Sei sulla pagina 1di 29

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/268398080

Melamine–Formaldehyde Adhesives

Article · August 2003


DOI: 10.1201/9780203912225.ch32

CITATIONS READS

12 9,344

1 author:

A.Pizzi Pizzi
University of Lorraine
866 PUBLICATIONS   15,922 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Isocyanate-Free Polyurethane View project

Densified wood surfaces View project

All content following this page was uploaded by A.Pizzi Pizzi on 20 January 2015.

The user has requested enhancement of the downloaded file.


32
Melamine–Formaldehyde Adhesives

A. Pizzi
Ecole Nationale Supérieure des Technologies et Industries du Bois,
Université de Nancy 1, Epinal, France

I. INTRODUCTION

Melamine–formaldehyde (MF) and melamine–urea–formaldehyde (MUF) resins are


among the most used adhesives for exterior and semiexterior wood panels and for the
preparation and bonding of both low- and high-pressure paper laminates and overlays.
Their much higher resistance to water attack is their main distinguishing characteristic
from urea–formaldehyde (UF) resins. MF adhesives are expensive. For this reason, MUF
resins which have been cheapened by addition of a greater or lesser amount of urea are
most often used. Notwithstanding their widespread use and economical importance, the
literature on melamine resins is only a small fraction of that dedicated to UF resins. Often
MFs and MUFs are described in the literature as a subset of UF amino resins. This is not
really the case, as they have peculiar characteristics and properties all of their own which
in certain respects are very different from those of UF adhesives.

II. USES FOR MF RESINS

Melamine–formaldehyde resins are used as adhesives for exterior- and semiexterior-grade


plywood and particleboard. In this application their handling is very similar to that of UF
resins for the same use, with the added advantage of their excellent water and weather
resistance. MF resins are also used for the impregnation of paper sheets in the production
of self-adhesive overlays for the surface of wood-based panel products and of self-adhesive
laminates. In this application the impregnation substrate, cellulose paper, is thoroughly
impregnated by immersing it in the resin solution, squeezing it between rollers, and drying
without curing it to proper flow by passing it through an airdraft tunnel oven at 70 to
120 C at  10 m/s. The dry MF-impregnated sheets can then be bonded by one of two
main processes:
1. The sheets of MF-impregnated paper, consisting of one surface layer or a few
surface layers, are bonded together and with a substrate of paper sheets impreg-
nated with phenolic resins to form laminates of variable thickness. In the
impregnated papers is the dry but still active MF resin, which functions as the
adhesive of the MF-impregnated sheet to both MF-impregnated sheets and at

Copyright © 2003 by Taylor & Francis Group, LLC


the interface between MF-impregnated and phenol–formaldehyde (PF)-impreg-
nated layers. These laminates are high-pressure laminates.
2. The MF in an impregnated paper sheet is not completely cured but still has a
certain amount of residual activity and is applied directly in a hot press, in a
single sheet, on a wood-based panel, to which it bonds by completing the MF
adhesive curing process.
Press platens are made from stainless steel or chromium-plated brass and copper.
The chromium layer preserves surface quality longer than does ordinary steel. The MF
laminates exhibit a remarkable set of characteristics. Because of their unusual chemical
inertness, nonporosity, and nonabsorbance, they resist most substances, such as mild
alkalies and acids, alcohols, solvents such as benzene, mineral spirits, natural oils, and
greases. No stains are produced on MF surfaces by these substances. In addition to almost
unlimited coloring and decorating possibilities, this remarkable resistance has resulted in
the extensive use of MF laminated wood-based panel products for tabletops, sales coun-
ters, laboratory benches, heavy-duty work areas in factories and homes, wall paneling, and
so on.

III. CHEMISTRY

A. Condensation Reactions
The condensation reaction of melamine (I) with formaldehyde (Fig. 1) is similar to but
different from the reaction of formaldehyde with urea. As for urea, formaldehyde
first attacks the amino groups of melamine, forming methylol compounds. However,

Figure 1 Methylolation (hydroxymethylation) and subsequent condensation reactions to form


melamine–formaldehyde adhesive systems.

Copyright © 2003 by Taylor & Francis Group, LLC


formaldehyde addition to melamine occurs more easily and completely than does addition
to urea. The amino group in melamine accepts easily up to two molecules of formalde-
hyde. Thus complete methylolation of melamine is possible, which is not the case with urea
[1]. Up to six molecules of formaldehyde can be attached to one molecule of melamine.
The methylolation step leads to a series of methylol compounds with two to six methylol
groups. Because melamine is less soluble than urea in water, the hydrophilic stage proceeds
more rapidly in MF resin formation. Therefore, hydrophobic intermediates of the MF
condensation appear early in the reaction. Another important difference is that MF con-
densation to give resins, and their curing, can occur not only under acid conditions, but
also under neutral or even slightly alkaline conditions. The mechanism of the further
reaction of methylol melamines to form hydrophobic intermediates is the same as for
UF resins, with splitting off of water and formaldehyde. Methylene and ether bridges
are formed and the molecular size of the resin increases rapidly. These intermediate con-
densation products constitute the large bulk of the commercial MF resins. The final curing
process transforms the intermediates to the desired MF insoluble and infusible resins
through the reaction of amino and methylol groups which are still available for reaction.
A simplified schematic formula of cured MF resins has been given by Koehler [2] and
Frey [3]. They emphasize the presence of many ether bridges besides unreacted methylol
groups and methylene bridges. This is because in curing MF resins at temperatures up to
100 C, no substantial amounts of formaldehyde are liberated. Only small quantities are
liberated during curing up to 150 C. However, UF resins curing under the same conditions
liberate a great deal of formaldehyde.
At the condensation stage attention must be paid to the formation of hydrolysis
products of the melamine before preparation starts. The hydrolysis products of melamine
are obtained when the amino groups of melamine are gradually replaced by hydroxyl
groups. Complete hydrolysis produces cyanuric acid (Formula 1).

Formula 1

Ammeline and ammelide can be regarded as partial amides of cyanuric acid. They
are acid and have no use in resin production. They are very undesirable by-products of the
manufacture of melamine because of their catalytic effect in the subsequent MF resin
production, due to their acidic nature. If present, both must be removed from crude
melamine by an alkali wash and/or crystallization of the crude melamine.

Copyright © 2003 by Taylor & Francis Group, LLC


B. Mechanisms and Kinetics
The mechanism of the initial stages of the reaction of melamine with formaldehyde leading
to the formation of methylol melamines is very similar to that of urea. The reaction
mechanism of the acid-catalyzed condensation reactions of methylol melamines to form
polymers and resins has been elucidated by Sato and Naito [4]. Melamine and formalde-
hyde react similarly to urea and formaldehyde, although basic differences are evident in
the reaction rates and mechanism. The primary products of reaction are methylolmela-
mines, and evidence indicates that such compounds are formed only at ambient or higher
temperature except in acid pH ranges. The reaction is reversible throughout the pH range.
Its forward rate is proportional to either [melamine][HCHO] or [melamine][HþCHOH] or
[melamineþ][HCHO], according to the pH used.
Methylolmelamine forms ‘‘dimers’’ by condensation with melamine under neutral
and acid conditions (70 C); this process is irreversible. The initial hydroxymethylation is
very rapid. Its rate is determined by the condensation of conjugated acids of methylolme-
lamines with melamine. The reaction rate is proportional to [melamine]2[HCHO] [5].
When the [mineral acid]/[melamine] ratio is 0.0 to 1.0, the early stage hydroxymethylation
of melamine is dependent on the concentration of the melamine molecule (base species)
MH and its conjugated acid MHþ 2 in the following manner [6]:

rate ¼ kH2 O ½MH ½HCHO þ kH ½MHþ þ


2 ½HCHO þ kMH2þ ½MH2 ½MH ½HCHO þ

þkMH ½MH 2 ½HCHO


in the absence of added acid, when the ratio [mineral acid]/[melamine] is ¼ 0, the rate of
the reaction can thus be represented as
rate ¼ kH2 O ½MH ½HCHO þ kMH ½MH 2 ½HCHO
The condensation reaction has been studied by investigating the kinetics of the initial stage
of the condensation of di- and trimethylolmelamine (MF2 and MF3) in the pH range 1 to
9. Regardless of pH, the initial rate is equal to [4]:
rate ¼ k½MFn 2 ðwith n ¼ 2 or 3Þ
In the presence of mineral acid, the main reaction at the early stage of the condensation is
the reaction between the methylolmelamine molecule and its conjugated acid (MFnHþ) [7].
This was found at an [acid]/[MFn] (n ¼ 2 or 3) ratio lower than 1.0 (pH 2.7). With an
[acid]/[MFn] ratio higher than 1.0 to 1.2 (pH<2), the main condensation takes place
between the conjugated acids themselves.
At equal pH values the condensation rate of trimethylolmelamine is considerably
faster than that of dimethylolmelamine. This is the opposite of the rates of mono- and
dimethylolurea. This means that while the nitrogen of the amido group in the case of urea
is more reactive and therefore more nucleophilic than the nitrogen of the amidomethylol
group, the opposite is true in the case of melamine. The reaction for MF2 is primarily
between the carbon of the methylol group next to the nitrogen in HMþCH2OH, and the
nitrogen of the amino group in MCH2OH. For MF3, the condensation is mainly between
the carbon of the methylol group next to the charged nitrogen in HþMCH2OH, and the
nitrogen of the aminomethylol group in MCH2OH [4]. The condensation rate therefore
increases with the increasing electrophilicity of the carbon of the methylol group and the
increasing nucleophilicity of the nitrogen of the amino group or aminomethylol group.
Therefore in MF3 the carbon in HMþCH2OH is more electrophilic than the same
carbon in MF2. On the other hand, the nitrogen of the aminomethylol group in

Copyright © 2003 by Taylor & Francis Group, LLC


HMþCH2OH of MF3 is less nucleophilic, and therefore less reactive, than the nitrogen of
the amino group of MF2. The effects of the carbon and nitrogen atoms are consequently
opposite to each other in the MFn condensation. Since the effect of the carbon is greater
than the effect of the nitrogen on the reaction rate, MF3 condenses faster than MF2. At
lower pH values the effect of the nitrogen becomes negligible and MF3 is even faster than
MF2 in condensing to polymers.
The difference between the kinetic behavior of urea and melamine can be ascribed to
the different effect of the nitrogen atom in the two compounds. With regard to the for-
mation of methylol compounds as a result of hydroxymethylation, the functionality of
melamine has been observed to be 6 against formaldehyde [1,8]. Similarly, melamine reacts
easily with formaldehyde to form MF3; it also forms MF6 in concentrated formaldehyde
[1,4,8]. For example, urea readily forms dimethylolurea, but forms trimethylolurea with
marked difficulty [1,8] and never forms tetramethylolurea. These results suggest that the
nitrogen of the amidomethylol group in methylolurea is considerably less nucleophilic
than the nitrogen of the amido group in urea. However, the nitrogen of the aminomethylol
group in methylolmelamine is not markedly less nucleophilic than the nitrogen of the
amino group in melamine. Presumably, this is due to the difference in basicity between
urea and melamine. The same is also true of their condensation reactions.

C. Mixed Melamine Resins


With regard to melamine–urea–formaldehyde, copolymers can be prepared which are
generally used to cheapen the cost of MF resins, but which also show some worsening
of properties. Copolymerization was proven by means of model compounds and poly-
condensates [9]. MUF resins obtained by copolymerization during the resin preparation
stage are superior in performance to MUF resins prepared by mixing preformed UF and
MF resins, especially because processing of such mixtures is quite difficult [10]. The rela-
tive mass proportions of melamine to urea used in these MUF resins is generally in the
melamine:urea range 50:50 to 30:70 [11]. Melamine–phenol–formaldehyde resins, which in
some respects show better properties than those of their corresponding MF and PF resins,
have also been prepared [12–14]. Analysis of the molecular structure of those resins in both
their uncured and cured states appeared to show that no co-condensates of phenol and
melamine form and that two separate resins coexist. This is due to the difference in
reactivity of the phenolic and melamine methylol groups as a function of pH. Also, in
their cured state an interpenetrating network of the separate PF and MF resins, as a
polymer blend, is formed, not a copolymer of the two [15–18]. Today MUF resins are
produced in greater amounts than MF resins in the field of adhesives due to the relatively
high cost of melamine: their formulation has progressed to such a level that often no
difference in performance exists between a good MUF resin and a pure MF resin. MF
resins are still more extensively used at this stage in the paper impregnation/laminates
fields although both MUF copolymers as well as separate, double application of UF
(paper core) and MF (paper surfaces) resins are making considerable inroads in this
area. MUF resins instead totally dominate today in the wood adhesives field. Paper
laminates and wood adhesives are the two main application areas of these resins.
A type of resin also used today is the so-called PMUF (or MUPF according to which
author is writing) adhesives. These are fundamentally MUF resins in which a minor
proportion of phenol (between 3 and 10%; phenol:melamine:urea by weight of 10:30:60
for example) has been assumed to have coreacted with to further upgrade weather resis-
tance of the bonded joint. Unfortunately the alleged superior performance of such resins is

Copyright © 2003 by Taylor & Francis Group, LLC


Figure 2 Schematic representation of the dependence on the type of formulation used of the fate of
phenol in a PMUF resin. (1) Phenol only present as unlinked free phenol/phenol derivatives but
mainly as a pendant group neither participating in resin cross-linking nor contributing to resin
performance and water resistance. (2) Intermediate case. (3) Case in which phenol is co-condensed
and participating in the cross-linked network.

often only wishful thinking as the phenol has frequently not been properly reacted with the
other materials, and consequently the PMUF resin will have a worse performance than a
comparable top of the range MUF resin. This was confirmed by the demonstration that it
depends exclusively on the resin manufacturing parameters and materials reaction order
used whether or not the phenol coreacts within many PMUF adhesives, showing that
often the phenol remains as a useless pendant group in the hardened aminoplastic
(MUF) network without contributing at all to its performance [19,20] (Fig. 2).
The best reaction order necessary to obtain PMUF resins in which phenol makes a
positive contribution to the performance of the hardened network has been reported [19].
PMUF resins are still used and some good resins of this type are indeed used in the
unrealistic hope that they outperform equivalent MUF resins, when it has been shown
clearly that they perform at best as a MUF adhesive presenting the same number of moles
of melamine for the total moles of phenol plus melamine of the PMUF itself. The idea that
the addition of small percentages of phenol to a MUF resin yields resins of better exterior
durability is then an incorrect myth perpetuated in the wood panels industry. Newer
formulations of MUF resins always outperform the corresponding PMUF. PMUFs are
not bad resins, they are simply resins in which one of the materials, phenol, is often wasted
for no purpose.

IV. RESIN PREPARATION, GLUE MIXING, AND HARDENING

Because of their characteristic rigidity and brittleness in their cured state, when MF resins
are used for impregnated paper overlays, small amounts (typically 3 to 5%) of modifying
compounds are often copolymerized with the MF resin during its preparation to give
better flexibility to the finished product and better viscoelastic dissipation of stress in
the joint. Most commonly used are acetoguanamine, e-caprolactam, and p-toluene-
sulfonamide (Formula 2).
The effect of these is to decrease cross-linking density in the cured resin due to the
lower number of amidic or aminic groups in their molecules. Thus in resin segments where

Copyright © 2003 by Taylor & Francis Group, LLC


Formula 2

they are included, only linear segments are possible, decreasing the rigidity and brittleness
of the resin. Acetoguanamine is most used for modification of resins for high-pressure
paper laminates, while caprolactame, which in water is subject to the following
equilibrium (Formula 3),

Formula 3

is used primarily for low-pressure overlays for particleboard. Small amounts of noncopo-
lymerized plasticizers such as diethylene glycol can also be used for the same purpose. Due
to the peculiar structure of the wood product itself, MF adhesives for particleboard
generally do not need the addition of these modifiers. Often, a small amount of dimethyl-
formamide, a good solvent for melamine, is added at the beginning of the reaction to
ensure that all the melamine is dissolved and is available for reaction. Sugar is often added
to lessen cost of the resin. The aldehyde group of sugars have been proven to be able to
condense with the amine groups of melamine and hence to copolymerize in the resin. Their
quantity in MF resins must be limited to very low percentages, and if possible, sugars
should not be used at all, as with aging they tend to cause yellowing, crazing, and cracking
of cured MF paper laminates and to have a bad effect on adhesive long-term water
resistance in both plywood and particleboard.
MF adhesive resins for plywood and particleboard must be prepared to quite dif-
ferent characteristics than those for paper impregnation. The latter must have lower
viscosity but still high resin solids content because they need to penetrate the paper sub-
strate to a high resin load, to be dried without losing adhesive capability, and only later to
be able to bond strongly to a substrate. Instead, MF adhesive resins for plywood and
particleboard are generally more condensed, to obtain lower penetrability of the wood
substrate (otherwise, some of the adhesive is lost by overpenetration into the substrate).
The reverse applies for paper substrates, where the contrasting characteristics desired—
good paper penetration and fast curing—can be obtained in several ways during resin
preparation. These characteristics can be achieved by producing, for example, a resin with
a lower degree of condensation and high methylol group content. Typically, a MF resin of
a lower level of condensation with melamine/formaldehyde molar ratio of 1:1.8 to 1:2 will
give the desired characteristics. Its high methylol content and somewhat lower degree of
polymerization will give low viscosity at a high resin solids content, favoring rapid wetting

Copyright © 2003 by Taylor & Francis Group, LLC


and impregnation of the paper substrate, while the high proportion of methylol groups
will give it fast cross-linking and curing capabilities.
A second, equally successful approach in to produce a MF resin of lower methylol
group content and higher degree of condensation to which a small second addition of
melamine (typically, 3 to 5% total melamine) is effected toward the end of resin prepara-
tion. The shift to lower viscosity and higher solids content given by a second addition of
melamine, shifting to lower values the average of the resin molecular mass distribution,
yields a resin of rapid impregnation characteristics. Conversely, the higher degree of
polymerization of the major part of the resin gives fast cross-linking, and curing, due to
the lower number of reaction steps needed to reach gel point. Typical total melamine/
formaldehyde molar ratios used in this system are 1:1.5 to 1:1.7.
Figures 3 and 4 show typical temperature and pH diagrams for the industrial man-
ufacture of MF and MUF resins for adhesives and other applications. The important
control parameters to take care of during manufacture are the turbidity point (the point
during resin preparation at which addition of a drop of MF reaction mixture to a test tube
of cold water gives slight turbidity) and the water tolerance or hydrophobicity point,
which marks the end of the reaction. The latter is a direct measure of the extent of
condensation of the resin and indicates the percentage of water or mass of liquid on the
reaction mixture that the MF resin can tolerate before precipitating out. It is typically set
for resins of higher formaldehyde/melamine ratios and lower condensation levels at
around 170 to 190%, but for resins of lower formaldehyde/melamine molar ratios and
higher condensation levels it is set at around 120%. As can be seen from the diagrams in
Fig. 3, once maximum reaction temperature is reached, pH is lowered to 9 to 9.5 to
accelerate formation of the polymer. Once the turbidity point is reached, pH is again
increased to 9.7 to 10.0, to slow down and more finely control the end point, determined
by reaching of the wanted value of the water tolerance point. Industrial MF resins are
generally manufactured to a 53 to 55% resin solids content with a final pH of 9.9 to 10.4
(but lower pH values are also used for low-condensation resins). To have acceptable rates
of curing if higher pH values are used, higher quantities of hardener need to be used,
which is clearly uneconomical. For typical MF resins for low pressure (particleboard),

Figure 3 Typical temperature and pH diagrams for the industrial manufacture of MF resins.

Copyright © 2003 by Taylor & Francis Group, LLC


Figure 4 Typical manufacturing diagram for 40:60 to 50:50 melamine/urea weight ratio
MUF resins.

self-adhesive overlay pressing times of between 30 and 60 s at 170 to 190 C press tempera-
ture are required according to the type of resin used. Pressing conditions for particleboard
and plywood adhesives are identical to those used for UF resins.
Glue mixing presents different requirements according to the final use of the MF
resin. Hardeners are either acids or materials that will liberate acids on addition to the
resin or on heating. In MF and MUF adhesives for bonding particleboard and plywood,
the use of small percentages of ammonium salts, such as ammonium chloride or ammonium
sulfate, is well established and is indeed identical to standard practice in UF resins. In MF
adhesives for low- and high-pressure self-adhesive overlays and laminates the situation is
quite different. Ammonium salts cannot be used for the latter application for three main
reasons. First, evolution of ammonia gas during drying and subsequent hot curing of the
MF impregnated paper would cause high porosity of the cured MF overlay. Second, the
stability of ammonium salts, in particular of ammonium chloride, might cause MF liquid
resin whitening and the MF-impregnated paper to cure and deactivate at ambient tempera-
ture after a short time in storage, causing the resin to have lost its adhesive capability by the
time it is needed in hot curing. Third, the elimination of ammonia during drying and curing
would leave the cured, finished paper laminate essentially very acid due to the residual acid
of the hardener left in the system. This badly affects the resistance to water attack of the
cured MF surface defeating the primarily advantage for which such surfaces have justly
become so popular. Thus a stable, self-neutralizing, non-gas-releasing hardener is needed
for such an application. Several have been prepared and one of the most commonly used is
the readily formed complex between morpholine and p-toluenesulfonic acid. Morpholine
and p-toluenesulfonic acid readily react exothermically to form a complex of essentially
neutral pH that is stable up to well above 65 C (Formula 4).

Formula 4

Copyright © 2003 by Taylor & Francis Group, LLC


Table 1 Typical Paper Impregnation Glue Mix for Self-Adhesive
Low-Pressure MF Overlays

Ingredient Parts by mass

MF resin, 53% solids content 99.1


Release agent 0.08
Wetting agent 0.16
Hardener (morpholine/p-toluenesulfonic acid complex) 0.64
Defoamer 0.02

During heat curing of the MF paper overlay in the press, the complex decomposes,
the MF resin is hardened by the acid that is liberated, morpholine is not vaporized and lost
to the system, and on cooling the complex is reformed, leaving the cured glue line essen-
tially neutral.
In MF glue mixing for overlays and laminates, small amounts of release agents to
facilitate release from the hot press of the cured bonded overlay are added. Small amounts
of defoamers and wetting agents to further facilitate wetting and penetration of the resin in
the paper are always added. A typical glue mix is shown in Table 1.
Two strong trends have appeared reasonably recently in the preparation of mela-
mine-impregnated paper laminates. First, impregnating machines capable of giving papers
in which much cheaper UF resin is substituting as much as 50% of the more expensive MF
resin have now been in operation for several years. This equipment is based on a double
impregnating bath application: the paper passes through a first bath where it absorbs the
UF resin first, the excess on the surfaces being scraped off in-line, and then passes through
a second bath where it absorbs the MF resin. The concept is to limit the UF resin to the
inside of the paper with the MF resin coating the outside of the paper: the hardened
surface after final curing will then have all the waterproof characteristics of a MF paper
laminate but at a lower price. Good results are obtained and many machines using this
type of process are today in industrial operation. A more recent trend has been to develop
MUF copolymers to use with the less costly single impregnating bath machines. A few
cases of this route to coping with the high cost of melamine are on record.

V. MUF ADHESIVE RESINS OF UPGRADED PERFORMANCE

Several effective techniques to consistently and markedly decrease the melamine content in
MUF wood adhesives without any loss of performance have also been recently developed.
Some of these formulation systems and techniques are already in the early stages of
industrialization. Among these melamine/acid salts, such as melamine acetate (Formula
5), function both as efficient hidden hardeners of UF resins for plywood as well as upgrad-
ing the performance of simple UF resins for plywood by approximately 10% by mass
melamine grafting to yield comparable strength durability of premanufactured MUF
resins of 30 to 40% melamine mass content, hence of resins of much higher mass content
of melamine. In short a MUF resin of melamine:urea weight ratio 10:90 will perform in
certain applications such as exterior plywood as a premanufactured MUF resin of mela-
mine:urea between 30:70 and 40:60 [21–24]. The system works both (i) by simple addition
of the melamine salt in the UF glue mix eliminating the need to premanufacture a MUF
resin. The effectiveness of melamine grafting in the glue mix and during hot pressing has

Copyright © 2003 by Taylor & Francis Group, LLC


been found to depend on the relative solubility of the melamine salt which depends on
both the acid strength of the acid as well as the number of acid functions in the salt. (ii) By
use of salts in which the excess acid has been eliminated from the salt, hence melamine
monoacetate with no loose acid residue. The salt can be added in the resin factory to a UF
resin and the mix sold as a MUF resin as pot life is indefinite and the resin needs the
addition of a classical hardener for aminoplastic resins such as sodium sulfate or sodium
chloride for hardening. The solubility of the salts used increases with increase in tempera-
ture. The reasons why traditional, premanufactured MUF resins waste 2/3 or more of the
melamine used in them, and why such a melamine salt addition system is so much more
effective by not wasting melamine were presented in the same study [21].

Formula 5

How is it possible that addition of a melamine salt to a UF glue mix in a melami-


ne:urea mass ratio of 10:90 yields plywood of comparable water resistance to a prereacted
MUF resin of melamine:urea mass ratio in the range 30:70 to 40:60? As a consequence of
what is presented above it is now possible to answer such a question. In the preparation of
precopolymerized MUF resins, hence of today’s normal, commercial MUF resins, during
the high temperature preparation reaction the melamine also reacts with formaldehyde to
form short MF chains which are then bound to the more abundant UF chains. Hardening
of MUF resins has been proven to occur almost exclusively by cross-linking through
–CH2– bridges connecting two melamines [20,25] as, due to its much lower reactivity,
urea is not greatly involved. The use of melamine salts at ambient temperature in the
glue mix instead ensures that only single melamine molecules are singly and separately
grafted on the UF resin chain.

UCH2MCH2MCH2M against UCH2M

to yield rather different cross-linked networks than those of a standard MUF reactor-
made resin [21–26].
As to cross-link the system only a very small amount of melamine molecules for each
UF chain is needed to achieve the same effect, to have several chains of MF as in standard
MUF resins does not improve the bond strength because (i) only one of the melamines in
the chain will react, the other not participating at all in final cross-linking, and (ii) the
bonding strength will also not be improved by having even all the melamines of the MF
chain react all in the same space zone of the network as shown in the first network formula
above: on the contrary, the highly localized position on vicinal sites in the network of a
high density of cross-links might well render the resin far too rigid and far too brittle

Copyright © 2003 by Taylor & Francis Group, LLC


(which indeed is the case for most melamine-based resins). It is then clear that at least 2/3
of the melamine presently used in MUF resins is actually wasted and does not contribute
much to the final results other than in a damaging manner, this being unavoidable as a
consequence of the system of preparation used. The new system presented greatly
improves on the present situation, not only on ease of handling (only a UF resin and a
melamine salt as a hardener are needed rather than a more sophisticated MUF resin), but
also on the amount of melamine needed (just approximately 1/3 of present consumption
for equal exterior-grade bonding performance) with potentially considerable economic
advantages as melamine is generally expensive.
The results of a 2 year field weathering test in Europe have confirmed that a UF resin
to which has been added 15% melamine acetate salt at the glue mix stage, to obtain a
melamine:urea mass ratio of 10:90 solids on solids, imparts a better durability and better
exterior performance to plywood glue lines than traditionally reactor-coreacted MUF
resins of melamine:urea mass ratio of 33:66 and even of commercial, prereacted PMUF
resin where the relative mass proportions of the materials in the resin are 10:30:60 [23].
Postcuring of aminoplastic-bonded wood joints has always been avoided due to the
evident degradation induced by heat and humidity on the aminoplastic resin hardened
network. This is a known fact and it is for this reason that boards bonded with UF, MUF,
and MF resins are traditionally cooled as rapidly as possible after manufacture. However,
tightening of formaldehyde emission regulations has caused considerable progress in ami-
noplastic formulations, especially much smaller molar ratios, and hence today’s amino-
plastic adhesives are indeed very different materials than those of 10–20 years ago. A
recent study [27,28] has shown that the postulate on the avoidance of postcuring of
aminoplastic resin bonded joints is under many conditions no longer valid. Thus, (1)
postcuring (for example by hotstacking in the simpler cases, by an oven or other heat
treatment in more sophisticated cases) can be used in principle and under well-defined
conditions to improve the performance of UF and MUF-bonded joints and panels
without any further joint and hardened adhesive degradation, as the value of the modulus
reached during postcuring is always consistently higher than the value at which the
modulus stabilizes after complete curing during the ‘‘pressing’’ cycle. (2) Postcuring
could also be used in principle and for the same reasons to further shorten the pressing
time of MUF-bonded joints and panels when well-defined postcuring conditions are used
or to decrease the proportion of adhesive used at parity of performance [27,28]. (3) There
is clear indication that under certain conditions, even when adhesive degradation starts,
the application of the posttreatment reestablishes the value of the joint’s strength to a
value higher than its maximum value obtained during curing. Some of the best posttreat-
ment schedules have also been presented [27] (Fig. 5).
The performance improvements in the internal bond (IB) strength of bonded wood
panels are introduced by the series of reactions pertaining to internal methylene ether
bridge rearrangements to a tighter methylene bridge network which have already been
observed and extensively discussed in the analysis of aminoplastic and phenolic resins
[27,29–31]. These are able to counterbalance well the degradative trend to which the
aminoplastic resin should be subjected. Furthermore, in modern resins of lower
F:(U þ M) molar ratio the amount of methylene ether bridges formed in curing is much
lower. Thus, disruption by postcuring of the already formed resin network by internal
resin rearrangements will be milder, if at all present, and will definitely not yield the
marked degradation and even collapse of the structure of the network which characterizes
older resins of much higher molar ratio when postcured under the same conditions [32,33].
In short, notwithstanding the internal rearrangement the network will stand and stand

Copyright © 2003 by Taylor & Francis Group, LLC


Figure 5 Thermomechanical analysis of a joint glued with a modern MUF adhesive showing the
advantage of hot-post-stacking for modern, low molar ratio aminoplastic adhesive bonded panels.
Note the maximum modulus achieved during isothermal heating (180 C for 8 min) (lower curve) and
maximum modulus achieved after cooling and reheating at 100 C for 8 min (upper curve): the
difference in modulus is the potential gain due to hot-post-stacking.

quite strongly: no, or hardly any decrease of IB strength will be noticeable. For modern,
lower molar ratio aminoplastic adhesives, since the resin network does not noticeably
degrade or collapse with postcuring, only the tightening of the network derived by further
bridge formation by reaction within the network of the few formaldehyde molecules
released by the now mild internal rearrangement will be noticeable: the IB strength
value will then improve with postcuring in boards bonded with modern, lower formalde-
hyde aminoplastic adhesives.
There are important differences in the behavior of MUF resins prepared in different
ways, and hence at the level of their performance as binders of wood panels, due both to
their differences at the level of the resin structure and to the type and distribution of the
molecular species formed before hardening, as well as to the differences in the structure of
the final hardened networks. An example of three types of MUF resins examined can
illustrate this point. (i) A sequential MUF in which the UF was prepared first and then
melamine coreacted afterwards once the UF polymer had been formed [8], a last small
urea addition also being carried out for a final (M þ U):F molar ratio of 1:1.5 and M:U
weight ratio of 47:53, (ii) a MUF resin in which the great majority of the urea and of the
melamine were premixed and then reacted simultaneously to form the resin, followed by
addition of small amounts of both last melamine and last urea, for a (M þ U):F molar
ratio of 1:1.5 and M:U weight ratio of 47:53, and (iii) a UF resin of molar ration 1:1.5 to
which has been added 15% by weight on resin solids of monoacetate of melamine in the
glue mix for a final (M þ U):F molar ratio of 1:1.39 and M:U weight ratio of 14:86. The
proportion and type of chemical species formed which can be calculated by the molar
proportions of the reagent, the manner in which these are combined during the reaction
under different conditions as well as the rate reaction constants of urea and melamine with
formaldehyde lead to the conclusion, confirmed by 13C nuclear magnetic resonance
(NMR), that the distribution of species for resins (i), (ii), and (iii) are as follows (their
relative proportions are indicated in Formulas 6, 7, and 8).
Case (i) above presents the following predominant chemical species (Formula 6),
where M attached to the UF polymer is in the form of both a single melamine as well as in
the form of a melamine formaldehyde short oligomer.

Copyright © 2003 by Taylor & Francis Group, LLC


Formula 6

Case (iii) above presents instead just UF oligomers and melamine salts (Formula 7),

Formula 7

where M is always in the form of a single melamine molecule.


Case (ii) above presents the following predominant chemical species (Formula 8),

Formula 8

Thus an MF resin drowned in mostly unreacted urea and where M attached to the UF
polymer is in the form of both a single melamine (M and M framed) as well as in the form
of a MF short oligomer (M framed).
The structure of the three resins when still in liquid form explains the appearance of
their structure after hardening. Thus, hardened MUF resins of formulation type (ii) will
present structures as presented in Formula 8 and thus will waste the benefit of a consider-
able proportion of the melamine used. Hardened MUF resins of type (i) will present
structures intermediate between those shown in Formulas 7 and 8 (but tending more to
the type of Formula 8) and thus while also wasting a considerable proportion of the

Copyright © 2003 by Taylor & Francis Group, LLC


melamine used, this will be less than for formulations of type (ii): the strength and water
resistance results of MUFs of type (iii) will then be noticeably better at parity of all other
conditions than what is obtainable with resins of type (ii), as indeed has been shown to be
the case. MUF resin formulations of type (iii), those of melamine acetate type, will give
hardened structures according to Formula 7 without wasting much melamine and giving
hence the best performance, with the limitation of proportion already mentioned and
explained above. This can be seen by comparing the strength results obtained by constant
heating rate thermomechanical analysis (TMA) [26]. A MUF formulation of type (iii)
containing 20% melamine acetate performs almost as well as a good formulation of
type (i) which contains two and a half times more melamine. They both perform much
better than a formulation of type (ii) [26] with some notable exceptions [38].
Another recent approach which has shown considerable promise in markedly
decreasing the percentage of adhesive solids on a board, and hence in markedly decreasing
melamine content, has been found almost by chance. It is based on the addition of certain
additives to the MUF resin. Additives have been found that are both able to decrease
melamine content in MUF resins at parity of performance, as well as able to decrease the
percentage of any MUF resin needed for bonding while still conserving the same adhesive
and joint performance. This second class of additives works for UF adhesives too, but less
well, while it gives acceptable results for PF resins, but it is at its best in the case of MUF
resins. This second class of additives is the acetals [34–36], methylal and ethylal being the
two most appropriate due to their cost to performance ratio, which do not release for-
maldehyde at pHs higher than 1 [35]. Methylal has according to results reported by the
Environmental Protection Agency (EPA) an LD50 value of 10,000 against that of 100 in
the case of formaldehyde, and is thus classed as nontoxic. The addition of these materials
to the glue mix of formaldehyde-based resin improves considerably its mechanical resis-
tance and the performance of the bonded joint.
This is in general valid for MUFs, UFs, and PFs, but the effect is particularly evident
for the MUF resins [35]. Decreases in MUF resin solids content of as much as 33% while
conserving the same performance are reported in the case of wood particleboard. In Fig. 6
are shown the continuous heating rate TMA curves of modulus as a function of tempera-
ture for an MUF resin of 1:1.2 (M þ U):F molar ratio. Similar but much less extreme

Figure 6 Thermomechanical analysis graph showing the increasing maximum values of the
modulus of a MUF-bonded joint with increasing amounts of methylal (an acetal) as an effectiveness
upgrading additive.

Copyright © 2003 by Taylor & Francis Group, LLC


trends are obtained also for UF and PF resins. In the case of MUF resins the addition of
10% additive on resin solids yields laboratory particleboard in which one can decrease the
percentage of resin solids on the board by between 20% and 25% without any loss of
performance. Similarly, at equal resin solids the strength of a particleboard is 33% higher
when 10% additive on resin solids is added to the glue mix. Addition of 20% methylal on
the board yields, in the case of the same resin, the same strength with 30% less adhesive
(and hence less melamine) [35].
What is the mechanism of action of methylal, ethylal, and some other acetals to
achieve such a feat? Their excellent solvent action on melamine and higher molecular
weight ligomers. The cases shown earlier in this chapter referring to melamine salts and
the loss of effectiveness due to wastage of melamine are applicable in this case too.
Melamine when added to a reacting mixture during resin manufacture is not really
soluble. It reacts then in heterogeneous phase with the other components of the resin,
some of it being in a transient state in equilibrium between being in solution and being
out of solution, and thus its efficacity is partially, but noticeably reduced. The intro-
duction of an excellent solvent, none better than these acetals was known before,
brings the totality of the reaction into homogeneous phase with a consequent, notice-
able improvement in both the effectiveness of reaction and the effectiveness of mela-
mine utilization.
A different class of additives from those above but also able to decrease melamine
content in MUF resins at parity of performance also exist. They are based on the addition
in the glue mix of 1 to 5% additive and allow preparation of MUF copolymers, prema-
nufactured in a traditional manner, in which either the proportion of melamine is lower,
for example a 20:80 by weight M:U resin to which the additive has been added performing
as well as a M:U 50:50 resin, or alternatively to upgrade a top of the range M:U 50:50
MUF adhesive to an exterior performance comparable and even superior to that of PF
resins [37–39]. Several different types of additives can achieve this but they are all based on
the preparation and acid stabilization of imines, or better of iminomethylene bases [38,39],
and of their addition to the MUF resin. Thus, the effect is still the same whether the
imines/iminomethylene bases, acid-anion stabilized, are prepared by coreaction of ammo-
nia and formaldehyde [38,39], or for instance as described for acid-anion stabilized decom-
position of hexamethylenetetramine [38,39] (Fig. 7). The structure of the imines and the
iminomethylene bases yielding this effect are very similar indeed to the structure of the
acetal additives presented above, the –NH– bridge of the imines having the same function
of the –O– bridge of the acetals. The imines/iminomethylene bases have the added dimen-
sion, however, that the nitrogen can function as a knot of tridimensional cross-linking
itself, which the oxygen bridge obviously cannot do. The amount of nitrogen-based addi-
tive that can be used is limited by its higher sensitivity to water in the hardened network.
This is not the case of the possibly less effective oxygen-based additives, which can be used
in greater amount: one property balances the other. The oxygen bridge conversely presents
perhaps a better longer-term thermal stability than the nitrogen-based bridges. These are
only very relative, rather subjective advantages. What is instead important is that the
similarity of structure indicates that in the main (but not completely) the mode of
action of all these additives may appear to be the same, but often different effects are at
work, namely first a considerable improvement of the viscoelastic dissipation of the energy
of the glue line and bonded joint without a drop in cross-linking density. The differences
between the different additives is then due to additional, although rather important effects
such as the solvent effect of the acetals in the MUF resins, and the increase in reaction rate
[25] and buffer effect [38] of the iminomethylene basis, as well as others. It is on the basis of

Copyright © 2003 by Taylor & Francis Group, LLC


Figure 7 Mechanism of hexamethylenetetramine decomposition leading to the formation of anion-
stabilized reactive iminomethylene bases. The same bases can be formed by reaction of ammonium
salts such as ammonium sulfate and formaldehyde and constitute a metastable intermediate between
hexamine and final decomposition products, and vice versa (after refs 26, 37–39).

this similarity of structure and effect that a scale of additives providing similar effects to
different levels has been established (see Formula 9) [40].
It must be pointed out that a TMA strength improvement of 100% on the MUF
resin without methylal (this is achieved by addition of 20% methylal on resin solids)
corresponds in the actual wood particleboard to an increase of IB strength of 33%.
This means that of all the compounds shown above only the acetals, such as methylal
and ethylal, as well as the similarly structured imine/iminomethylene bases discussed
above (for which the effect on strength is more marked) are capable of marked improve-
ments in IB strength at the actual wood panel level.
These developments are of use for MUF resins not just in the field of wood adhe-
sives, or of other binders in general, but also to improve and upgrade the performance of

Copyright © 2003 by Taylor & Francis Group, LLC


Formula 9

resins in other applications such as that of melamine-based impregnated paper laminates,


where they have been shown to improve considerably the storage stability of paper
impregnating resins [36].
As regards the more application-bound physical aspects of MUF resins these can be
applied in different ways, this too sometimes having a bearing on other types of additives
used. Thus to the normal case of a MUF plus its hardener one can add cases in which a
formaldehyde depressant such as a low condensation MF, UF, or MUF precondensate or
one of their mixes is added; sometimes this is in combination with an accelerator based on
the same principle. Such an approach is more used in other resins, but it has been shown
and reported as being feasible also for MUF resins [41].

VI. TEMPERATURE–TIME-TRANSFORMATION AND CONTINUOUS-


HEATING-TRANSFORMATION CURING DIAGRAMS OF
MUF RESINS WHEN ALONE AND HARDENING IN A
WOOD JOINT (OR OTHER INTERACTIVE SUBSTRATE)

Temperature–time-transformation (TTT) and continuous-heating-transformation (CHT)


curing diagrams for polycondensation resins are starting to acquire more importance in
the deductions of the behavior of different resins during hardening. They are a type of
state diagram. TTT and CHT diagrams of resins by themselves or on noninteracting
substrates show similar trends as exemplified by the case of epoxy resins on glass fibers
[42,43] (Fig. 8a). Different trends than those for TTT and CHT diagrams of epoxy resins
reported in the literature (Fig. 8a) occur, however, in the higher and lower temperature
zones of the diagrams of waterborne formaldehyde-based resins hardening on wood. CHT
and TTT diagrams have already been reported for PF, UF, MUF, phenol–resorcinol–
formaldehyde (PRF), and tannin–formaldehyde thermosetting resins [31,44–46] (Figs. 8b
and c). The higher temperature zone of the CHT diagrams for MUF resins in a wood
joint, reported in Fig. 8c, shows the same trends (and for the same reasons, namely the
interactive nature of the substrate and movement of water from resin to substrate and vice
versa) observed for UF and PF resins.
However, the experimental TTT diagram in Fig. 8b shows quite a different trend
from the CHT diagram for the same resins and for the TTT diagrams reported in the
literature for epoxies on noninteracting substrates such as glass fiber. To start to under-
stand the trend shown in Figs. 8b and c it is first necessary to observe what happens to the
modulus of the wood substrate alone (without a resin being present) when examined under

Copyright © 2003 by Taylor & Francis Group, LLC


Figure 8 (a) Schematic classical TTT and CHT curing diagrams of epoxy resins and any other
polycondensation resin, such as MUFs, on a nonreactive, noninterfering substrate. (b) Schematic
TTT curing diagram of MUF, PRF, UF, and PF wood adhesives on wood as an interacting
substrate. (c) Schematic CHT curing diagram of PRF, MUF, UF, and PF wood adhesives on
wood as an interacting substrate.

Copyright © 2003 by Taylor & Francis Group, LLC


the same conditions of a wood joint during bonding. No significant degradation occurs up
to a temperature of 180 C as shown by the relative stability of the value of the elastic
modulus as a function of time. Some slight degradation starts to occur at 200 C, but after
some initial degradation the elastic modulus again settles to a steady value as a function of
time and at a value rather comparable to the steady value obtained at lower temperatures.
Evident degradation starts to be noticeable in the 220–240 C range and this becomes even
more noticeable at higher temperatures. The effect of substrate degradation on the TTT
diagram in Fig. 8b can then only start to influence the trends in gel and vitrification curves
at temperatures higher than 200 C and it is for this reason that the region of the curves
higher than 200 C is indicated by dashed lines in Fig. 8b. At a temperature  200 C the
trends observed are due to the resin only. In this range of temperature the eventual turning
to longer time and stable temperature of the vitrification curve, characteristic of the TTT
diagrams of epoxy resins, becomes also evident for the TTT diagrams of the waterborne
PRF and MUF resins on lignocellulosic substrates indicating that diffusion hindrance at a
higher degree of conversion becomes for these resins too the determinant parameter
defining reaction rate. What differs, however, from previous diagrams is that the trend
of all the curves, namely the gelation curve, initial pseudogel (entanglement) curve, and
start and end of vitrification curve is the same. In epoxy resin TTT diagrams the trend of
the gelation curve is completely different from that reported here. The result shown in
Fig. 8b is, however, rather logical because if diffusion problems alter the trend of the
vitrification curve, then the same diffusional problem should also alter the gelation and
pseudogel curves. This is indeed what the experimental results in Fig. 8b indicate. It may
well be that in waterborne resins the effect is more noticeable than in epoxy resins. This is
the reason why it is possible to observe it for PF, UF, PRF, and MUF resins. With the
data available and with the limitation imposed by the start of wood substrate degradation
of higher temperatures it is not really possible to say if the gelation curve and the vitrifica-
tion curve run asymptotically towards the same value of temperature at time ¼ 1
although the indications are that this is quite likely to be the case. What is also evident
in the trend of the two curves is the turn to the left, hence the inverse trend of their
asymptotic tendency towards Tg1. This turn cannot be ascribed to substrate degradation
because for very reactive resins, such as PRFs, such a turn already occurs at a temperature
lower than 150 C, hence much lower than the temperature at which substrate degradation
becomes significant. This inverse trend can only be attributed to movements of water
coming from the substrate towards the resin layer as the trend of the curves indicates
an easing of the diffusional problem already proven to occur at such a high degree of
conversion [30,39].
Two other aspects of the TTT diagrams in Figs. 8b and c must be discussed, these
being the trend of the curves at temperatures higher than 200 C and the trend of the
devitrification (or resin degradation) curve. The dashed line trend and experimental points
of all the curves at temperatures higher than 200 C are clearly only an effect caused by the
ever more severe degradation of the substrate: degradation of the substrate implies a
greater mobility of the polymer network constituting the substrate, hence the continuation
of the curves as shown in their segmented part. That this is the case is also supported by
the virtual negative times yielded by the TMA equipment when the temperature becomes
extreme, as well as by the trend of the resin’s higher degradation curve which tends to
intersect the vitrification curve at about 200–220 C or higher, this being a clear indication
that one is measuring the changes in the reference system, the substrate itself, and that
these are at this stage so much more important than the small changes occurring in the
resin as to be able to dominate the whole complex system which is the bonded joint.

Copyright © 2003 by Taylor & Francis Group, LLC


The CHT and TTT diagrams pertaining to waterborne formaldehyde-based poly-
condensation resins on a lignocellulosic substrate should then appear in their entirety as
shown in Figs. 8b (TTT) and 8c (CHT) rather than as the classical diagrams of epoxies on
noninterfering substrates such as glass fiber shown in Fig. 8a.

VII. COLD-SETTING MUF ADHESIVES

MUF resin can be used as a cold-setting wood laminating adhesive for glulam and fin-
gerjointing by the use of adequate acid hardeners. In all semiexterior and protected exter-
ior structural applications where a clear/invisible glue line is preferred for aesthetic reasons
then a MUF adhesive is preferred to the classical PRF adhesives used for this purpose. It is
then more a question of fashion cycles, but notwithstanding this MUF resins have taken a
considerable hold today in Europe (contrary to North America where PRFs are by far
preferred) and confidence in them for this application has been steadily growing.
PRF ‘‘honeymoon’’ fast-set, separate application adhesives for exterior-grade struc-
tural glulam and fingerjointing have now been used industrially for about twenty years
[1,8] in several relevant variations developed over the years. MUF resins are now taking
the same ‘‘honeymoon’’ direction: the use of a melamine resin and a resorcinol separate
component system [47] has been reported. However, for all the improvements made to the
commercial MUF resins of this type in all their different variations, they were still based
on some resorcinol or resorcinol-aided component. Thus, using as one component a MUF
resin of high melamine content and resorcinol as a second component is just unusual in its
use of a MUF rather than a PRF resin; a very acceptable resin concept but for the fact that
it is coupled with a phenol such as resorcinol. The coupling of an acid-setting MUF
adhesive and of resorcinol might well present no advantages or even some potentially
serious disadvantages. It has been shown for example that thermosetting PMUF resins
do not present a better performance than equivalent MUF resins and that often, depend-
ing on their sequence of manufacture, present instead a much worse performance. There
are very well-defined technical and chemical reasons for this [19,20] that boil down to the
relevant differences in reactivity of the two materials, namely the phenol (here resorcinol)
and melamine. The reactivity of melamine and even urea at the acid-setting pHs they need
is much greater than that of any phenol, even resorcinol, as this pH range is that of the
lowest reactivity of any phenol. Thus, even resorcinol runs the risk of being linked very
little to the MUF matrix, especially in a fast-setting system such as a honeymoon, and at
best it will remain as a bypassed pendant side group not able to fully achieve the function
for which it has been added.
More recently an exclusively MUF-based honeymoon adhesive for glulam and fin-
gerjoints has been developed and reported in which one component is a high performance
MUF resin, while the second separate application component is based on just slightly
acidified water thickened to the same viscosity of the first component by the addition of
1.5% carboxymethyl cellulose (CMC) [48,49] (Fig. 9). The system has also been tried
successfully in industry for both fast production of fingerjointing (Fig. 10) and glulam
and also for the fast production of ambient temperature pressed plywood [48,49]. MUF-
based, honeymoon-type, fast-setting, separate application adhesive systems which do not
need any resorcinol are then capable of performing as adhesives for structural exterior-
grade joints and glulam and of satisfying all the requirements of the relevant adhesive
specifications for such an application. The parameters that were shown to be determining
are mainly the performance of the MUF resin, if and once an excellent resin formulation is

Copyright © 2003 by Taylor & Francis Group, LLC


Figure 9 Tensile strength increase as a function of time of beech joints (BS 1204, Part 1) bonded
with MUF-based honeymoon adhesive systems: effect of the variation of the initial application pH of
the resin (component A).

Figure 10 Four-point bending strength increase as a function of time of pine (Pinus sylvestris)
fingerjoints bonded with MUF-based honeymoon adhesive systems.

available both the ratio of melamine to urea and the molar ratio having a lesser effect,
performance only starting to drop lower than the requirements of relevant standards when
M:U weight ratios fall well below 20:80 and of the order of 10:90. Addition of resorcinol at
these failing levels while improving slightly the performance did not solve the problem;
resorcinol addition then does not allow specification requirements to be satisfied [48,49].

Copyright © 2003 by Taylor & Francis Group, LLC


At the higher M:U ratios such as M:U ¼ 47:53, but even at lower melamine contents,
addition of resorcinol does not improve the results at all, its addition again revealing itself
superfluous. The reasons for such a behavior are those already presented and explained
above. The MUF honeymoons present all the other usual advantages associated with
honeymoon adhesives, namely high curing rate, long pot life, tolerance to higher moisture
content of the substrate, and tolerance to even quite severe imbalances in viscosity and
proportions between the two components.

VIII. CHEMICAL AND PHYSICAL ANALYSIS

The analysis of these resins is difficult when unknown products, particularly fully cured
have to be tested for UF and MF resins. Widmer [50] offers a method for the identification
of UF and MF resins in technical products. This involves preparing crystalline products of
urea and melamine and identifying them under the microscope. Melamine (in the form
of melamine crystals) and urea (in the form of long, crystalline needles of urea dixanthate)
can be seen. This method allows one to distinguish between urea and melamine even in
a cured adhesive joint.
Quantitative determination of MF resins is also rather difficult. A method was
developed by Widmer [50] for the quantitative determination of melamine in MF
condensation products. In this method the resins are destroyed under pressure by
aminolysis leaving the melamine intact. This is then converted to melamine picrate,
which is easily crystallized and weighed. The Widmer method makes it possible to
determine quantitatively the presence of urea and melamine in intermediate condensation
products and in cured UF and MF resins (even when they have been mixed). Estimations
are seldom in error by more than a few percent.
Hirt et al. [51] have published an effective and rapid method for the detection
of melamine by ultraviolet spectrophotometry. This method can be used for products
containing MF. It makes use of the strong absorption of the melamine ion at 235 nm.
The resin is extracted from comminuted MF samples by hydrolyzing to melamine
by boiling under reflux in 0.1 N hydrochloric acid. Stafford [52] also gives a method for
the identification of melamine in wet-strength paper.
Uncured MF resin analyses are carried out by gel permeation chromatography
(GPC) and 13C NMR. GPC was an inconvenient method for MF resins, although it
has become much more accepted for this purpose in recent years. Dimethylformamide,
dimethylsulfoxide, or salt solutions are generally used as solvents with a differential
refractometer as a detector. Derivatives of the resin, such as those obtained by silylation,
are generally used to decrease molecular association by hydrogen bonding. 13C NMR is a
more convenient technique, and the chemical shifts of the different structural groups in the
resin can easily and readily be identified [52,61]. This method is also quite convenient
in comparing MF and MUF resin structures obtained by different manufacturing
methodologies [52–61].
More recently effective equations correlating the results obtained by liquid-phase
13
C NMR of the liquid MF and MUF resin before hardening with the strength and degree
of cross-linking of the resin in the hardened state, hence of the IB strength and formalde-
hyde emission of boards bonded with them were developed and reported by two different
groups [41,62–64]. One set of equations applied to different formulations can be used
to determine the chemical characteristics of the resin without knowing anything
of its manufacturing parameters, procedures or molar ratio [63,64] while a different

Copyright © 2003 by Taylor & Francis Group, LLC


set of equations from a different group requires the previous knowledge of the molar
ratio of the resin [62]. They are both very useful tools for the characterization of
such resins.
Recently, thermomechanical analysis (TMA) has been used to characterize
the performance of polycondensation resins, including MF and MUF adhesives, by cor-
relating the deflection in bending of a beech wood joint bonded with a thermosetting resin
with the IB strength the same resin can give when used to prepare a wood particleboard
under given, industrially significant conditions [15,18,48] according to the equation
IB ¼ a (1/f) þ b, where f is the deflection obtained by TMA in three-point bending and a
and b are coefficients characteristic of the type of resin used [65–67]. The experiments can
be performed on a preprepared joint in isothermal mode or starting with the liquid glue
between the veneer substrates in nonisothermal mode and hence following the hardening
of the resin in situ within the joint. What is measured is the narrowest deflection of the
cured composite joint at whatever temperature this is achieved, this being correlated
through a factor with the inverse of the modulus of the bonded joint, as well as the
increase of modulus as a function of time or temperature characterizing in situ the kinetic
performance of a particular MUF or MF resin. The system can also give the
characteristics of the resin-hardened network by calculation of both the energy of
adhesion of interaction with the substrate as well as the level of tightness of the hardened
network through the average size of segments between cross-linking nodes (and
entanglement nodes) [68–70].
MF and today also MUF resins produce high quality plywood and particleboard
because their adhesive joints are boilproof. Considerable discussion has occurred and
many investigations have been carried out on the weather resistance of MF and MUF
adhesives. Many authors uphold the good weather resistance of the more recently
developed MF and MUF adhesives, especially those in which small amounts of phenol
(PMUF or MUPF) have been incorporated. The more general trend, however, is to
consider the wood products manufactured with these resins as capable of resistance to
limited weather and water exposure only, such as in flooring applications, rather than
being capable of true exterior-grade weather resistance for which phenolic adhesives are
preferred. Figure 11 shows the different behavior of traditional older generation
MF-bonded and PF-bonded particleboard in a series of wet–dry cycles. Whereas
PF-bonded boards initially deteriorate rapidly then stabilize to a constant swelling
value, MF particleboards have slower initial deterioration but never stabilize and continue
deteriorating with time and additional wet–dry cycles. This indicates that MF-bonded

Figure 11 Typical trends of irreversible thickness swelling characteristics of wood particleboard


bonded with MF and PF adhesives during a wet–dry cycle test.

Copyright © 2003 by Taylor & Francis Group, LLC


wood is not completely impervious to further water attack, indicating the fundamental
susceptibility of the aminoplastic bond to water. The rate of deterioration, and therefore
bond hydrolysis, increases as the temperature increases. Considering the insolubility
of melamine in cold water, this is quite understandable. Recent developments in MUF
resins (as indicated in this chapter) have, however, introduced a question mark over this
rather conservative definition of the weather resistance of MF and MUF resins in relation
to phenolics, and clear indications exist that with wise formulation MUFs can indeed
perform as fully fledged, heavy-duty exterior adhesives capable of competing successfully
with both phenolic and diisocyanate resins.

IX. FORMULATIONS

For starting experiments in MF resins, the following formulations are suggested.

A. MF Formulation for Exterior Particleboard


In a reaction vessel charge at 25 C, 44.4 parts by mass of water, add 30% NaOH solution
to pH 11.2 to 12.0, followed by 15.5 parts of 91% paraformaldehyde prills, 34.4 parts of
melamine powder, 2.8 parts of caprolactam, and 2.5 parts of N,N0 -dimethylformamide
while maintaining the temperature at 25 C; heat in  40 min to 92 to 95 C. When the
temperature reaches 80 C, adjust the pH with 30% NaOH solution, if necessary, to pH
9.9. At 93 C, cool to 90 C and maintain the temperature there. Adjust pH to 9.55 to 9.65
with formic acid. Hold the pH at this value while checking, adjusting, and recording the
pH value every  10 min. Check for the turbidity point at 10-min intervals until the
turbidity point is reached. At this time bring pH up to 9.95 to 10.05. Check, adjust,
and record the pH every 10 min. Start distilling water under vacuum to a solid of  53
to 55%. Check the water tolerance at 10-min intervals until it is 170 to 180%. Then apply
full vacuum and cool the resin to 30 to 35 C.

B. Formulation for Low-Pressure MF Paper-Impregnated Overlays


Follow the same procedure as for formulation A, but at the end of the water vacuum
distillation add 1.7 to 1.9 parts by mass of the second melamine and heat the reaction
mixture to 95 C again and maintain this temperature for 5 to 6 min, then cool rapidly.

C. MUF Formulation for Exterior Particleboard


This sequential MUF formulation can be successfully manufactured at different
(M þ U):F molar ratios according to exactly the same procedure, and not only for the
proportions as indicated in the example that follows. Thus the same formulation gives
excellent results for example at (M þ U):F molar ratios of 1:1.5 and 1:1.7, but not only
these ratios. For more MUF formulations see references [8,11,38].
To 113 parts by weight of Formurea (a formaldehyde concentrate stabilized by urea,
of mass content 57% formaldehyde and 23% urea. NB: Formurea comes in other con-
centrations too) are added 13 parts of urea and 30 parts of water. The pH is set at 10 to
10.4 and the temperature brought to 92 to 93 C under continuous mechanical stirring. The
pH is then lowered to 7.8 and the reaction continued at the same temperature, allowing the
pH to fall by itself over a period of 1 h 30 min to 1 h 35 min to a pH of 5.2 (one should

Copyright © 2003 by Taylor & Francis Group, LLC


strictly prevent the pH falling under 5 to avoid both uncontrollable reactions taking hold
as well as a decrease in the finished adhesive performance later). To bring the pH back to
9.5 or higher, 22% NaOH water solution was added, followed by 41 parts by weight of
melamine premixed with 19 parts of water. One part of dimethylformamide and 2 parts of
diethylene glycol are then added to the reaction mixture, maintaining a temperature of
93 C. The water tolerance is checked every 10 min while the pH is allowed to fall by itself.
When the water tolerance reached is 180 to 200% (this is often reached after 35 to 40 min,
and the pH reached is of 7.2), 6.5 parts by weight of second urea is added and the pH is
again brought up to 9.5. The reaction is continued until the water tolerance reached is
lower than 150% (the pH has reached generally 7.7 at this stage). The pH is then corrected
to 9–10.2 again and the reaction mixture cooled and stored. Resins produced using this
procedure have solids contents of 58 to 65%, a density of 1.260 to 1.280 at 20 C, a
viscosity of 70 to 150 cP, free formaldehyde of approximately 0.32, and gel times with
3% NH4Cl of 51 to 57 s at 100 C. Increasing and lowering of pH where the pHs indicated
are not reached by the reaction time can be done by addition of 22 to 33% NaOH water
solution (pH increases) and by addition of formic or acetic acid (pH decrease). The
preparation diagram of this resin is shown in Fig. 4.

REFERENCES

1. A. Pizzi, in Wood Adhesives Chemistry and Technology, Vol. 1 (A. Pizzi, ed.), Marcel Dekker,
New York, 1983, Chap. 2.
2. R. Koehler, Kunststoffe Tech. 11: 1 (1941); Kolloid Z. 103: 138 (1943).
3. R. Frey, Helv. Chim. Acta 18: 491 (1935).
4. K. Sato and T. Naito, Oikym. J. (Japan) 5(2): 144 (1973).
5. M. Akano and Y. Ogata, J. Am. Chem. Soc. 74: 5728 (1952).
6. K. Sato and S. Ouchi, Polymer J. (Japan) 10(1): 1 (1978).
7. A. Takahaski, Chem. High Polymer (Japan) 7: 115 (1950); Chem. Abstr. 46: 438 (1952); Chem.
High Polymer (Japan) 9: 15 (1952); Chem. Abstr. 48: 1730 (1954).
8. A. Pizzi, Advanced Wood Adhesives Technology, Marcel Dekker, New York, 1994.
9. D. Braun and H.-J. Ritzert, Angew. Makromol. Chem. 156:1 (1988); 135: 193 (1985).
10. D. Braun and H.-J. Ritzert, Kunststoffe 77: 1264 (1987).
11. T. A. Mercer and A. Pizzi, Holzforschung Holzverwertung 46(3): 51 (1994).
12. A. Bachmann and T. Bertz, Aminoplaste, VEB Verlag fur Grundstoffindustrie, Leipzig, 1967, p.
81.
13. A. Knop and W. Scheib, Chemistry and Application of Phenolic Resins, Springer Verlag, Berlin,
1979, p. 134.
14. K. Bruncken, in Kunststoffhandbuch, Vol. 10 (R. Vieweg and E. Becker, eds.), Hanser, Munich,
1968, p. 352.
15. D. Braun and W. Krausse, Angew. Makromol. Chem. 108: 141 (1982).
16. D. Braun and W. Krausse, Angew. Makromol. Chem. 118: 165 (1983).
17. D. Braun and H.-J. Ritzert, Angew. Makromol. Chem. 125: 9 (1984).
18. D. Braun and H.-J. Ritzert, Angew. Makromol. Chem. 125: 27 (1984).
19. C. Cremonini, A. Pizzi, and P. Tekely, Holz Roh Werkstoff 54(2): 85 (1996).
20. M. Higuchi, J.-K. Roh, S. Tajima, H. Irita, T. Honda, and I. Sakata, Polymeric structures of
melamine-based composite adhesives, in Proceedings No. 4735 of the Adhesives and Bonded
Wood Symposium, Forest Products Society, Madison, WI, 1994, pp. 429–449.
21. M. Prestifilippo, A. Pizzi, H. Norback, and P. Lavisci, Holz Roh Werkstoff 54(6): 393 (1996).
22. C. Cremonini and A. Pizzi, Holzforschung Holzverwertung 49(1): 11 (1997).

Copyright © 2003 by Taylor & Francis Group, LLC


23. C. Cremonini and A. Pizzi, Holz Roh Werkstoff 57(5): 318 (1999).
24. C. Kamoun and A. Pizzi, Holz Roh Werkstoff 56(1): 86 (1998).
25. A. Pizzi and L. A. Panamgama, J. Appl. Polymer Sci. 58: 109 (1995).
26. A. Pizzi, High performance MUF resins of low melamine content by a number of novel
techniques, Wood Adhesives 2000 Proceedings, Forest Products Society, Madison, WI,
2000.
27. X. Lu and A. Pizzi, Holz Roh Werkstoff 56(6): 393 (1998).
28. C. Zhao and A. Pizzi, Holz Roh Werkstoff 58(5): 307 (2000).
29. R. Garcia and A. Pizzi, J. Appl. Polymer Sci. 70(6): 1111 (1997).
30. C. Kamoun, A. Pizzi, and R. Garcia, Holz Roh Werkstoff 56(4): 235 (1998).
31. A. Pizzi, X. Lu, and R. Garcia, J. Appl. Polymer Sci. 71(6): 915 (1999).
32. B. Meyer, Urea-formaldehyde Resins, Addison-Wesley, Reading, MA, 1979.
33. A. Pizzi, Aminoplastic wood adhesives, in Wood Adhesives Chemistry and Technology
(A. Pizzi, ed.), Marcel Dekker, New York, 1983, Chap. 2.
34. Lambiotte & Co., 1999, Technical data sheet, Brussels.
35. A. Pizzi, M. Beaujean, C. Zhao, M. Properzi, and Z. Huang, J. Appl. Polymer Sci. 84: 2561
(2002).
36. M. Zanetti, A. Pizzi, M. Beaujean, H. Pasch, K. Rode, and P. Dalet, J. Appl. Polymer Sci. 86:
1855 (2002).
37. A. Pizzi, P. Tekely, and L.A. Panamgama, Holzforschung 50: 481 (1996).
38. M. Zanetti, A. Pizzi, and C. Kamoun, Holz Roh Werkstoff 61: 55 (2003).
39. F. Pichelin, C. Kamoun, and A. Pizzi, Holz Roh Werkstoff 57(5): 305 (1999).
40. M. Zanetti and A. Pizzi, J. Appl. Polymer Sci., in press (2003).
41. L. A. Panamgama and A. Pizzi, J. Appl. Polymer Sci. 59: 2055 (1996).
42. J. B. Enns and J. K. Gillham, J. Appl. Polymer Sci. 28: 2831 (1983).
43. G. Wisanrakkit, J. K. Gillham, and J. B. Enns, J. Appl. Polymer Sci. 41: 1895 (1990).
44. A. Pizzi, C. Zhao, C. Kamoun, and H. Heinrich. J. Appl. Polymer Sci. 80(12):
2128–2139 (1999).
45. S. Garnier and A. Pizzi, J. Appl. Polymer Sci. 81: 3220 (2001).
46. M. Properzi, A. Pizzi, and L. Uzielli, J. Appl. Polymer Sci. 81: 2821 (2001).
47. New Zealand Forestry Res. Inst., Greenweld, US Patent 5,674,338 (1997).
48. M. Properzi, A. Pizzi, and L. Uzielli, Honeymoon MUF adhesives for exterior grade glulam,
Holz Roh Werkstoff 59 413 (2001).
49. M. Properzi, A. Pizzi, and L. Uzielli, Holzforschung Holzverwertung 53: 114 (2001).
50. G. Widmer, Paint Oil Chem. Rev. 112: 18, 26, 28, 30, 32 (1949); Kunststoffe 46(8): 359 (1956).
51. C. Hirt, F. T. King, and R. G. Schmitt, Anal. Chem. 26(8): 1273 (1954).
52. R. W. Stafford, Paper Trade J. 120: 51 (1945).
53. H. Schindlebauer and J. Anderer, Angew, Makromol. Chem. 79: 157 (1979).
54. B. Tomita and H. Ono, J. Polymer Sci. Chem. Ed. 17: 3205 (1979).
55. L. A. Panamgama and A. Pizzi, J. Appl. Polym. Sci. 55: 1007 (1995).
56. B. Tomita and C. -Y. Hse, Mokuzai Gakkaishi 39(11): 1276 (1993).
57. B. Tomita, M. Ohyama, A. Itoh, K. Doi, and C.-Y. Hse, Mokuzai Gakkaishi 40(2): 170 (1994).
58. Y. Yoshida, B. Tomita, and C.-Y. Hse, Mokuzai Gakkaishi 41(6): 547 (1995).
59. Y. Yoshida, B. Tomita, and C.-Y. Hse, Mokuzai Gakkaishi 41(6): 555 (1995).
60. Y. Yoshida, B. Tomita, and C.-Y. Hse, Mokuzai Gakkaishi 41(7): 652 (1995).
61. B. Tomita and C.-Y. Hse, J. Polym. Sci., Polym. Chem. 30: 1615 (1992).
62. V. M. L. J. Aarts, M. L. Scheepers, and P. M. Brandts, Analysis of MF resins , Proceedings of
1995 European Plastic Laminates Forum, Heidelberg, Germany, 1995, pp. 17–25.
63. T. A. Mercer and A. Pizzi, J. Appl. Polymer Sci. 61(9): 1687 (1996).
64. T. A. Mercer and A. Pizzi, J. Appl. Polymer Sci. 61(9): 1697 (1996).
65. C. Kamoun and A. Pizzi, Holz Roh Werkstoff 58(4): 289 (2000).
66. Y. Laigle, C. Kamoun, and A. Pizzi, Holz Roh Werkstoff 56(3): 154 (1998).
67. C. Zhao, S. Garnier, and A. Pizzi, Holz Roh Werkstoff 56(6): 402 (1998).

Copyright © 2003 by Taylor & Francis Group, LLC


68. A. Pizzi, J. Appl. Polymer Sci. 63: 603 (1997).
69. A. Pizzi, J. Appl. Polymer Sci. 65: 1843 (1997).
70. C. Kamoun, A. Pizzi, and R. Garcia, Holz Roh Werkstoff 56(4): 235 (1998).

Copyright © 2003 by Taylor & Francis Group, LLC

View publication stats

Potrebbero piacerti anche