences in the hydrodynamic volumes even at the same molar mass cannot be excluded totally. This uncertain calibration of the columns also induces a great uncertainty in the calculation of molar mass averages on the basis of the chromatograms obtained. Molar mass distributions of UF resins have been reported by several authors [22,125,340345], as have mass distributions of MUF resins [71,346348]. The molecular characterization of PF resins can also be performed without any major problems by GPC [128,134,139,349,350]. Due to newer GPC methods, modication of the PF resin before the analysis is no longer necessary. Figure 3 shows chromatograms of two PF resins, one with a distinct high molecular weight portion, and the other with rather lower molar masses [128]. The averages of the molar mass can be (i) calculated from the gel chromatograms, taking into consideration the above-mentioned problems with the calibration of the columns, and (ii) measured by (a) vapor pressure osmometry for the number average molar mass (UF resins [19,22,341,344], MF resins [351], PF resins [121,352,353]) and (b) light scattering for the weight average molar mass (UF resins [19,22,341], PF resins [354]). The weight average molar mass at each elution volume can also be monitored directly during each GPC run using GPCLALLS. If the weight average molar mass in this case is determined directly in the eluent by light scattering, no standard calibration of

Figure 3 GPC plot of two PF resins: ( ) PF resin with a distinct high molecular weight portion; (- - - -) PF resin with rather low molar masses. Column set: Merck HIBAR LiChrogel PS1 PS4 PS20 PS400. Solvent and mobile phase: THF. Detection: UVVIS, 254 and 280 nm, respectively. Concentration of samples: 1 mg/ml. Flow rate: 0.5 ml/min. (After Ref. 128.)
Copyright 2003 by Taylor & Francis Group, LLC

Figure 4 GPC coupled with low angle laser light scattering (GPCLALLS) of a UF resin: e(V) concentration signal; E(V) normalized response of the LALLS detector; log Mw(V) E(V)/e(V) measured weight average molar mass as a function of the elution volume V. Column set Varian Toyo Soda TSK G4000 H8 G3000 H8 G2000 H8 G1000 H8. Solvent and mobile phase: 0.01 m solution of LiBr in DMF. Temperature: 40 C, ow rate: 1.1 ml/min, concentration of samples: 1015 mg/ml. (After ref. 18.)

the column is necessary (GPCLALLS). The eluent with the dissolved molecules passes a light scattering cell and the weight average molar mass is measured directly during each chromatographic run. However, this method is laborious and, therefore, described only in a few cases in the literature (UF [18] Fig. 4; PF [109,142,355,356]). During each run two curves are obtained: one is the concentration peak, and the other the light scattering peak, which is directly related to the actual molar mass average in the detection cell at each moment. Using these two curves, an individual calibration curve can be derived for each run. However, it must be taken into consideration that the light scattering signal can only be evaluated in the higher molar mass region and, therefore, the calibration curve is valid with sucient accuracy only in this part of the chromatogram. E. Monitoring of Gelling and Hardening

During gelling and hardening of the condensation resins in the hot press one can distinguish between the chemical advancement of the condensation reaction during curing of the thermosetting resin (build up of the three-dimensional network) and the progressive development of the mechanical strength of the joint (increase in cohesive bond strength). The two quantities do not progress at the same rate. The test methods that are used to follow the progression of the hardening of the resin are shown in Table 20. The extent of chemical curing can be monitored using DTA and DSC. The exothermic behavior of the curing process is then measured as a temperature dierence or directly as heat ow. Figure 5 shows a DSC plot of a PF resin [2]. The DSC run was done with pressure sealed capsules at a heating rate of 10 C/min.
Copyright 2003 by Taylor & Francis Group, LLC

Table 20

Test Methods Used to Follow Building Up of Bonding Strength Description Measures the dierence in temperature between two cells, these two cells are heated at a certain heating rate; one ofthe two cells contains the sample under investigation. Uses a similar type of instrument as DTA, but measures directly the heat ow of the exothermic and endothermic reactions occurring. The data obtained that are of interest are: shape of the curve, temperatures of the onset and the top of an exothermic or an endothermic peak, slope of the upcurve, width of the peak. DMA uses a small sheet of glass ber mats as a substrate, which is impregnated with the resin. This sample then undergoes periodic oscillations, at the same time the sample is heated following a special temperature program. The curing of the resin leads to an increase in the strength of the sample which then can be correlated with the increase of the cohesive bonding strength. Similar to DMA but follows the adhesive hardening in situ on the real wood substrate (rather than on glass ber). Thin wood strips are used to sandwich a liquid glue line which is then hardened. The curing of the resin leads to an increase in the strength of the sample which can then be correlated with the increase of the cohesive bonding strength as well as with the internal bond strength of wood particleboard using the same adhesive. It has been used both at constant heating rate and in isothermal mode. The damping behavior of the torsion of a glass ber probe impregnated with the resin is characteristic for the increase of stiness. The ABES consists of a small press and a tiny testing machine in a single unit. It enables bonds to be formed under highly controlled conditions; the joints that contain the bonds which are to be measured are pressed against heated blocks for a certain time, cooled within a few seconds, and pulled immediately thereafter in shear mode. Repetition of this procedure at dierent curing times and temperatures yields the points (a point for each specimen) of a near-isothermal strength development curve. References 357359

Test method Dierential thermal analysis (DTA) Dierential scanning calorimetry (DSC)

15,360363

Dierential mechanical analysis (DMA)

364367

Thermomechanical analysis (TMA)

368377

Torsional braid analysis (TBA) Automatic Bonding Evaluation System (ABES)

169,171,378

96,97

Copyright 2003 by Taylor & Francis Group, LLC

Figure 5 DSC plot of a PF resin. Pressure sealed capsules, heating rate 10 C/min. (From Ref. 2.)

During the curing of the resin the cohesive bonding strength develops step by step. Monitoring the eective strength increase (dened as the degree of mechanical curing) enables conclusions to be drawn about the suitability or not of a resin for a certain application. The best methods to use for this purpose are DMA (Fig. 6), TMA (Fig. 7) and ABES [96] (Fig. 1). In the TMA plot in Fig. 7 it is possible to note the interactive nature of the substrate on the curing of the PF adhesive. For example, the modulus of elasticity (MOE) increase curve shows two sections (and a two peak rst derivative curve). This indicates formation of entanglement networks of the resin in wood which is not possible on noninteractive substrates such as glass as in Fig. 6. Of course DMA and TMA give equally good results when used on the same wood substrate [379,380]. The ABES technique is also linearly correlated with TMA and DMA results as has been demonstrated by the linear relationship that has been found for both MUF and tanninformaldehyde adhesives in the results of TMA and ABES [381]. In board manufacturing, when the press opens, a certain level of mechanical hardening and with this a certain bond strength is necessary to withstand the internal steam pressure in the pressed board. The full chemical curing, however, can be attained outside of the press during hot stacking. Advanced formation of the bond strength already at the same degree of chemical curing will enable shorter press times and will, therefore, increase the production capacity and reduce production costs. Plotting the chemical and mechanical degrees of curing in an xy diagram shows the dierent hardening behaviors of various resins; such a correlation plot of the degree of chemical cure (e.g., measured by DSC) and the increase of mechanical strength (e.g., measured by TMA, DMA, or ABES) can be regarded as a ngerprint of the curing behavior of a resin [154].
Copyright 2003 by Taylor & Francis Group, LLC

Figure 6

DMA plot of a PF resin on glass ber. Heating rate 10 C/min. (After Ref. 2.)

Figure 7 TMA plot of the curing of a PF resin on beech wood. Heating rate 10 C/min. Numbers in the gure are temperatures in  C. (After ref. 370.) (*) MOE curve; (4) rst derivative curve.

Copyright 2003 by Taylor & Francis Group, LLC

XI.

WOOD AS AN INFLUENTIAL PARAMETER IN WOOD GLUING

The properties of wood-based panels are determined, in principle, by three parameters: wood, adhesive, and processing conditions. Only if all of these three parameters are correct and well balanced in the wood bonding process, can proper bonding results be achieved. The inuence of the rst parameter, wood, involves several factors. Bonded wood often is described as a chain of several links: wood (substance), wood surface, interface between wood and adhesive, surface of the glue line (boundary layer), and glue line itself. As is true for all such chains, the weakest link determines the strength of the chain, and in wood gluing this is in most cases the interface. The strength of an adhesive bond depends on various parameters: strength of the glue line and its behavior against stresses; inuence of humidity, wood moisture content, and wood preservatives added; wood properties, which can inuence the strength of the glue line and might cause internal stresses; and mechanical properties of the wood material. Hence wood, especially the wood surface and its interface with the bondline plays a crucial role in the quality of bonding and therefore the quality of the wood-based panels. Low or even no bonding strength can be caused by unfavorable properties of the wood surface, e.g., low wettability. A. Inuence of Wood Species on the Properties of Wood-Based Panels

In the wood-based panels industry a great variety of wood species are used as raw materials. The choice of the wood species used is often determined just by the availability and the price of the raw material. Furthermore, large amounts of wood residues from the primary wood processing industry (e.g., saw mill waste) as well as old (recycled) wood are used. It is more than a proverb to say that the quality of a wood-based panel has already been established, to a great extent, before the wood reaches the wood storage area of the panelproducing mill. The mills generally try to maintain as constant in time as possible the composition of the wood species mix as well as the mix of wood origins and preparation modes for a certain board type. For various board types, dierent wood mixes (species, shape and size of the particles) are used. This is rather based on practical and empirical long term experience and often not on any reasoned thinking. Economic reasons (availability of special wood, price) can also play an important role in the choices made. Many papers deal with special wood species in the production of wood-based panels, but the total knowledge available on this subject is not really satisfactory. Neusser and coworkers [382,383] are two of the rare examples in the literature giving a broader overview on this aspect: they have compared 18 dierent Austrian wood species by producing and testing laboratory particleboards. The test results obtained allowed adjustments for properties and density of laboratory boards. The best results were found for ash, followed by white beech and oak. However, these results may not be valid for all types of wood and all types of boards. B. Wood Particle Size and Shape Before Pressing

The strength of a bond in a wood panel increases with the value of wood density for the range of approximately 0.7 to 0.8 g/cm3. Above this density a decrease of the bond
Copyright 2003 by Taylor & Francis Group, LLC

strength occurs. The performance and properties of wood-based panels are strongly inuenced by the properties of the wood used. Thus, wood anisotropy as well as its heterogeneous nature, the variability of its properties, and its hygroscopicity have to be taken into account in all bonding processes. Equally, the orientation of the wood bers and the grain angle in bonding solid wood have to be considered. Particles as raw material for particleboards show a great variety in wood species, origin, method of preparation, age, and especially size and shape. If wood is ground into particles, a mixture of particles of very dierent sizes and shapes is always obtained. Particles can be described in a simple way as squared at pieces with certain values for length l (mm), width b (mm), thickness d (mm), and slenderness ratio s l/d. The volume of a particle is then given as V lbd mm3 Considering particles with l ) d, the eective gluing surface area is F 2lb mm2 The area form factor [384] can be considered as measure of the eective gluing surface area based on the volume. It is inversely proportional to the thickness of the particles: F 2 2s V d l The inuence of particle size and shape on mechanical and hygroscopic properties of boards is well described in several papers in the literature [385390]. The central statement of these papers is an increase of bending strength, and compression and tension strength in the board plane, but a decrease of internal bond strength with increasing particle length. In particleboards the particles overlap, and thus the overlapping areas must be large enough to guarantee the transmission of the wood strength to the strength of the whole assembly. C. Chemical Composition of Wood Extractives contained in wood can inuence the gluing process in the physical as well as chemical sense. Several authors [391393] have indicated that the chemical composition of a wood surface after processing might be dierent due to the concentration on it of polar and apolar substances coming from the wood itself. Even the ber direction of the wood surface (longitudinal, radial, tangential) can inuence this composition. Extractives soluble in water or steam can migrate during the drying process to the wood surface and can decrease its wettability. In particular fatty substances and waxes might cover the wood surface. As a consequence of this, chemical weak boundary layers (CWBLs) are formed [394,395]. A chemical-induced eect can also occur if the wood extractives have a strong acidic or alkaline behavior. This might cause acceleration or retardation of the hardening process of the adhesives based on polycondensation resins. Dierent wood species can show great dierences in pH as well as in the buering capacity. Even within a single wood species dierences might occur due to seasonal variations, position of origin within the tree log, pH of the soil, age of the tree, time span after cutting, and drying and processing parameters.
Copyright 2003 by Taylor & Francis Group, LLC

D. Wood Surfaces The wood surface is a complex and heterogeneous mixture of polymeric substances such as cellulose, hemicellulose, and lignin. It is also inuenced by factors such as polymer morphology, wood extractives, and processing parameters. During the processing of wood and the generation of new surfaces, damage to the wood material and to the surface can occur, which might cause low quality bonding and low bond strengths. This often shows as low percent wood failure or only as a thin ber layer. The reason for this can be a mechanical destruction of the uppermost wood layer, usually described as a mechanical weak boundary layer (MWBL) [396398]. This layer consists of damaged wood cells caused by processing. A fracture of a bond at the interface between the wood and the adhesive can be caused by a cohesive fracture of such a weak boundary layer [399] or by a real adhesion failure at the interface [396,400]. 1. Contact Angles of Wood Surfaces A precondition for the gluing of two wood surfaces is the wetting of these surfaces by the liquid or liquied adhesive. Wetting here includes the value for the contact angle, the spreading of the liquid on the surface, and the partial penetration of the liquid into the porous adherend. Good wetting enables the creation of high adhesion forces between the wood surface and the adhesive. However, direct correlations between the contact angle and the bonding strength achieved are rather rare [401] or seem not to be universal [402]. Low contact angles ( < 45 ) indicate good wetting behavior. Contact angles greater than 90 lead to incomplete wetting, which might cause low bond strengths. The main parameters that inuence the surface tension of the adhesive, when on the substrate, and therefore the possible bond strengths are: wood species [2426,403] roughness of the surface [404406] cutting direction (radial/tangential) [2426] earlywood, latewood [2426,407,408] direction of the spreading of the droplet during measurement of the contact angle (along or lateral to the direction of the bers) [409] wood moisture content [410412] ber angle [413] age of the wood surface [414,415] pH of the wood surface [416418] type and amount of wood extractives [401,419,420] pretreatment of the surface, e.g., by extraction with various solvents [421] type of adhesive: UF resins [2426]; PF resins [406408] During the production of wood-based panels a certain portion of the adhesive penetrates into the wood surface. An overpenetration causes starved glue lines, whereas too low a penetration limits the contact surface between the wood and the adhesive; low penetration often is the consequence of a poor wetting behavior. 2. Modications of the Wood Surface

Modications of the wood surface can be implemented using various physical, mechanical, and chemical treatments. Chemical treatments are performed in particular to enhance dimensional stability of the panel, but also to improve physical and mechanical properties
Copyright 2003 by Taylor & Francis Group, LLC

or to yield a higher resistance against physical, chemical, and biological degradation. To render the wood substrate hydrophobic, e.g., by acetylation, decreases the number of hydrophilic sites [422]. The OH groups of the cellulose react with acetic anhydride forming an ester. The hygroscopicity of the wood substrate decreases, and hence swelling and shrinking of the panel can be lowered [423]. Use of acetylated bers for the production of MDF boards showed marked reduction in their thickness swelling [424,425]. It has also been reported that wood acetylation can yield reactions of the anhydride with the aromatic ring of the lignin, although the exact reaction paths are not known [426]. This chemical attack at the aromatic rings can yield some crosslinking of the constituents of the wood substrate and can, therefore, contribute to the improved wood dimensional stability. 3. Seasonal Variations of Wood Quality in the Wood Panelboard Industry Hanetho [427] has discussed the experiences of the particleboard industry regarding the inuence of seasonal variations of the wood quality. Some problems do occur using wood that has been harvested in the winter time and which goes into board production immediately. When these logs or chips have been stored for some time, these problems disappear. The contact angles of water and adhesive on wood are higher in the case of freshly harvested wood compared to stored chips. This means that the surface of the wood particles obtained from such a fresh wood is more hydrophobic, inuencing negatively the wetting and penetration and thus the substrate gluability. It has been determined that the reason for the lower wettability of freshly harvested wood is the higher content of wood extractives. These results, however, must not be confused with the better wettability of a freshly prepared surface, independently of whether it is freshly harvested or stored wood. Hydrophobic wood extractives and components oxidize or polymerize during storage after harvesting, as also can be seen from their lower extractibility [428]. Because of this eect the ability of wood extractives to migrate to a new surface is also reduced. Figure 8 shows this eect by plotting contact angles versus time after the

Figure 8 Contact angles of a UF resin on the surfaces of wood particles, as a function of the contact time, hence the time elapsed after the application of the droplet. The surfaces have been cut from a freshly harvested log and from a log stored for 3 months. (After ref. 428.) Water extracts from the particles made from freshly harvested wood have higher pH values, but lower buer capacities than the surfaces made from stored chips. The lower buer capacity might lead to prehardening if the usual amount of hardener is used, with a consequent decrease of the board strength.
Copyright 2003 by Taylor & Francis Group, LLC

application of the urea resin droplet onto the surfaces of freshly harvested wood and stored wood.

XII. A.

PROCESSING CONDITIONS DURING PRODUCTION AS PARAMETERS INFLUENCING WOOD GLUING Adhesive Consumption and Glue Spread in the Production of Particleboards

Several aspects regarding the proportion of adhesive to be used in the production of particleboards must be evaluated to obtain good results: proportion of adhesive on individual particles proportion of adhesive in particle mixtures and fractions proportion of adhesive in the total particle mix distribution of the adhesive on the surface of the particles, and proportion of the particles surface area covered with adhesive. The resin load content on wood as a measure of the consumption of adhesive is one of the more important parameters to consider during the production of particleboards. From a technological standpoint a certain minimum amount of resin is necessary to obtain the desired properties of the boards resulting in sucient bonding of the individual particles. However, an excessive resin load imparts some technological disadvantages, such as high moisture content and hence possible problems with high vapor pressure during hot pressing. Furthermore, for economical reasons, the consumption of adhesive should be as low as possible as the resin contributes considerably to the costs of the nished boards. The resin load, however, is only an overall average on the total mixture of particles, without considering dierences in particle size distribution and the shape of the individual particles. Moreover, the resin load gives no direct indication of the area-specic consumption of the adhesive, which is the amount of resin solids content based on the surface area of the particles. The expression resin-robbing by the nes is well known and describes the exceedingly high consumption of adhesive based on mass of particles owing to the great surface area of the ne particles [429,430]. The resin load on wood chips can be described in the following two ways: mass resin load (percent or grams of resin solids content per 100 g dry particles) and surface-specic resin load (grams of resin solids content per square meter of surface area). If one of these two terms is known, the other can be calculated assuming a uniform distribution of the resin on the particle surfaces and estimating the total surface area of the particles. In the production of particleboards mixtures of particles are always used as raw material, and thus the particles dier in size and shape. A size grading of the particles can be performed by sieving, where two of the three dimensions of the particle must be smaller than the standard measure of the actual sieve mesh to be passed. An exact sieving of the particles according to their size, therefore, is only possible for particles of rather similar shapes. Particles can dier widely in shape. A simplication to describe their shape is to assume that they are squared, at with length l, width b, and thickness d for medium and coarse particles and rather cubic for the nes. Since the sieve mesh is usually graduated according to a logarithmic scale, for the theoretical calculations of the particle size
Copyright 2003 by Taylor & Francis Group, LLC

Figure 9 Example of a particle size distribution, the calculated mass resin load (gluing factor), and the distribution of the resin solids content. The overall adhesive resin consumption was assumed to be 8% resin solids content/dry wood. (After ref. 429.)

distribution this was also assumed to be logarithmic and similar to a gaussian distribution. Distributions on an industrial scale might dier from this model. Each particle fraction has a certain relation to its resin load according to the size of the particles. Because of the great surface area of the ne particles their resin load increases strongly (linearly with the term d1). Even if there is only a small proportion of a mass fraction of very ne particles in the mixture, the high consumption of resin solids content of this fraction has a negative impact on the resin load of the coarse particles. Figure 9 shows an example of a particle size distribution with the calculated mass resin loads and the distribution of the resin solids content on the dierent fractions of the particle size distribution. Particle length was assumed to vary from 25 mm for the coarsest particles to 0.6 mm for wood dust, according to experience with industrial particle mixtures. Because of the reasons discussed above, usually core layers and face layers are glued separately. In the core layer rather coarse particles predominate and in the face layer rather ne particles predominate. This separate gluing enables the use of dierent compositions of the glue mixes (e.g., dierent addition of water and hardener) and dierent resin loads (gluing factors) for the two layer types. An example of separate gluing is shown in Fig. 10, with separate gluing of the core layer CL (6.5% mass gluing factor) and in the face layer FL (11.0% mass gluing factor). The mass ratio of the layers CL:FL is 60:40. Figure 10 shows the particle size distributions and the mass gluing factors of the individual particle size fractions for this example of separate gluing. Samples of industrial core layer and face layer particles, before and after gluing, can be fractionated by sieving, and thus sampling has to be done at the same time before and after blending. In the case of aminoplastic adhesives each particle fraction, glued or not, can be investigated for its nitrogen content. By knowing (i) the content of nitrogen as well as the resin solids content in the glue mix and (ii) the moisture content of the particles glued or not in the various fractions, the mass gluing factor of each glued particle size fraction can be calculated. Figure 11 shows the results of one such calculation for glued core layer and face layer particles. Even if the absolute values may dier from the calculated ones, the resin load (by weight) and the particle size show that the same shape of distribution curve is obtained.
Copyright 2003 by Taylor & Francis Group, LLC

Figure 10 Particle size distribution and mass gluing factor of the individual particle size fractions for the separate gluing example CLFL. The resin consumption was assumed to be 6.5% resin solids content/dry wood in the core layer (CL) and 11.0% in the face layer (FL). The mass proportions are CL:FL 60:40. (After ref. 429.)

Figure 11 Fractionated mass gluing factors of industrially glued core layer particles. The mass gluing factor during blending was 9.5% resin solids/dry particles. (After ref. 429.)

Assuming that the gluing of particles of dierent sizes is performed randomly with their surface area as the decisive parameter, for various homogeneous particle size fractions and for dierent particle size mixtures the theoretical mass gluing factors and the distribution of the resin solids content can be calculated and correlated with the same values obtained experimentally, by analysis. There are some indications [431433], however, that glue distribution is not exclusively inuenced by the surface area of the particles, but has a certain preference for coarser particles. This may be due to the eectiveness of the adhesive application, thus to the separation and distribution of resin droplets, or to the mixing action in the blender after application of the resin on the wood particles (wiping eect). The concept that the particle surface area exclusively inuences gluing is quite clearly invalid, if glue droplets and the surface to be glued have similar size. Meinecke and Klauditz [431] mentioned diameters of glue droplets of 8 to 110 mm, depending on the type of spraying and Lehmann [434] mentioned up to 200 mm. The latter values are of the same order of magnitude as the size of the nest particles used for the calculations above. Besides the surface area of the particles several other parameters also have some inuence on the necessary resin consumption, e.g., type of boards, thickness of the sanding
Copyright 2003 by Taylor & Francis Group, LLC

zone, type and capacity of the blenders, separation and spraying of the resin (depending on if only the wiping/spreading eect occurs during blending or if instead spraying of the resin is used), shape of the particles for the same particle sizes, dependence of the slenderness ratio on particle length, concentration and viscosity of the glue resin, or a partial size degradation of the coarser particles in the blender. New strategies in blending take into account the reality of the higher resin consumption by the ner particles, e.g., by removing the dust and the nest particles from the particle mix before blending. Also an exact screening and classifying of the particles before blending can improve the distribution of the resin on the particle surfaces and can help to spare some resin. A lower consumption of resin not only means lower costs for the raw materials, but also helps to avoid various technological disadvantages. With the resin, water is also applied to the particles; as long as this amount of water is low enough, especially in the core layer, no problem should occur with a too high vapor pressure during hot pressing. Often, however, the moisture content of the glued core particles is too high, due to an excessive gluing factor. The high vapor pressure in the board at the end of the press cycle tends to expand the fresh board; if venting is not done very carefully, blistering of the boards at the end of the continuous press or after the opening of the press might occur. Additionally, the heat transfer by the steam shock can be delayed if the vapor pressure dierence betweeen the face layer and the core layer is small. If the moisture content of the glued core layer particles is high, the moisture in the glued face layer particles must be reduced. Also spraying water onto the belt before the forming station and onto the surface of the formed mat cannot be done due to the problems with the too high moisture content in the mat and hence with the too high vapor pressure. Gluing of particles is usually done in quickly rotating blenders by spraying the resin mix into the blender. Due to the rotation of the blender a partial degradation in the size of the particles can occur. While blending OSB strands this degradation must be avoided; this is done by using slowly rotating big blender drums with a diameter of approximately 3 m. The liquid adhesive is distributed by several atomizers in this blender drum. Gluing of bers in MDF production is usually done in the so-called blowline between the rener and the dryer. The advantage of this method is that it avoids resin spots at the surface of the board. The disadvantage, however, is the fact that the resin passes the dryer and can suer part precuring. This causes some loss of usable resin (approximately 0.5 to 2% in absolute gures); therefore the glue consumption in blowline blending is higher than in the mechanical blending. Due to this fact mechanical blenders have lately been installed again in a few factories. The theory of turbulent ow blowlinegluing is not yet clearly dened [435,436]. However, some equations attempting to describe it have been recently presented [436]. B. Wood Moisture Content

The wood moisture content inuences several important processes such as wetting, ow of the adhesive, penetration into the wood surface, and hardening of the adhesive in the gluing and production of wood-based panels. In bonding solid wood usually a wood moisture content of 6 to 14% is seen as optimal. Lower wood moisture contents can cause a quick dryout of the glue spread due to a strong absorption of the water into the wood surface as well as wetting problems. High moisture contents can lead to a high ow and an enhanced penetration into the wood, causing starved glue lines.
Copyright 2003 by Taylor & Francis Group, LLC

Additionally a high steam pressure can be generated which might give problems of blistering when the press opens or at the end of the continuous press. Also the hardening of a condensation resin might be retarded or even hindered. During the hot press cycle of the particleboard or MDF production, quick changes of temperature, moisture content, and steam pressure occur. The gradients of temperature and moisture content determine signicantly the hardening rate of the resin and hence the board properties. These gradients together with the mechanical pressure applied to densify the mat are decisive for generating the density prole and hence for the application properties and performance of the boards. The higher the moisture content of the glued face layer particles, the steeper the moisture gradient between the surface and the core of the mat and the quicker the heating up of the mat occurs. In the ber mat in MDF production no dierences are seen in the moisture content of the outer layer and the inner layer due to the temperature applied to the mat, nevertheless a vapor pressure gradient occurs. The moisture content of the glued particles is the sum of the wood moisture content and the water that is part of the applied glue mix. Therefore, the moisture content of the glued particles mainly depends on the gluing factor. Usual moisture contents of glued particles are: (a) for UF, 6.58.5% in the core layer and 1013% in the face layer; (b) for PF, 1114% in the core layer and 1418% in the face layer. The optimal moisture content of the glued and dried MDF bers in the mat before the press is in the region of 911%. The higher the moisture content of particles, the easier the face layer can be densied at the start of the press cycle; this leads to a lower density in the core layer. Blistering at the end of the press cycle or at the end of the continuous press occurs if the steam pressure within the fresh, and still hot, board exceeds the internal bond strength of the board. It should be noted that the bond strengths at higher temperatures are always lower than after cooling the board. If blistering occurs using resins with low formaldehyde content, press time should be shortened instead of prolonged, because a longer press time would not increase the bond strength but certainly would increase the steam pressure in the board. Careful venting as well as decreasing the moisture content of the glued particles and reducing the press temperature will help.

C. Press Cycle During the hot press cycle the hardening of the resin and possible reactions of the adhesive with the wood substance take place. The inuential parameters are especially the press temperature and the moisture content in the mat. Additional parameters are the wood density, porosity, swelling and shrinking behavior of the wood, structure at the surface, and wetting behavior. During the press cycle several processes take place: transport of heat and moisture densication, increasing internal stresses, followed by relaxation processes adhesion between the particles or bers increase of the bond strength in the glue line (cohesion). Models describing what occurs in a panel during hot pressing have been published [437443]. These take into consideration various conditions occurring during the hot press cycle such as heat transfer, temperature gradients, moisture content, steam pressure, bond strengths, and presence or absence of postcuring [437443].
Copyright 2003 by Taylor & Francis Group, LLC

Table 21

Press Strategy for Production of Particleboards

Dierent particle structures: coarser in the core, ner in the face layer. Press temperature: As high as possible, to enable a quick heating up of the core layer due to an optimal steam shock eect. In continuous lines press temperatures decrease from the entrance to the outlet of the press. In the last zone of the press even active cooling in a few cases is possible (decreasing steam pressure in the core layer). Moisture content of the glued particles: Core layer as dry as possible (ca. 67% in the case of UF resins), face layer as high as possible (1114%, depending on the proportion of the face layer in the board). Too high a moisture content can cause blistering. Spraying of water onto both surfaces in order to enhance the steam shock, amount ca. 2040 g/m2. Press pressure prole: The variation of pressure during hot pressing can follow dierent sequences. Quick densication with pressure maximum to enable a high density of the face layer and hence high modulus of elasticity (MOE). Sometimes a second densication step is used.

Table 22

Press Strategy for Production of MDF

Despite the uniform ber material, a certain density prole is created due to the action of heat and compression. Two-step pressure prole with quick densication at the start of the hot press cycle and a second densication step for the inner layer. Uniform moisture content of the glued and dried bers across the thickness of the mat. Higher moisture content in the outer layer would require a three-layer mat or spraying of water.

Tables 21 and 22 summarize the usual press strategies for the production of particleboards and MDF. The warming up of the mat is performed by the so-called steam shock eect [442447]. The precondition for this is the high permeability to steam and gases of the particle or ber mat [442,443,448,449]. High moisture contents of the face layers and spraying of water on the surface layers sustain this eect. The press temperature inuences the possible press time and by this the capacity of the production line. The minimum press time has to guarantee that the bond strength of the still hot board can withstand the internal steam pressure as well as the elastic springback in board thickness at press opening.

XIII.

CONCLUSIONS

Wood is a very complex material. Wood adhesives technology is an advanced science which blends the technology of adhesive preparation and formulation with a multitude of advanced application technologies to dierent wood products. In many elds other than wood, good bonding depends mainly on the use of a good adhesive. The situation is not as straightforward in wood gluing: in general one can obtain excellent wood panels when using a decidedly poor adhesive if the parameters governing the technology of manufacture of the wood product are well mastered. This indicates the extent to which a high level application technology can play a predominant role in this eld. This is not
Copyright 2003 by Taylor & Francis Group, LLC

valid for all wood products. Of course, good results are better or easier to obtain if one uses an excellent adhesive. However, just the use of a good adhesive gives no assurance of good bonding in this eld. It is the essential interaction of the equally important adhesive and its application technology that this chapter has tried to describe. It is exactly this interaction that is so important in a eld that comprises more than 60% by volume of all the adhesives used today in the world for any application. Without mastering this interaction between adhesive technology and wood product manufacturing technology there cannot be wood bonding of any consequence.

REFERENCES
1. 2. 3. 4. M. Dunky, in Duroplaste (Kunststo-Handbuch Bd. 10) (W. Woebeken, ed.), Carl Hanser, Verlag, Munich, 1988, pp. 593614. M. Dunky, in Polymeric Materials Encyclopedia (J.C. Salamone, ed.), CRC Press, Boca Raton, FL, 1996, Vol. 11. M. Dunky, Int. J. Adhesion Adhesives 18: 95107 (1998). M. Dunky, in (M. Dunky, A. Pizzi and M. Van Leemput, eds.) State of the Art-Report, COST-Action E13, part I (Working Group 1, Adhesives), European Commission, Brussels, Belgium, 2002. K. Lederer, in Polymere Werkstoe, Vol. III (H. Batzer, ed.), Thieme, Stuttgart, 1984, pp. S.95291. B. Meyer, UreaFormaldehyde Resins, Addison-Wesley, London, 1979. A. Pizzi, ed., Wood Adhesives: Chemistry and Technology, Marcel Dekker, New York, 1983, pp. 59104. A. Pizzi, Advanced Wood Adhesives Technology, Marcel Dekker, New York, 1994, pp. 1966. A. Pizzi, in Handbook of Adhesive Technology (A. Pizzi and K.L. Mittal, eds.), 1st Edition, Marcel Dekker, New York, 1994, pp. 381392. V. Horn, G. Benndorf, and K. P. Radler, Plaste Kautsch. 25: 570575 (1978). M. G. Kim and L. W. Amos, Ind. Eng. Chem. Res. 29: 208212 (1990). C. Soulard, C. Kamoun, and A. Pizzi, J. Appl. Polym. Sci. 72: 277289 (1999). BASF AG, German patent DE 2,207,921 (1972). BASF AG, German patent DE 2,550,739 (1975). Y. Su, Qu. Ran, W. Wu, and X. Mao, Thermochim. Acta 253: 307316 (1995). Rohm & Haas Co., U.S. patent 2,605,253 (1950). Allied Chemical & Dye Co., U.S. patent 2,683,134 (1951). J. Billiani, K. Lederer, and M. Dunky, Angew. Makromol. Chem. 180: 199208 (1990). M. Dunky, K. Lederer, and E. Zimmer, Holzforsch. Holzverwert. 33: 6171 (1981). G. E. Myers, Forest Prod. J. 34(5): 3541 (1984). M. Dunky, Holzforsch. Holzverwert. 37: 7582 (1985). M. Dunky and K. Lederer, Angew. Makromol. Chem. 102: 199213 (1982). Ch. Huber and K. Lederer, J. Polym. Sci., Polym. Lett. Ed. 18: 535540 (1980). M. Scheikl and M. Dunky, Holzforsch. Holzverwert. 48: 5557 (1996). M. Scheikl and M. Dunky, Holz. 54: 113117 (1996). M. Scheikl and M. Dunky, Holzforschung 52: 8994 (1998). M. Dunky, unpublished results, 1985. Y. Nakarai and T. Watanabe, Wood Industry 17: 464468 (1962). J. T. Rice, Forest Prod. J. 15: 107112 (1965). E. E. Ferg, M.Sc. Thesis, University of the Witwatersrand, Johannesburg, South Africa, 1992. H. Neusser and W. Schall, Holzforsch. Holzverwert. 24: 4550 (1972). J. S. Sodhi, Holz Roh. Werkst. 15: 9296 (1957). M. Dunky, unpublished results, 1997.

5. 6. 7. 8. 9. 10. 11. 12. 13. 14. 15. 16. 17. 18. 19. 20. 21. 22. 23. 24. 25. 26. 27. 28. 29. 30. 31. 32. 33.

Copyright 2003 by Taylor & Francis Group, LLC

34. 35. 36. 37. 38. 39. 40. 41. 42. 43. 44. 45. 46. 47. 48. 49. 50. 51. 52. 53. 54. 55. 56. 57. 58. 59. 60. 61. 62. 63. 64. 65. 66. 67. 68. 69. 70. 71. 72. 73. 74. 75. 76. 77. 78. 79.

VEB Leuna-Werke, German patent DE 2,655,327 (1976). BASF AG, European patent EP 1,596 (1978). Ch.-Y. Hse, R. L. Geimer, W. E. Hsu, and R. C. Tang, Forest Prod. J. 45(1): 5762 (1995). E. Kehr, G. Riehl, E. Hoferichter, E. Roael, and B. Dix, Holz Roh. Werkst. 52: 253260 (1994). A. Pizzi, J. Valenzuela, and C. Westermeyer, Holzforschung 47: 6972 (1993). C. Simon, B. George, and A. Pizzi, Holzforschung 56(3): 327334 (2002). H.-J. Deppe, Holz Roh. Werkst. 35: 295299 (1977). H.-J. Deppe and K. Ernst, Holz Roh. Werkst. 29: 4550 (1971). A. Tinkelenberg, H. W. Vassen, K. W. Suen, and P. G. J. Leusink, J. Adhesion 14: 219231 (1982). Methanol Chemie Nederland, European patent EP 25,245 (1980). M. Dunky, unpublished results, 1986. E. Roael, B. Dix, H. Miertzsch, T. Schwarz, E. Kehr, M. Scheithauer, and E. Hoferichter, Holz Roh. Werkst. 51: 197207 (1993). M. Higuchi, H. Shimokawa, and I. Sakata, Mokuzai Gakkaishi 25: 630635 (1979). G. E. Troughton and S. Chow, Holzforschung 29: 214217 (1975). A. Pizzi, J. Adhesion Sci. Technol. 4: 573578 (1990); 4: 589595 (1990). A. Pizzi, J. Adhesion Sci. Technol. 1: 191200 (1987). R. O. Ebewele, J. Appl. Polym. Sci. 58: 16891700 (1995). R. O. Ebewele, G. E. Myers, B. H. River, and J. A. Koutsky, J. Appl. Polym. Sci. 42: 29973012 (1991). R. O. Ebewele, B. H. River, G. E. Myers, and J. A. Koutsky, J. Appl. Polym. Sci. 43: 14831490 (1991). R. O. Ebewele, B. H. River, and G. E. Myers, J. Appl. Polym. Sci. 49: 229245 (1993). R. O. Ebewele, B. H. River, and G. E. Myers, J. Appl. Polym. Sci. 52: 689700 (1994). H. Yamaguchi, M. Higuchi, and I. Sakata, Mokuzai Gakkaishi 35: 199204 (1980). H. Yamaguchi, M. Higuchi, and I. Sakata, Mokuzai Gakkaishi 35: 801806 (1989). G. E. Myers, Wood Sci. 15: 127138 (1982). G. E. Myers and J. A. Koutsky, Holzforschung 44: 117126 (1990). G. E. Myers, Proc. Wood Adhesives 1985: Status and Needs, Madison, WI, 1985, 119156. G. E. Myers, Forest Prod. J. 33(4): 4957 (1983). M. Higuchi and I. Sakata, Mokuzai Gakkaishi 25: 496502 (1979). M. Higuchi, K. Kuwazuru, and I. Sakata, Mokuzai Gakkaishi 26: 310314 (1980). K. Ezaki, M. Higuchi, and I. Sakata, Mokuzai Kogyo 37: 225230 (1982). J. Dutkiewicz, J. Appl. Polym. Sci. 28: 33133320 (1983). Lentia GmbH, German patent DE 2,455,420 (1974) BASF AG, German patent DE 3,442,454 (1984). Methanol Chemie Nederland, European patent EP 62,389 (1982). Westinghouse Electric Corp., U.S. patent 4,123,579 (1978). Georgia-Pacic Resins, Inc., U.S. patent 5,681,917 (1996). T. A. Mercer and A. Pizzi, Holzforsch. Holzverwert. 46: 5154 (1994). R. Maylor, Proc. Wood Adhesives 1995, Portland, OR, 1995, pp. 115121. BASF AG, German patent DE 3,116,547 (1981). BASF AG, European patent EP 52,212 (1981). M. Prestilippo, A. Pizzi, H. Norback, and P. Lavisci, Holz Roh. Werkst. 54: 393398 (1996). C. Cremonini and A. Pizzi, Holzforsch. Holzverwert. 49: 1115 (1997). C. Cremonini and A. Pizzi, Holz Roh. Werkst. 57: 318 (1999). C. Kamoun and A. Pizzi. Holz Roh. Werkst. 56: 86 (1998). A. Weinstabl, W. H. Binder, H. Gruber, and W. Kantner, J. Appl. Polym. Sci. 81: 32313235 (2001). S. Chow and K. J. Pickles, Wood Sci. 9: 8083 (1976).

Copyright 2003 by Taylor & Francis Group, LLC

80. 81. 82. 83. 84. 85. 86. 87. 88. 89. 90. 91. 92. 93. 94. 95. 96. 97. 98. 99. 100. 101. 102. 103. 104. 105. 106. 107. 108. 109. 110. 111. 112. 113. 114. 115. 116. 117. 118. 119. 120. 121. 122. 123. 124.

H. Neusser and W. Schall, Holzforsch. Holzverwert 24: 108116 (1972). G. Lehmann, Proc. Klebstoe fur Holzwerkstoe und Faserformteile, Braunschweig, Germany, 1997. F. Wolf, Proc. First European Panel Products Symposium, Llandudno, Wales, 1997, pp. 243249. C. Cremonini, A. Pizzi, and P. Tekely, Holz Roh. Werkst 54: 8588 (1996). M. Prestilippo and A. Pizzi, Holz Roh. Werkst. 54: 272273 (1996). BASF AG, German patent DE 2,020,481 (1970). A. Huttermann and A. Haars, Peiderer AG, German patent DE 3,621,218 (1986). A. Haars, A. Huttermann, and A. Kharazipour, Peiderer AG, German patent DE 3,644,397 (1986). M. Higuchi, J.-K. Roh, S. Tajima, H. Irita, T. Honda, and I. Sakata, Proc. Adhesives and Bonded Wood Products, Forest Prod. Society, Madison, WI, 1994, pp. 429449. L. A. Panamgama and A. Pizzi, J. Appl. Polym. Sci. 58: 277281 (1995). D. Braun and W. Kraue, Angew. Makromol. Chem. 108: 141159 (1982). D. Braun and W. Kraue, Angew. Makromol. Chem. 118: 165182 (1983). D. Braun and H.-J. Ritzert, Angew. Makromol. Chem. 125: 926 (1984). D. Braun and H.-J. Ritzert, Angew. Makromol. Chem. 125: 2736 (1984). J. Vehlow, WKI-Report 22, Wilhelm Klauditz Institut, Braunschweig, 1990 M. Higuchi, S. Tohmura, and I. Sakata, Mokuzai Gakkaishi 40: 604611 (1994). P. E. Humphrey, Proc. Third Pacic Rim Bio-based Composites Symposium, Kyoto, Japan, 1996, pp. 366373. P. E. Humphrey, US patent USP 5,176,028 (1990). A. G. Krems Chemie, European patent EP 436,485 (1990). E. E. Ferg, A. Pizzi, and D. C. Levendis, J. Appl. Polym. Sci. 50: 907915 (1993). E. E. Ferg, A. Pizzi, and D. C. Levendis, Holzforsch Holzverwert. 45: 8892 (1993). T. A. Mercer and A. Pizzi, J. Appl. Polym. Sci. 61: 16971702 (1996). T. A. Mercer and A. Pizzi, J. Appl. Polym. Sci. 61: 16871695 (1996). L. A. Panamgama and A. Pizzi, J. Appl. Polym. Sci. 59: 20552068 (1996). N. J. L. Megson, Phenolic Resin Chemistry, Butterworth, London, 1958. H. G. Peer, Rec. Trav. Chim. 78: 851 (1959). H. G. Peer, Rec. Trav. Chim. 79: 825 (1960). E. Kumpinsky, Ind. Eng. Chem. Res. 33: 285291 (1994). C. M. Chen and J. T. Rice, Forest Prod. J. 26(6): 1723 (1976). L. Gollob, Ph.D. thesis, Oregon State University, Corvallis, OR, 1982. A. Pizzi, ed., Wood Adhesives: Chemistry and Technology, Marcel Dekker, New York, 1983, pp. 105176. R. Mueller, in Duroplaste (Kunststo-Handbuch Ed. 10), Carl Hanser Verlag, Munich, 1988, pp. 614629. T. Sellers, Jr., Plywood and AdhesiveTechnology, Marcel Dekker, New York, 1985. A. R. Walsh and A. G. Campbell, J. Appl. Polym. Sci. 32: 42914293 (1986). Monsanto, U.S. patent USP 3,342,776 (1967). Borden Chemical Company Ltd., U.S. patent USP 4,433,120 (1981). A. Pizzi, Advanced Wood Adhesives Technology, Marcel Dekker, New York, 1994, pp. 89148. C. Zhao, A. Pizzi, and S. Garnier, J. Appl. Polym. Sci. 74: 359378 (1999). P. R. Steiner, G. E. Troughton, and A. W. Andersen, Proc. Adhesives and Bonded Wood Products, Seattle, WA, 1991, pp. 205214. M. R. Clarke, P. R. Steiner, and A. W. Anderson, U.S. patent USP 4,824,896 (1988). S. Tohmura, M. Higuchi, and I. Sakata, Mokuzai Gakkaishi 38: 5966 (1992). S. Chow, P. R. Steiner, and G. E. Troughton, Wood Sci. 8: 343349 (1975). M. Duval, B. Bloch, and S. Kohn, J. Appl. Polym. Sci. 16: 15851602 (1972). S. So and A. Rudin, J. Appl. Polym. Sci. 41: 205232 (1990). E. R. Wagner and R. J. Gre, J. Polym Sci. A1 9: 21932207 (1971).

Copyright 2003 by Taylor & Francis Group, LLC

125. K. Oldoerp, Proc. Klebstoe fur Holzwerkstoe und Faserformteile, Braunschweig, Germany, 1997. 126. K. Oldoerp and H. Miertzsch, Holz Roh. Werkst. 55: 97102 (1997). 127. S. Ellis and P. R. Steiner, Proc. Wood Adhesives 1990, Madison, WI, 1990, pp. 7685. 128. G. Gobec, M. Dunky, T. Zich and K. Lederer, Angew. Makromol. Chem. 251: 171179 (1997). 129. M. G. Kim, L. W. Amos, and E. E. Barnes, ACS Div. Polym. Chem. Polym. Prepr. 24(2): 173174 (1983). 130. W. L-S. Nieh and T. Sellers, Jr., Forest Prod. J. 41(6): 4953 (1991). 131. E. Johnson, St. and F. A. Kamke, J. Adhesion 40: 4761 (1992). 132. E. Johnson, St. and F. A. Kamke, Wood Fiber Sci. 26: 259269 (1994). 133. R. A. Haupt and T. Sellers, Jr., Forest Prod. J. 44(2): 6973 (1994). 134. S. Ellis, Forest Prod. J. 43(2): 6668 (1993). 135. S. Ellis and P. R. Steiner, Forest Prod. J. 42: 814 (1992). 136. B. D. Park, B. Riedl, E. W. Hsu, and J. Shields, Holz Roh. Werkst. 56: 155161 (1998). 137. J. Perlac, Holztechnol. 5: 4548 (1964). 138. M. D. Peterson, Proc. Wood Adhesives 1985, Madison, WI, 1985, pp. 8297. 139. R. S. Stephens and N. P. Kutscha, Wood Fiber Sci. 19: 353361 (1987). 140. J. B. Wilson, G. L. Jay, and R. L. Krahmer, Adhesives Age 22: 2630 (1979). 141. J. B. Wilson and R. L. Krahmer, Proc. 12th Washington State University Int. Symposium on Particleboards, Pullmann, WA, 1978, pp. 305315. 142. L. Gollob, R. L. Krahmer, J. D. Wellons, and A. W. Christiansen, Forest Prod. J. 35: 4248 (1985). 143. R. H. Young, E. E. Barnes, R. W. Caster, and N. P. Kutscha, ACS Div. Polym. Chem. Polym. Prepr. 24(2): 199200 (1983). 144. W. Werner and O. Barber, Chromatographia 15: 101106 (1982). 145. R. A. Haupt and T. Sellers, Jr., Ind. Eng. Chem. Res. 33: 693697 (1994). 146. A. Pizzi and A. Stephanou, J. Appl. Polym. Sci. 49: 21572170 (1993). 147. A. Pizzi and A. Stephanou, Holzforschung 48: 3540 (1994). 148. X. Lu and A. Pizzi, Holz Roh. Werkst. 56: 339346 (1998). 149. A. Pizzi, B. Mtsweni, and W. Parsons, J. Appl. Polym. Sci. 52: 18471856 (1994). 150. S. So and A. Rudin, J. Polym. Sci. Polym. Lett. Ed. 23: 403407 (1985). 151. R. H. Young, Proc. Wood Adhesives 1985: Status and Needs, Madison, WI, 1985, pp. 267276. 152. R. G. Schmidt and C. E. Frazier, Wood Fiber Sci. 30: 250258 (1998). 153. R. G. Schmidt and C. E. Frazier, Int. J. Adhesion Adhesives 18: 139146 (1998). 154. R. L. Geimer, R. A. Follensbee, A. W. Christiansen, J. A. Koutsky, and G. E. Myers, Proc. 24th Washington State University Int. Particleboard/Composite Materials Symposium, Pullmann, WA, 1990, pp. 6583. 155. M. Lecourt, P. Humphrey, and A. Pizzi, Holz Roh. Werkst. 61(1): 75(2003). 156. A. Pizzi, R. Vosloo, F. A. Cameron, and E. Orovan, Holz Roh. Werkst. 44: 229234 (1986). 157. A. W. Christiansen, Forest Prod. J. 35: 4754 (1985). 158. A. Pizzi, R. Garcia, and S. Wang, J. Appl. Polym. Sci. 66: 255266 (1997). 159. B. Riedl and B.-D. Park, Proc. Forest Products Society Annual Meeting, Merida, Mexico, 1998, pp. 115121. 160. S. Tohmura and M. Higuchi, Mokuzai Gakkaishi, 41: 11091114 (1995). 161. S. Tohmura, J. Wood Sci. 44: 211216 (1998). 162. T. Gramstad and J. Sandstroem, Spectrochim. Acta 25A: 3135 (1969). 163. M. G. Kim, L. W. Amos, and E. E. Barnes, Ind. Eng. Chem. Res. 29: 20322037 (1990). 164. E. Scopelitis and A. Pizzi, J. Appl. Polym. Sci. 48: 21352146 (1993). 165. K. Oldoerp and R. Marutzky, Holz Roh. Werkst. 56: 7577 (1998). 166. B. Tomita and Ch.-Y. Hse, J. Polym. Sci. Part A, Polym. Chem. 30: 16151624 (1992). 167. B. Tomita and Ch.-Y. Hse, Mokuzai Gakkaishi, 39: 12761284 (1993). 168. B. Tomita, M. Ohyama, and Ch.-H. Hse, Holzforschung, 48: 522526 (1994).
Copyright 2003 by Taylor & Francis Group, LLC

169. 170. 171. 172. 173. 174. 175. 176. 177. 178. 179. 180. 181. 182. 183. 184. 185. 186. 187. 188. 189. 190. 191. 192. 193. 194. 195.

196. 197. 198. 199. 200. 201. 202. 203. 204. 205. 206. 207.

B. Tomita, M. Ohyama, A. Itoh, K. Doi, and Ch.-H. Hse, Mokuzai Gakkaishi, 40: 170175 (1994). B. Tomita and Ch.-Y. Hse, Proc. Adhesives and Bonded Wood Products, Seattle, WA, 1991, pp. 462479. M. Ohyama, B. Tomita, and C. Y. Hse, Holzforschung 49: 8791 (1995). C. Zhao, A. Pizzi, A. Kuhn, and S. Garnier, J. Appl. Polym. Sci. 77: 249259 (1999). A. Pizzi, A. Stephanou, I. Antunes, and G. de Beer, J. Appl. Polym. Sci. 50: 22012207 (1993). Y. Yoshida, B. Tomita, and Ch.-Y. Hse, Mokuzai Gakkaishi, 41: 652658 (1995). B. Tomita and Ch.-Y. Hse, Int. J. Adhesion Adhesives 18: 6979 (1998). A. Pizzi, ed., Wood Adhesives: Chemistry and Technology, Marcel Dekker, New York, 1983, pp. 177247. A. Pizzi, Advanced Wood Adhesives Technology, Marcel Dekker, New York, 1994, pp. 148225. E. Kulvijk, Adhesives Age 20: 3334 (1977). C.-M. Chen, Forest Prod. J. 32: 3540 (1982). C.-M. Chen, Forest Prod. J. 32(11/12): 1418 (1982). C.-M. Chen, Holzforschung 36: 109116 (1982). B. Dix and R. Marutzky, Adhaesion 26(12): 410 (1982). R. M. Drilje, FAO-Report World Consultation on Wood Based Panels, New Delhi, India, 1975. R. Long, Holz Roh. Werkst. 49: 485487 (1991). L. Suomi-Lindberg, Paperi ja Puu 2: 6569 (1985). R. F. Buchholz, G. A. Doering, and C. A. Whittemore, Proc. Wood Adhesives 1995, Portland, OR, 1995, pp. 241246. C. M Chen, Holzforchung 49: 153157 (1995). B. Danielson and R. Simonson, J. Adhesion Sci. Technol. 12: 923939, 941946 (1998). A. Trosa and A. Pizzi, Holz Roh. Werkst. 56: 229233 (1998). L. Zhao, B. F. Griggs, C.-L. Chen, J. S. Gratzl, and C.-Y.Hse, J. Wood Chem. Technol. 14: 127145 (1994). A. Pizzi and T. Walton, Holzforschung 46: 541547 (1992). A. Pizzi, J. Valenzuela, and C. Westermeyer, Holzforschung 47: 6972 (1993). C. Simon, B. George, and A. Pizzi, J. Appl. Polym. Sci., 86: 3681 (2002). T. Adcock, M. P. Wolcott, and S. M. Peyer, Proc. 4th European Panel Products Symposium, Llandudno, Wales, 1999, pp. 6776. K. W. Haider, J. W. Rosthauser, and T. R. Miller, Proc. Forest Products Society Annual General Meeting, Baltimore, 2001; Extended Abstracts, Wood Adhesives 2000, Lake Tahoe, 2000, pp. 85, 86. J. Zheng and C. E. Frazier, Proc. Forest Products Society Annual General Meeting, Baltimore, 2001; Extended Abstracts, Wood Adhesives 2000, Lake Tahoe, 2000, pp. 121, 122. J. J. Marcinko, C. Phanopoulos, and P. Y. Teachey, Extended Abstracts, Wood Adhesives 2000, Lake Tahoe, 2000, pp. 23, 24. T. Enomoto, T. Kitayama, M. Takatani, and T. Okamoto, Extended Abstracts, Wood Adhesives 2000, Lake Tahoe, 2000, p. 80. L. A. Panamgama and A. Pizzi, J. Appl. Polym. Sci. 55: 10071015 1995. G. Loew and H. I. Sachs, Proc. 11th Washington State University Int. Symp. on Particleboards, Pullman, WA, 1977, pp. 473492. I.H. Sachs, Holz Zentralblatt 103: 295296 and 384 (1977). I.H. Sachs, Polyurethane (Kunststo-Handbuch Bd.7) (G. Oertel, ed.), Carl Hanser Verlag, Munich, 1983, pp. 598604. O. Wittman, Holz Roh. Werkst. 34: 427431 (1976). H. Roll, Thesis, University of Munich, 1993. H. Roll, Proc. Holzwerkstosymposium, Mobil Oil AG, Magdeburg, Germany, 1995. D. Grunwald, Proc. 2nd European Wood-Based Panel Symposium, Hannover, Germany, 1999. M. N. Schreyer, W.-D. Domke, and S. Stini, J. Chromatogr. Sci. 27: 262266 (1989).

Copyright 2003 by Taylor & Francis Group, LLC

208. D. G. Lay and P. Cranley, in Handbook of Adhesive Technology (A. Pizzi, K.L. Mittal, eds.), 1st Edition, Marcel Dekker, New York, pp. 405429. 209. D. R. Larimer, Proc. 2nd European Wood-Based Panel Symp. Hannover, Germany, 1999. 210. J. J. Marcinko, W. H. Newman, C. Phanopoulos, and M. A. Sander, Proc. 29th Washington State University Int. Particleboard/Composite Materials Symposium, Pullman, WA, 1995, pp. 175183. 211. W. E. Johns, in Wood Adhesives: Chemistry and Technology, Vol.2 (A. Pizzi, ed.), Marcel Dekker, New York, 1989, pp. 7596. 212. C. E. Frazier, R. G. Schmidt, and J. Ni, Proc. Third Pacic Rim Bio-Based Composites Symposium, Kyoto, Japan, 1996, pp. 383391. 213. J. Kramer, Holz-Kunststoverarb. 33: 6264 (1998). 214. K. Umemura, A. Takahashi, and S. Kawai, J. Appl. Polym. Sci. 74: 18071814 (1999). 215. F. W. Abbate and H. Ulrich, J. Appl. Polym. Sci. 13: 19291936 (1969). 216. K. C. Frisch, L. P. Rumao, and A. Pizzi, in Wood Adhesives: Chemistry and Technology, Marcel Dekker, New York, 1983, pp. 289318. 217. A. Gudehn, Thesis, University of Umea, Sweden, 1984. 218. R. D. Palardy, B. R. Grenley, F. H. Story, and W. A. Yrjana, Proc. Wood Adhesives 1990, Madison, WI, 1990, pp. 124128. 219. R. Prather, D. Martone, and G. Nelson, Proc. 29th Washington State University Int. Particleboard/Composite Materials Symposium, Pullman, WA, 1995, pp. 165174. 220. L. Bolangier, Proc. Klebstoe fur Holzwerkstoe und Faserformteile, Braunschweig, Germany, 1997. 221. B. Dix, Holz Roh. Werkst. 44: 228 (1986). 222. B. Dix, Holz Roh. Werkst. 44: 328 (1986). 223. B. Dix, Holz Roh. Werkst. 45: 350 (1987). 224. B. Dix, Holz Roh. Werkst. 45: 389 (1987). 225. B. Dix, Holz Roh. Werkst. 45: 428 (1987). 226. B. Dix, Holz Roh. Werkst. 45: 487494 (1987). 227. J. Tomkinson, in (M. Dunky, A. Pizzi, and M. Van Leemput, eds.) State of the Art-Report, COST-Action E13, part I (Working Group 1, Adhesives), European Commission, Brussels, Belgium, 2002. 228. A. Pizzi, Forest Prod. J. 28(12): 4247 (1978). 229. Y. Yazaki and W. E. Hillis, Holzforschung 34: 125130 (1980). 230. Z. Guangcheng, L. Yunlu, and Y. Yazaki, Holzforschung 42: 407408 (1988). 231. Y. Yazaki, Holzforschung 37: 8790 (1983). 232. Y. Yazaki, Holzforschung 38: 7984 (1984). 233. Y. Yazaki, Holzforschung 39: 7983 (1985). 234. G. Vazquez, G. Antorrena, J. C. Parajo, and J. L. Francisco, Holz Roh. Werkst. 47: 491494 (1989). 235. K. F. Plomley, Commonwealth Scientic and Industrial Organization (CSIRO), Div. Forest Prod. Victoria, Technol. Paper 39, (1966). 236. C. Ayla, Thesis University Hamburg, Germany, 1980. 237. A. Pizzi, Int. J. Adhesion Adhesives 1: 107 (1980); 2: 213214 (1981). 238. H. M. Saayman and C. H. Brown, Forest Prod. J. 27(4): 2125 (1977). 239. C. Ayla and G. Weimann, Holz Roh. Werkst. 39: 9195 (1981). 240. V. Sealy-Fisher and A. Pizzi, Holz Roh. Werkst. 50: 212220 (1992). 241. A. Pizzi, Colloid Polym. Sci. 257: 3740 (1979). 242. A. Pizzi, Ph.D. thesis, Univeristy of the Orange Free State, South Africa, 1978. 243. A. Pizzi and A. Stephanou, Holzforsch Holzverwert. 44: 6268 (1992). 244. A. Pizzi and A. Stephanou, J. Appl. Polym. Sci. 51: 21092124 (1994); 51: 21252130 (1994). 245. N. Meikleham, A. Pizzi, and A. Stephanou, J. Appl. Polym. Sci. 54: 18271845 (1994). 246. A. Pizzi, N. Meikleham, and A. Stephanou, J. Appl. Polym. Sci. 55: 929933 (1995). 247. A. Pizzi, N. Meikleham, B. Dombo, and W. Roll, Holz Roh. Werkst. 53: 201204 (1995).
Copyright 2003 by Taylor & Francis Group, LLC

248. 249. 250. 251. 252. 253. 254. 255. 256. 257. 258. 259. 260. 261. 262. 263. 264. 265. 266. 267. 268. 269. 270. 271. 272. 273. 274. 275. 276. 277. 278. 279. 280. 281. 282. 283. 284. 285. 286. 287. 288. 289. 290. 291. 292. 293. 294. 295.

B. Dix and R. Marutzky, J. Appl. Polym. Sci., Appl. Polym. Symp. 40: 91100 (1984). C.-M. Chen, Holzforschung 36: 6570 (1982). B. Dix and R. Marutzky, Holz Roh. Werkst. 41: 4550 (1983). S. Inoue, M. Asaga, T. Ogi, and Y. Yazaki, Holzforschung 52: 139145 (1998). O. Liiri, H. Sairanen, H. Kilpelainen, and A. Kivisto, Holz Roh. Werkst. 40: 5160 (1982). C.-M. Chen, Holzforschung 45: 711 (1991). C.-M. Chen, Holzforschung 45: 303306 (1991). C.-M. Chen and J. K. Pan, Holzforschung 45: 155159, (1991). G. Vazquez, G. Antorrena, and J. C. Parajo, Holz Roh. Werkst. 44: 415418 (1986). G. Vazquez, G. Antorrena, and J. C. Parajo, Wood Sci. Technol. 21: 155166 (1987). G. Vazquez, G. Antorrena, J. Gonzales, and J. C. Alvarez, Holz Roh. Werkst. 54: 9397 (1996). E. Voulgaridis, A. Grigoriou, and C. Passialis, Holz Roh. Werkst. 43: 269272 (1985). Y. Yazaki, Holzforschung 39: 267271 (1985). Y. Yazaki and P. J. Collins, Holz Roh. Werkst. 52: 185190 (1994). A. B. Anderson, R. J. Breuer, and G. A. Nicholls, Forest Prod. J. 11: 226227 (1961). A. Pizzi, Advanced Wood Adhesives Technology, Marcel Dekker, New York, 1994, pp. 149217. A. Pizzi, in Handbook of Adhesive Technology (A. Pizzi and K. L. Mittal, eds.), 1st Edition, Marcel Dekker, New York, 1994, pp. 347358. A. Pizzi, J. Polym. Sci. Polym. Lett. Ed. 17: 489 (1979). A. Pizzi, ed., Wood Adhesives: Chemistry and Technology, Marcel Dekker, New York, 1983, pp. 174244. A. Pizzi and A. Stephanou, Holz Roh. Werkst. 52: 218222 (1994). H. A. Coppens, M. A. F. Santana, and F. J. Pastore, Forest Prod. J. 30(4): 3842 (1980). A. Pizzi, Holz Roh. Werkst. 52: 229 (1994). A. Pizzi and P. Sorfa, Holzforsch. Holzverwert. 31: 113115 (1979). R. Long, Adhaesion 35(5): 3739 (1991). A. Pizzi, Adhesives Age 20(12): 2730 (1977). A. Pizzi and F.-A. Cameron, Holz Roh. Werkst. 39: 255260 (1981). C. Ayla and G. Weimann, Holz Roh. Werkst. 40: 1318 (1982). A. Pizzi, Proc. Forest Products Society Annual Meeting, Merida, Mexico, 1998, pp. 1330. H. Heinrich, F. Pichelin, and A. Pizzi, Holz Roh. Werkst. 54: 262 (1996). A. Pizzi, Holz Roh. Werkst. 52: 286 (1994). A. Pizzi, P. Stracke, and A. Trosa, Holz Roh. Werkst. 55: 168 (1997). S. Wang and A. Pizzi, Holz Roh. Werkst. 55: 174 (1997). A. Pizzi, W. Roll, and B. Dombo, Bakelite AG, U.S. patent 5,532,330 (1996). F. Pichelin, C. Kamoun, and A. Pizzi, Holz Roh. Werkst. 57: 305317 (1999). A. Pizzi, N. Meiklham, B. Dombo, and W. Roll, Holz Roh. Werkst. 53: 201204 (1995). A. Pizzi, H. Scharfetter, and E. W. Kes, Holz Roh. Werkst. 39: 8589 (1981). A. Pizzi, J. Valenzuela, and C. Westermeyer, Holz Roh. Werkst. 52: 311315 (1994). A. Pizzi, J. Appl. Polym. Sci. 23: 27772792 (1979). L. Calve, G. C. J. Mwalongo, B. A. Mwingira, B. Riedl, and J. A. Shields, Holzforschung 49: 259268 (1995). B. Dix and R. Marutzky, Holz Roh. Werkst. 42: 209217 (1984). A. Pizzi, J. Appl. Polymer Sci. 23: 27772792 (1979). A. Pizzi, Holz Roh. Werkst. 40: 293301 (1982). C.-M. Chen, Holzforschung 46: 433438 (1992). C.-M. Chen, Holzforschung 47: 7275 (1993). C.-M. Chen, Holzforschung 48: 517521 (1994). C.-M. Chen, T.-Y. Chen, and J. Dong, Holzforschung 47: 435438 (1993). C.-M. Chen and D. L. Nicholls, Forest Prod. J. 50(3): 8186 (2000). C.-M. Chen and P. M. Winistorfer, Holzforschung 47: 507512 (1993).

Copyright 2003 by Taylor & Francis Group, LLC

296. 297. 298. 299. 300. 301. 302. 303. 304. 305. 306. 307. 308. 309. 310. 311. 312. 313. 314. 315. 316. 317. 318. 319. 320. 321. 322. 323. 324. 325. 326. 327. 328.

329. 330. 331. 332. 333. 334.

A. Trosa and A. Pizzi, Holz Roh. Werkst. 55: 306 (1997). C. Ayla and N. Parameswaran, Holz Roh. Werkst. 38: 449459 (1980). F. W. Herrick and L. H. Bock, Forest Prod. J. 8: 269274 (1958). G. Vazquez, G. Antorrena, J. L. Francisco, and J. Gonzales, Holz Roh. Werkst. 50: 253256 (1992). H. M. Saayman and J. A. Oatley, Forest Prod. J. 26(12): 2733 (1976). G. Vazquez, G. Antorrena, J. L. Francisco, M. C. Arias, and J. Gonzales, Holz Roh. Werkst. 51: 221224 (1993). A. Pizzi and D. G. Roux, J. Appl. Polym. Sci. 22: 27172718 (1978). A. Pizzi and G. M. E. Daling, Holzforsch. Holzverwert. 32: 6467 (1980). A. Pizzi and G. M. E. Daling, J. Appl. Polym. Sci. 25: 10391048 (1980). R. W. Hemingway, R. E. Kreibich, J. Appl. Polym. Sci. Appl. Polym. Symp. 40: 7990 (1984). A. Pizzi, E. Orovan, and F. A. Cameron, Holz Roh. Werkst. 46: 6771 (1988). D. Gornik, R. W. Hemingway, and V. Tisler, Holz Roh. Werkst. 58: 2330 (2000). A. Pizzi, E. Orovan, and F. A. Cameron, Holz Roh. Werkst. 42: 1217 (1984). A. Pizzi, D. du T. Rossouw, W. Knuel, and M. Singmin, Holzforschung Holzverwertung 32(6): 140150 (1980). A. Pizzi and F.-A. Cameron, Forest Prod. J. 34(9): 6167 (1984). E. Scopelitis and A. Pizzi, J. Appl. Polym. Sci. 47: 351360 (1993). A. Pizzi and M. Merlin, Int. J. Adhesion Adhesives 1: 261 (1981). A. Pizzi, E. P. von Leyser, J. Valenzuela, and J. G. Clark, Holzforschung 47: 168174 (1993). B. Dix and R. Marutzky, Holz Roh. Werkst. 43: 198 (1985). A. H. Grigoriou, Holz Roh. Werkst. 55: 269274 (1997). X. Lu and A. Pizzi, Holz Roh. Werkst. 56: 78 (1998). H. H. Nimz, in Wood Adhesives: Chemistry and Technology, Marcel Dekker, New York, 1983, pp. 247288. A. Pedersen and J. Jul-Rasmussen, Dansk Spaanplade Kompagni, German patent DE 1,303,693 (1962). K. C. Shen, Forest Prod. J. 24(2): 3844 (1974). K. C. Shen, Forest Prod. J. 27(5): 3238 (1977). K. C. Shen, German patent DE 2,410,746 (1974). K. C. Shen, D. P. C. Fung, and L. Calve, U.S. patent USP 4,265,846 (1979). K. C. Shen and D. P. C. Fung, Forest Prod. J. 29(3): 3439 (1979). K. C. Shen, L. Calve, and P. Lau, Proc. 13th Washington State University Int. Symp. on Particleboards, Pullmann, WA, 1979, pp. 369379. H. Nimz, A. Razvi, I. Mogharab, and W. Clad, Helmitin-Werke, German patent DE 2,221,353 (1972). H. H. Nimz and G. Hitze, Cellulose Chem. Technol. 14: 371382 (1980). A. Huettermann, GIT Fachz. Lab. 943950 (1989). A. Haars, A. Kharazipour, H. Zanker, and A. Huettermann, in (R. W. Hemingway and A. H. Conner, eds.), Adhesives from Renewable Resources, ACS Symposium Series 385, 1989, pp. 126134. A. Kharazipour, A. Haars, O. Milstein, M. Shekholeslami, and A. Huettermann, Proc. First Eur. Workshop Lignocell. Pulp, Hamburg, 1991, pp. 103115. A. Kharazipour, A. Haars, M. Shekholeslami, and A. Huettermann, Adhaesion 35(5): 3036 (1991). K. Nonninger, Proc. Klebstoe fur Holzwerkstoe und Faserformteile, Braunschweig, Germany, 1997. A. Haars and A. Huttermann, German patent DE 3,037,992 (1980). A. Huttermann, O. Milstein, A. Haars, K. Wehr, and G. Lovas, Peiderer AG, German patent DE 3,611,676 (1986). A. Haars and A. Huttermann, U.S. patent USP 4,432,921 (1984).

Copyright 2003 by Taylor & Francis Group, LLC

335. 336. 337. 338. 339. 340. 341. 342. 343. 344. 345. 346. 347. 348. 349. 350. 351. 352. 353. 354. 355. 356. 357. 358. 359. 360. 361. 362. 363. 364. 365. 366. 367. 368. 369. 370. 371. 372. 373. 374. 375. 376. 377. 378. 379.

A. Kharazipour and A. Huttermann, in Forest Products Biotechnology, (A. Bruce and W. Palfreyman eds.), Taylor and Francis, London, (1998). W. E. Hsu, Proc. 27th Washington State University Int. Particleboard/Composite Materials Symposium, Pullman, WA, 1993, pp. 155166. F. Kollmann, F. Schnulle, and K. Schulte, Holz Roh. Werkst. 13: 440449 (1955). M. Dunky and H. Schoergmaier, Holzforsch. Holzverwert. 47: 2630, (1995). H. Staudinger and K. Wagner, Makromol. Chem. 12: 168235 (1954). P. R. Ludlam and J. G. King, J. Appl. Polym. Sci. 29: 38633872 (1984). M. Dunky, Ph.D. thesis, Montanuniversitaet Leoben, Austria, 1980. T. Hlaing, A. Gilbert, C. Booth, Brit. Polym. J. 18: 345348 (1986). Ch.-Y. Hse, Z.-Y. Xiz, and B. Tomita, Holzforschung 48: 527532 (1994). S. Katuscak, M. Thomas, and O. Schiessl, J. Appl. Polym. Sci. 26: 381394 (1981). K. Kumlin and R. Simonson, Angew. Makromol. Chem. 93: 4354 (1981). D. Braun, M. de L. Abrao, and H.-J. Ritzert, Angew. Makromol. Chem. 135: 193210 (1985). D. Braun and H.-J. Ritzert, Angew. Makromol. Chem. 135: 193210 (1985). B. Tomita and H. Ono, J. Polym. Sci. Chem. Ed. 17: 32053215 (1979). T. Holopainen, L. Alvila, J. Rainio, and T. T. Pakkanen, J. Appl. Polym. Sci. 66: 11831193 (1997). B. Riedl and L. Calve, J. Appl. Polym. Sci. 42: 32713273 (1991). D. Braun and W. Pandjojo, Angew. Makromol. Chem. 80: 195205 (1979). R. Gnauck, G. Ziebarth, and W. Wittke, Plaste Kautsch. 27: 427428 (1980). K. Kamide and Y. Miyakawa, Makromol. Chem. 179: 359372 (1978). M. G. Kim, W. L. Nieh, T. Sellers Jr., W. W. Wilson, and J. W. Mays, Ind. Eng. Chem. Res. 31: 973979 (1992). A. W. Christiansen and L. Gollob, J. Appl. Polym. Sci. 30: 22792289 (1985). J. D. Wellons and L. Gollob, Proc. Wood Adhesives 1980, Madison, WI, 1980, pp. 1722. S. Chow and P.R. Steiner, Holzforschung 29: 410 (1975). O. B. Denisov, Holztechnol. 19: 139141 (1978). H. Matsuda and S. Goto, Can. J. Chem. Eng. 62: 108111 (1984). A. Sebenik, U. Osredkar, M. Zigon, and I. Vizovisek, Angew. Makromol. Chem. 102: 8185 (1982). G. E. Myers and J. A. Koutsky, Forest Prod. J. 37(9): 5660 (1987). M. Szesztay, Z. Laszlo-Hedvig, E. Kovacsovics, and F. Tudos, Holz Roh. Werkst. 51: 297300 (1993). M. Szesztay, Z. Laszlo-Hedvig, P. Nagy, and F. Tudos, Holz Roh. Werkst. 54: 399402 (1996). K. Umemura, S. Kawai, Y. Mizuno, and H. Sasaki, Mokuzai Gakkaishi 41: 820827 (1995). K. Umemura, S. Kawai, Y. Mizuno, and H. Sasaki, Mokuzai Gakkaishi 42: 489496 (1996). K. Umemura, S. Kawai, R. Nishioky, Y. Mizuno, and H. Sasaki, Mokuzai Gakkaishi, 41: 828836 (1995). K. Umemura, S. Kawai, H. Sasaki, R. Hamada, and Y. Mizuno, J. Adhesion 59: 87100, (1996). A. Pizzi, F. Probst, and X. Deglise, J. Adhesion Sci. Technol. 11: 573590 (1997). A. Pizzi, J. Appl. Polym. Sci. 63: 603617 (1997). R. Garcia and A. Pizzi, J. Appl. Polym. Sci. 70: 11111116 (1998). A. Pizzi, R. Garcia, and X. Deglise, J. Appl. Polym. Sci. 67: 16731678 (1998). C. Zhao, S. Garnier, and A. Pizzi, Holz Roh Werkst. 56(6): 402 (1998). Y. Laigle, C. Kamoun, and A. Pizzi, Holz Roh Werkst. 56(3): 154 (1998). A. Pizzi, X. Lu, and R. Garcia, J. Appl. Polym. Sci. 71: 915925 (1999). C. Kamoun and A. Pizzi, Holz Roh Werkst. 58(4): 288289 (2000). C. Zhao and A. Pizzi, Holz Roh Werkst. 58(5): 307308 (2000). C. Kamoun, A. Pizzi, and R. Garcia, Holz Roh. Werkst. 56: 235243 (1998). P. R. Steiner and S. R. Warren, Forest Prod. J. 37(1): 2022 (1987). S. Garnier, Ph.D. thesis, University of Nancy 1, Epinal, France, 2002.

Copyright 2003 by Taylor & Francis Group, LLC

380. 381. 382. 383. 384. 385. 386. 387. 388. 389. 390. 391. 392. 393. 394. 395.

396. 397. 398. 399. 400. 401. 402. 403. 404. 405. 406. 407. 408. 409. 410. 411. 412. 413. 414. 415. 416. 417. 418. 419. 420. 421. 422. 423. 424.

S. Garnier, O. C. Vorster, and A. Pizzi, unpublished results, 1999. M. Lecourt, P. Humphrey, and A. Pizzi, Holz Roh Werkst. 61: 75 (2003). H. Neusser, U. Krames, and M. Zentner, Holzforsch. Holzverwert. 28: 7987 (1976). H. Neusser and M. Zentner, Holzforsch. Holzverwert. 26: 5463 (1974). T. F. Duncan, Forest Prod. J. 24(6): 3644 (1974). W. F. Lehmann, Forest Prod. J. 24(1): 1926 (1974). H.-A. May and G. Keserue, Holz Roh. Werkst. 40: 105110 (1982). H. Neusser and U. Krames, Holzforsch. Holzverwert. 21: 7780 (1969). P. W. Post, Forest Prod. J. 8: 317322 (1958). P. W. Post, Forest Prod. J. 11(1): 3437 (1961). G. Rackwitz, Holz Roh. Werkst. 21: 200209 (1963). M. Jaic, R. Zivanovic, T. Stevanovic-Janezic, and A. Dekanski, Holz Roh. Werkst. 54: 3741 (1996). E. Liptakova, J. Kudela, Z. Bastl, and I. Spirovova, Holzforschung 49: 369375 (1995). E. Zavarin, in The Chemistry of Solid Wood (R. Rowell, ed.,) Am. Chem. Soc., Adv. in Chem. Ser. 207, 1984, pp. 349400. J. J. Bikerman, The Science of Adhesive Joints, Academic Press, New York, 1961. P. Pulkkinen and L. Suomi-Lindberg, in (M. Dunky, A. Pizzi and M. Van Leemput, eds.) State of the Art-Report, COST-Action E13, Part I (Working Group 1, Adhesives), European Commission, Brussels, Belgium, 2002. R. J. Good, J. Adhesion 4: 133154 (1972). I. Johansson and M. Stehr, Proc. Forest Products Society Annual Meeting, Vancouver, 1997. M. Stehr, J. Seltman, and I. Johansson, Holzforschung 53: 93103 (1999). J. J. Bikerman, Ind. Eng. Chem. 59(9): 4044 (1967). G. J. Crocker, Rubber Chem. Technol. 42(1): 3070 (1969). C.-M. Chen, Forest Prod. J. 20(1): 3641 (1970). M. Scheikl, Thesis, University of Agricultural Sciences, Vienna, Austria, 1995. M. Kazayawoko, A. W. Neumann, and J. J. Balatinecz, Wood Sci. Technol. 31: 8795 (1997). A. Bogner, Holz Roh. Werkst. 49: 271275 (1991). V. R. Gray, Forest Prod. J. 12: 452461 (1962). Ch.-Y. Hse, Holzforschung 26: 8285 (1972). T. F. Shupe, C. Y. Hse, and W. H. Wang, Proc. Forest Products Society Annual Meeting, Merida, Mexico, 1998, pp. 132136. T. F. Shupe, C. Y. Hse, E. T. Choong, and L. H. Groom, Forest Prod. J. 48(6): 9597 (1998). Q. Shen, J. Nylund, and J. B. Rosenholm, Holzforschung 52: 521529 (1998). G. Elbez, Proc. Wood-Based Composite Products CSIR Conference, Pretoria, South Africa, 1985. P. O. Rozumek and G. Elbez, Holzforschung 39: 239243 (1985). J. D. Wellons, Forest Prod. J. 30(7): 5355 (1980). K. Suchsland, Holz Roh. Werkst. 15: 385390 (1957). A. Herczeg, Forest Prod. J. 15: 499505 (1965). T. Nguyen and W. E. Johns, Wood Sci. Technol. 13(1): 2940 (1979). E. Kehr and W. Schilling, Holztechnol. 6: 225232 (1965). E. Plath, Holz Roh. Werkst. 11: 392400 (1953). R. Popper, Holzbau 44: 168170 (1978). E. Roael and W. Rauch, Holz Roh. Werkst. 32: 182187 (1974). D. Narayanamurti, Holz Roh. Werkst. 15: 370380 (1957). D. C. Maldas and D. P. Kamdem, Forest Prod. J. 49(11/12): 9193 (1999). R. M. Rowell, Wood Sci. 15: 172182 (1982). H. Tarkow, A. J. Stamm, and E. C. O. Erickson, Forest Prod. Lab. Rep. 1593, USDA Forest Service, Forest Products Laboratory, Madison, WI, 1950. M. Gomez-Bueso, J. Westin, R. Torgilsson, P. O. Olesen, and R. Simonson, Holz Roh. Werkst. 57: 433438 (1999).

Copyright 2003 by Taylor & Francis Group, LLC

425. 426. 427. 428. 429. 430. 431. 432. 433. 434. 435. 436. 437. 438. 439. 440. 441. 442. 443. 444. 445. 446. 447. 448. 449.

M. Gomez-Bueso, J. Westin, R. Torgilsson, P. O. Olesen, and R. Simonson, Holz Roh. Werkst. 58: 914 (2000). A. Pizzi, A. Stephanou, M. J. Boonstra, and A. J. Pendlebury, Holzforschung 48: Suppl. 9194 (1994). P. Hanetho, Proc. FESYP-Tagung, Federation Europeenne du Syndacat des fabricants de Panneaux, Munchen, 1987, pp. 129136. E. L. Back, Forest Prod. J. 41(2): 3036 (1991). M. Dunky, Holzforsch. Holzverwert. 40: 126133 (1988). M. Dunky, Proc. Second European Panel Products Symposium, Llandudno, Wales, 1998, 206217. E. Meineke and W. Klauditz, Research Report, Nordrhein-Westfalen Provincial Government, Germany, 1962. J. B. Wilson and M. D. Hill, Forest Prod. J. 28(2): 4954 (1978). G. A. Eusebio and N. C. Generalla, FPRDI J. 12: 1219 (1983). W. F. Lehmann, Forest Prod. J. 15: 155161 (1965). G. D. Waters, Proc. National Particleboard Association (NPA) Resin and Blending Seminar, Irving, Texas, 1990, pp. 5661. D. Robson, M. Riepen, J. Hague, C. Loxton, and R. Quinney, Proc. First European Panel Products Symposium, Llandudno, Wales, 1997, pp. 203210. P. E. Humphrey, Thesis, University of Wales, Bangor, Wales, 1982. P. E. Humphrey, Proc. 25th Washington State University Int. Particleboard/Composite Materials Symposium, Pullman, WA, 1991, pp. 99108. P. E. Humphrey and S. Ren, J. Adhesion Sci. Technol. 3: 397413, (1989). P. E. Humphrey and D. Zavala, J. Testing Evaluation 17: 323328 (1989). X. Lu and A. Pizzi, Holz Roh Werksto 56(5): 393401 (1998). F. Pichelin, A. Pizzi, A. Fruhwald, and P. Triboulot, Holz Roh Werksto 59(4): 256265 (2001). F. Pichelin, A. Pizzi, A. Fruhwald, and P. Triboulot, Holz Roh Werksto 60(1): 917 (2002). F. Fahrni, Holz Roh. Werkst. 14: 810 (1956). R. Keylwerth, Holzforsch. Holzverwert. 11: 5157 (1959). F. Kollmann, Holz Roh. Werkst. 15: 3544 (1957). M. D. Strickler, Forest Prod. J. 9: 203215 (1959). G. V. Haas, Thesis, University of Hamburg, Germany, 1998. G. V. Haas, A. Steen, and A. Fruehwald, Holz Roh. Werkst. 56: 386392 (1998).

Copyright 2003 by Taylor & Francis Group, LLC

Potrebbero piacerti anche