Вы находитесь на странице: 1из 22

6

Resorcinol–Formaldehyde Resins
and Hydroxymethyl Resorcinol
(HMR and n-HMR)

6.1 Introduction
Resorcinol or meta-hydroxy benzene is a phenol with the following
structure:
OH

HO

The reaction of resorcinol and formaldehyde (RF) is much more rapid


than the reaction between phenol and formaldehyde. It is known that
the base- or acid-catalyzed resorcinol–formaldehyde (RF) reactions can
lead to polymeric resins, which are currently used as wood adhesives and
composites. RF phenol–resorcinol–formaldehyde (PRF) resin can cure at
ambient temperatures. Thus, they have an advantage over the plain phenol–
formaldehyde (PF) resin. Hence RF and PRF resins are ideally suited for
producing exterior grade glue-lam and finger joints. It has to be men-
tioned that pure RF resins are generally not used as wood adhesive due to
their high cost. To limit such costs, resorcinol is co-reacted with PF resols,
obtaining PRF co-condensed resins. The reaction between resorcinol and
formaldehyde can lead to the formation of linear polymer if the ratio of
resorcinol to formaldehyde is 1:1 or less.
It is to be noticed that the C4 and C6 sites of the aromatic ring of resorci-
nol are highly activated towards electrophilic substitution due to their high
electron density [1]. The C2 site, although strongly activated, suffers from
the steric hindrance induced by the two vicinal hydroxyl groups. The pro-
portion of the methylene (-CH2-) linkages in C4 plus C6 relative to those
in C2 is in the proportion 10.5:1 [2].

R. N. Kumar and A Pizzi. Adhesives for Wood and Lignocellulosic Materials, (147–168) © 2019
Scrivener Publishing LLC

147
148 Adhesives for Wood and Lignocellulosic Materials

Due to the similarity of chemical structure of resorcinol to phenol, it


is natural to expect that the reaction products of resorcinol and formal-
dehyde are similar to those of phenol and formaldehyde in that methylol
resorcinols are first formed with the methylol groups in the o- and
p-positions initially. These methyolol derivatives can undergo further con-
densation to form dimers and oligomers in the same manner as their phe-
nolic counterparts though reactions between (1) the two hydroxymethyl
groups and/or (2) hydroxymethyl group and the hydrogen atom of the
unsubstituted o- or p-positions of another resorcinol molecule to form
methylene bridges.

6.2 Reaction between Resorcinol and Formaldehyde


The reaction between resorcinol and formaldehyde under alkaline condi-
tions leads to the formation of methylol groups (hydroxymethyl groups)
and, by further conversion, to quinone methides (QMs). This is given in
Figure 6.1.
QMs are the intermediates for subsequent chain extension and cross-
linking reactions similar to the reaction between phenol and formaldehyde.
From a kinetic point of view, the initial reaction to form dimers is much
faster than the subsequent condensation of these dimers and higher oligo-
mers. The reaction of resorcinol with formaldehyde on an equal molar
basis proceeds at a rate that is approximately 10 to 15 times that of the PF

OH O
O=
O=

CH2O
:
OH CH2O H
HO HO HO HO

OH
O=

O=

CH2 CH2OH CH2OH


=

HO HO
HO
Quinone methide

Figure 6.1 Reaction of resorcinol with formaldehyde through its different methylol
derivatives up to quinone methides.
RF Resins and Hydroxymethyl Resorcinol 149

system. Such a high reactivity of the RF system renders it impossible to


produce these adhesives in resol form. Hence, only the novolak-type resins
without free methylol groups are produced. In the novolak resin, all the
resorcinol nuclei are linked by methylene bridges with no methylol groups
or methylene–ether linkages [2].
In general, for the formulation of PRF resins, the formulator tries to
achieve the best performance with the minimum proportion of the very
expensive resorcinol. The industrial PRF resins have evolved with respect
to cost/performance in such a way that a general proportion of resorcinol
is between 16% and 18% of the total resin solids. In general, for the lower-
molecular-weight PF resins in which resorcinol needs to be grafted, the
proportion is around 1.15–1.2 molar resorcinol for 1 molar PF trimer. A
lower proportion may lead either to long linear polymers, or if even lower
may result in gelling of the resin. A higher proportion of resorcinol would
be an unnecessary addition of cost.
For a more detailed information on RF, PR, urea–resorcinol–
formaldehyde (URF), tannin–resorcinol–formaldehyde (TRF), and lignin–
resorcinol–formaldehyde (LRF), readers should refer to previous reviews [2].

6.3 Comparison between Resorcinol and Phenol


Significant difference exists between PF and RF resins in that the curing
of PF resin generally requires a hot-press temperature of above 160°C,
whereas RF and PRF cure at ambient temperature; i.e., they are cold-
setting resins. This indicates that RF condensations present a much lower
energy barrier and occur at much faster rate even at room temperature.
Superficially, higher reactivity may be attributed to the stronger nucleo-
philicity of resorcinol since it has two phenolic hydroxyl groups, which
lead to higher electron density on the aromatic ring.
Proposal of a suitable mechanism at the molecular level requires an
understanding of the formation of reactive intermediates since they deter-
mine the rate of the overall condensation reaction [3].
Thus, the comparison of the kinetics of the formation of intermediates
during the reactions between PF and RF will throw light on the difference
in the mechanism of the two reactions. However, the experimental kinetic
studies seem to be very difficult since the capture and quantitative eval-
uation of such short-lived intermediates are currently almost impossible
especially for solution reactions. Theoretical approach instead is expected
to give better clarity on the complex reactions involved. In the field of
wood resin adhesives, pioneering investigations employing a molecular
150 Adhesives for Wood and Lignocellulosic Materials

mechanics approach are on record [4–8]. These studies have focused on the
interactions between resin components and wood cellulose, with a view to
establish theoretical models for resin–wood surface interaction, thereby giv-
ing insight into the relationship between resin structure and performance.
Molecular mechanics is a technique based on force field theory and
mainly deals with all the secondary interactions such as hydrogen bond-
ing, van der Waals and electrostatic forces, and their influences on molec-
ular energy and conformations.
Li et al. on the other hand consider quantum chemistry calculations
to be necessary for analyzing the resorcinol–formaldehdye condensation
reaction [3]. This method is based on quantum mechanics and deals with
formation and breakage of chemical bonds, kinetics, and thermodynam-
ics of chemical reactions through rigorous calculations on the electronic
structures of the involved species.
By employing a quantum chemistry method, Li et al. recently studied
the base-catalyzed PF reactions. The QM was confirmed to be the key
intermediate that initiates condensations [9].
RF condensation reactions may share the same mechanism, but the
chemistry of RF resins is obviously more complex than that of the PF one.
Resorcinol has two hydroxyl groups and therefore dissociation of the pro-
tons can produce a singly charged or doubly charged anion (dianion) in
the presence of a base.
As a result, different QM intermediates may be formed.

OH O O
O
H2C CH2

HO HO HO HO
R 4 QM
2 QM CH2
6 QM

O O
H 2C CH2

O O
2 QMA 4 QMA

Figure 6.2 Possible quinone methide intermediates formed by reaction of resorcinol with
formaldehyde.
RF Resins and Hydroxymethyl Resorcinol 151

As can be seen in Figure 6.2, five possible intermediates may be formed


according to the proposed mechanism. 2-QM, 4-QM, and 6-QM are neu-
tral species that are similar to phenol QMs. However, 2-QMA and 4-QMA
are anions that may exhibit different reactivity.
According to the results for PF reactions, the formation of the QM is
the rate-determining step of the overall condensation reaction [9]. Thus,
it can be speculated that formations of RF QMs encounter lower energy
barriers to the reaction of condensation.

6.4 Reactive Positions and Types of Linkages


Comparison between Resorcinol and Phenol
Theoretically, resorcinol has three reactive sites on the ring, namely,
the three sites in ortho to the two hydroxyl groups. Thus, three types of
methylene linkages through 2,2’-, 2,4-, and 4,4’-condensations should be
formed. However, it is interesting that some 13C NMR studies indicated
that the 2,4’- and 4,4’-linkages were found to be dominant, whereas the
2,2’-linkage was observed to be minor or absent [10–13].
Further, the 4,4’-linkage appeared to be more favorable than
2,4’-linkages. These observations imply that the C4-site is more reac-
tive than the C2-site.
The results of the theoretical calculations of Li et al. showed that the
formation of 6-QM and 4-QMA have relatively lower energy barriers [9].
This rationalized previous experimental observations that the 2,4’- and
4,4’-methylene linkages were dominant, whereas the 2,2’-linkage was
almost absent.

6.5 Hydroxymethyl Resorcinol


6.5.1 Introduction
In 1995, a wood primer was developed, which mitigated the damaging
effects caused by stringent simulated weathering. This chemical treat-
ment is known as hydroxymethyl resorcinol (HMR) (Figure 6.3). HMR is
a 5% solids, dilute alkaline solution of resorcinol and formaldehyde, and
is currently the most well-known and effective wood primer available; it
remarkably enhances the moisture durability of wood–adhesive bonds.
The treatment is suggested to enhance the dimensional stability of wood,
152 Adhesives for Wood and Lignocellulosic Materials

OH
OH
HOH2C CH2OH
+ HCHO
NaOH OH
OH
CH2OH
Resorcinol Trihydroxymethyl resorcinol

OH
CH2 CH2OH

HO OH
CH2
HO

OH

Hydroxymethyl resorcinol trimer

Figure 6.3 Reaction of resorcinol with formaldehde to form HMR trimer.

and its effectiveness is attributed to its ability to cross-link within the


wood cell wall [14–17].
Both heat curable phenolic resin adhesives as well as room temperature
curable RF adhesives have produced durable bonds with wood for decades.
However, these resins do not form bonds of adequate durability in the case
of the following circumstances [18]:

(1) In the case of bonding wood treated with preservatives


such as chromated copper arsenate (CCA)
(2) In the case of bonding composites of wood with non-
wood materials such as plastics or metals that may require
epoxies or polyurethane adhesives
(3) In the case of bonding wood with epoxy resins to conform
to the requirements for resistance to delamination as spec-
ified in ASTM D2559 [19]

The above problems were addressed by researchers at the USDA


Forest Service, Forest Products Laboratory, Madison, WI, who first
RF Resins and Hydroxymethyl Resorcinol 153

published results on HMR as a new coupling agent for wood products


[20–22].

6.5.2 Normal HMR


HMR was disclosed in 1995 by Vick and co-workers at the United States
Forest Products Laboratory [21]. Ever since its introduction in the mid-
1990s, HMR gained in reputation among the wood adhesion community
as an effective coupling agent in producing durable bonds under stringent
exterior conditions for a number of resin-adhesive systems, although it
was originally developed for use with epoxies. Subsequently, the appli-
cation was extended to other systems such as emulsion polymer isocy-
anate, polymeric diphenylmethane diisocyanate (pMDI) [23], and PRF
resin [20, 22], for bonding wood to unsaturated polyester-based fiber-
reinforced plastic (FRP). HMR also promotes the exterior durability of
the joints bonded with adhesives such as epoxy, PRF resin [23], and vinyl
ester [24].
HMR was found to dramatically improve the moisture durability of
epoxy bonds to Sitka spruce [21]. The effectiveness of the HMR was
established since it conforms to the requirements of the ASTM D 2559
[19] delamination test. This procedure was a critical hurdle towards
the structural certification of wood adhesives, involving a harsh cyclic
water-weathering process. With no HMR treatment, bisphenol-A epoxy
adhesives failed the ASTM test, while HMR-treated samples passed this
rigorous test.
HMR has also been found to promote bonding of preservative-treated
wood, an extremely difficult substrate for adhesive bonding. Vick et al.
found that the HMR treatment significantly increased internal bond
strength of flakeboards made with CCA-treated wood flakes and a PF
adhesive [22]. However, the HMR treatment showed no effect on the
mechanical properties of flakeboards made with preservative-free flakes
[22]. Christiansen and Conner [25] found that the HMR treatment
enhanced the durability of bonds between CCA-treated lumber and epoxy,
pMDI, EPI, and PRF as confirmed by the conformity to the requirements
of the ASTM D 2559 delamination test. HMR was found to dramatically
improve the moisture durability of epoxy bonds to Sitka spruce [21].
HMR resin was found to function effectively as coupling agent between
wood and for bisphenol-A epoxy adhesive. It is of noteworthy signifi-
cance that the bonded wood joint passed the harsh ASTM D 2559 cyclic
water-weathering delamination test [19].
154 Adhesives for Wood and Lignocellulosic Materials

6.5.3 Formulation of HMR


The HMR coupling agent is applied to the wood surface as an aqueous
solution with low solids content (5%) and is analogous to a low-molecular-
weight RF resin. It contains four main components, which are resorcinol,
formaldehyde, sodium hydroxide, and water. The Ingredients of original
HMR formulation are given in Table 6.1.

6.5.3.1 Mixing Procedure


The original HMR coupling agent is prepared at the time of use by mixing
resorcinol with formaldehyde [formaldehyde-to-resorcinol (F/R) molar
ratio of 1.54] at pH 9 in a 5% solids content aqueous solution at room
temperature. This version of HMR becomes most effective 3 to 4 h after
mixing, and it can be used for up to about 8 h. The duration of reaction
time between preparing the solution and applying it to a wood surface
strongly affects the molecular weight distribution and residual reactivity
of HMR. Reaction times either shorter or longer than the optimum range
result in bonds that are less water resistant to delamination. A fresh mix
must be prepared for each new batch of HMR.
For wood-to-wood bonding using epoxy adhesives, the original ver-
sion of HMR was used by Vick et al. [21], Vick and Okkonen [26], and
Vick et  al. [27]. Accordingly, the ingredients in Table 6.1 are allowed to
react for 4 h at room temperature. After 4 h, the solution is applied onto a
freshly planed wood surface. Because of the amount of water contained in
the HMR solution (95%), the HMR-treated surface needs to dry out prior
to adhesive application. During drying, water evaporates from the wood
surface while the resorcinol and formaldehyde react. A drying time of 18
to 24 h is used under ambient conditions to bond non-aqueous adhesives,
such as epoxy resins [21].

Table 6.1 The ingredients of the original HMR formulation. F/R molar
ratio = 1.54.
Ingredient Percentage
Resorcinol (crystalline) 3.34
Water deionized 90.43
Sodium hydroxide 3 molar solution (10.8% by weight) 2.44
Formaldehyde solution 3.79
RF Resins and Hydroxymethyl Resorcinol 155

6.5.3.2 Limitations to the Use of HMR


Three main obstacles to the commercial use of HMR became apparent:

(1) HMR has no storage life. Thus, every batch has to be mixed
on-site from accurately measured proportions of starting
chemical, which is time-consuming and prone to error.
Each batch has only 3 to 4 h of storage life.
(2) The second obstacle was another time constraint, namely,
a waiting period of 3 to 4 h before the material can be used,
i.e., a duration almost equal to the storage life of the mix.
This makes HMR commercially cumbersome to use, too.
Care of freshly HMR-treated wood is required, because
contamination with, e.g., dust or chemical vapor can
decrease the effectiveness of the HMR.
(3) Furthermore, there was a need to reduce the reaction time
of the HMR solution.

6.6 Novolak-Based HMR


In an effort to make HMR more versatile, Christiansen et al. developed
the so-called “novolak-HMR” (n-HMR) [18, 28]. This is a stable novolak
prepolymer, in which formaldehyde and resorcinol are mixed at a low
mole ratio (0.39 F:R); at this stage, storage life is essentially infinite [18].
Prior to wood application, the mole ratio is increased to 1.54 [18]. Once
the additional formaldehyde is added, the n-HMR solution can be applied
to the wood substrate immediately and up to 7 h later [18]. Weather dura-
bility tests using n-HMR treatment and an epoxy adhesive yielded results
similar to those of Vick et al. [21]. The term “novolak” signifies the fact that
it is a two-stage resin system. In the first stage, the mole ratio of F/R is 0.39,
and in the second stage, the ratio is increased to 1.54 and the final cure is
achieved by a second addition of formaldehyde. Christiansen’s use of the
term novolak reflects the two-stage curing process, but n-HMR is prepared
under alkaline conditions.
As described above, the n-HMR system is carried out in two stages:

(1) The first stage is the preparation of an n-HMR solution.


(2) The second stage is the preparation of the actual coupling
agent by adding additional formaldehyde to n-HMR. This
is called the “activated stage” [18].
156 Adhesives for Wood and Lignocellulosic Materials

Christiansen et al. tested different F/R ratios from 0.23 to 0.46 for the
novolak stage [14, 18, 21]. In this work, an F/R ratio of 0.39 was used.
By using the n-HMR, the advantage is that the coupling agent can be
prepared in the factory and then shipped it to a wood-bonding plant where
the novolak would simply be mixed with the final formaldehyde.

6.6.1 Preparation of n-HMR


The mixing procedure used for n-HMR at 5% solids content is described
by Christiansen et al. [14, 18, 21].

6.7 Bonding Mechanism Using HMR


6.7.1 Mechanism based on the Material Properties of HMR
HMR was designed to be a low-solids-content, low-molecular-weight cou-
pling agent that could penetrate the wood cell wall at the molecular level
[29]. The low molecular weight nature of HMR solution was confirmed by
Vick et al. [27].
HMR in its reactive, application state mostly occurs as monomers,
dimers, and trimers that can penetrate the wood cell wall because the
molecular weight of these HMR molecules is less than 1000 [27].
Based on theoretical calculations of solubility parameters for wood cell
wall components and HMR, it is postulated that HMR most likely will
associate with lignin because of their similar solubility parameters [30].
Protonated HMR has a solubility parameter of 27.5 (J1/2 cm3/2) while lignin,
hemicellulose, and cellulose have solubility parameters of 31.1, 36.3, and
38.6 (J1/2 cm3/2), respectively. Further support for the preference of HMR
for lignin is the impact of HMR on possibly lowering the glass transition
temperature (Tg) of lignin in dynamic mechanical analysis (DMA) experi-
ments on HMR-soaked wood veneers [30].
In order to impart a durable wood–adhesive bond, HMR should form
a cross-linked structure at the cell wall. To prove the importance of the
cross-link density, Christiansen employed substituted resorcinol, namely,
2-methyl resorcinol, to decrease the cross-link density [14]. This substitu-
tion resulted in decreased wood–epoxy bond durability. By virtue of its low
molecular weight, HMR can diffuse into the cell wall to form an interpen-
etrating polymer network.
RF Resins and Hydroxymethyl Resorcinol 157

6.7.2 Mechanism Based on Surface Chemistry


Two primary means to measure surface chemical characteristics of wood
are X-ray photoelectron spectroscopy (XPS) and contact angle analysis. XPS
provides a measure of the elemental composition of the wood surface includ-
ing carbon and oxygen percentages as well as the carbon oxidation state [29].
Contact angle analysis can be used to probe the surface energetics (polar
and dispersive character) as well as adhesive wettability of wood [31, 32].
This XPS result correlates well with the surface energetics, where it was
found that HMR-treated wood reduced the dispersive (nonpolar) charac-
ter, but increased the polar character of the wood surface.
Upon treatment with HMR, treated wood surfaces were found to exhibit an
increase in non-oxidized carbons and a decrease in carbon–oxygen bonds [32].
Any carbonyl or carboxyl functionality remained unaltered or slightly
decreased compared to untreated wood. However, despite the increase in
non-oxidized carbon and the decrease in carbon–oxygen ratios, the oxygen/
carbon (O/C) ratios either remain unaffected or become higher. This XPS
result correlates well with the surface energetics, where it was found that
HMR-treated wood reduced the dispersive (nonpolar) character, but
increased the polar character of the wood surface [31]. The enhanced polar
interaction can be attributed to the hydroxymethyl groups of the HMR.

6.8 Applications of HMR and n-HMR


6.8.1 Bonding to Preservative-Treated Wood
It is known that bonding of wood treated with preservatives has been very
difficult. One of the important preservatives used for wood preservation is
CCA. In the past, it was essentially impossible to both treat wood and then
adhesively bond it to make bigger assemblies, even with durable PRF adhe-
sives since CCA preservatives interfere with bond formation of phenolic-
based adhesives on CCA-treated southern pine. Cr+3 ions and Cu+2 ions
are known to form a complex with phenol and formaldehyde and to affect
the rate of cure of the PF resin [33–36], thereby detracting from achieving
optimum bond strength. However, if CCA-treated southern pine is primed
with HMR, PRF adhesives could produce durable bonds that would meet
the requirements of the ASTM D 2559 delamination test [20]. Without the
primer, the PRF adhesive could meet the requirements. In the past, the
American Institute of Timber Construction also did not certify southern
158 Adhesives for Wood and Lignocellulosic Materials

pine lumber for glulam if the lumber had previously been treated with
CCA until HMR was also used as primer [37].
HMR was also used to make flakeboards from CCA-treated wood [23].
HMR treatment significantly increased internal bond strength of flake-
boards made with CCA-treated wood flakes and a PF adhesive. However,
the HMR treatment showed no effect on the mechanical properties of
flakeboards made with preservative-free flakes [22]. Christiansen and
Conner [25] also found the HMR treatment to enhance the durability of
epoxy, pMDI, EPI, and phenol–resorcinol–formaldehyde (PRF) bonds to
CCA-treated lumber in the ASTM D 2559 delamination test [17].
Alkaline copper quat (ACQ) and copper azole (CA-B), the most prom-
inent substitutes for CCA, are difficult to bond consistently using a PRF
adhesive formulated for bonding to CCA-treated wood.
The bond performance of ACQ- and CA-B-treated wood primed with
n-HMR was improved. Delamination of the PRF-bonded CA-B-treated
southern yellow pine samples was found to be reduced in the ASTM
delamination test [38].

6.8.2 Epoxy–Wood Adhesion


Epoxy resins develop strong bonds to wood in the dry state. Once they
are exposed to repeated soaking in water and drying, the bonds fail to
meet requirements of exterior grade structural wood adhesive. No epoxy
adhesives are known to meet the requirements of ASTM Specification D
2559 [19]. If wood surfaces were coated with HMR before bonding with
bisphenol-A epoxy adhesives, the durability of the bond was dramatically
increased in comparison with the bonding without the coupling agent [21].

6.8.3 Bonding of Fiber-Reinforced Polymer–Glulam Panels


In recent years, wood or wood panels are structurally upgraded by lami-
nation with a thinner layer of FRP composite to give higher strength and
defect-free surface for application as architectural beams and bridge mem-
bers. Such composites also provide a chemically resistant surface. HMR
can function as coupling agent between wood and FRP. For instance, HMR
was found to be effective in bonding E-glass/vinyl ester resin-based FRP
and eastern hemlock glulam panels.
Fiber-reinforced polymers (FRPs) for glulam reinforcement have been
thoroughly researched due to their light weight and high tensile strength
properties [17, 24, 39]. Such glulam panels approximately 4’ wide by 6”
thick have been employed in timber girder bridges [40].
RF Resins and Hydroxymethyl Resorcinol 159

Lopez-Anido et al. investigated the use of an E-glass/vinyl ester resin as


a means of reinforcing eastern hemlock glulam panels [24]. HMR was used
to improve the durability of the wood/vinyl ester resin bond.
The ASTM D 1101 (1997) [41] cyclic delamination test was used to test
the durability of the wood–FRP laminates [17, 24]. The HMR treatment
was found to promote strong, exterior-grade bonds between the wood and
FRP. These bonds were comparable in strength and durability to PRF adhe-
sive bonds; PRF adhesives are commonly considered as the most durable
structural wood adhesives [17, 24].
The wood–FRP composite requires bonding high-polarity wood to fairly
low-polarity thermoset “plastic”. In this application, HMR has proven to be
very useful [23, 42]. Though the PRF adhesive bonds well to wood, it can-
not maintain a durable bond to vinyl ester FRP. However, epoxy bonds well
to HMR-primed wood and to vinyl ester FRP.
Eisenheld employed IR heat for accelerating the drying of n-HMR-
treated wood before making the vinyl ester–glass–wood composites [43,
44]. The standard HMR drying procedure is 24 h at 23 ± 2°C. The rein-
forcement method adopted in the work was SCRIMPTM (an acronym that
stands for Seemann Composites Resin Infusion Molding Process), a resin
transfer-molding process.

6.8.4 Priming Agent for Bondability of Wax-Treated Wood


Various pretreatments have been used to improve the bonding proper-
ties of inactivated/hydrophobic wood surfaces. Among these were cou-
pling agents, bio-products, such as enzymes (xylanases), or other chemicals,
such as tris (polyoxyethylene) sorbitan monooleate [45], sodium hydroxide
(NaOH), calcium hydroxide, nitric acid, hydrogen peroxide, and borax [46].
Kurt et al. employed HMR priming wax-treated wood prior to bonding
with PVAc and MF and reported that it increased the shear strength of
PVAc-bonded specimens under wet conditions and of MF-bonded speci-
mens under dry and wet conditions [47]. The effect on MF-bonded speci-
mens, however, was much more pronounced under wet conditions.

6.9 Special Adhesives of Reduced Resorcinol Content


6.9.1 Fast-Setting Adhesive for Fingerjointing and Glulam
Together with the more traditional fingerjointing adhesives, a series of
ambient temperature fast-curing separate application systems have also
been developed. These eliminate the long delays caused by the use of more
160 Adhesives for Wood and Lignocellulosic Materials

conventional PRF adhesives, which require lengthy periods to set. These


types of resorcinol adhesives are applied separately. They were first devel-
oped in the United States to bond large components where presses were
impractical [48, 49]. Kreibich describes these separate application or
“honeymoon” systems as follows [49]:
Component A is a slow-reacting PRF resin with a reactive hardener.
Component B is a fast-reacting resin with a slow-reacting hardener.
When A and B are mated, the reactive parts of the component react within
minutes to form a joint that can be handled and processed further. Full
curing of the slow-reacting part of the system takes place with time. The
m-aminophenol used for component B is a frightfully expensive chemi-
cal, and for this reason, these systems were discarded and not used indus-
trially [50]. In their original concept, component A is a traditional PRF
cold-setting adhesive at its standard pH of between 8 and 8.5 to which
formaldehyde hardener has been added. Flour fillers may be added or
omitted from the glue mix. Component B is a phenol/meta-aminophenol/
formaldehyde resin with a very high pH (and therefore a high reactivity),
which contains no hardener or only a very slow hardener.
More recently, a modification of the system described by Kreibich has
been used extensively in industry with good success [51, 52]. Part A of the
adhesive is again a standard PRF cold-setting adhesive with powder hard-
ener added at its standard pH. Part B can be either the same PRF adhesive
with no hardener and the pH adjusted to 12 or a 50% to 55% tannin extract
solution at a pH of 12–13, provided that the tannin is of the condensed or
flavonoid type, such as mimosa, quebracho, or pine bark extract with no
hardener [50, 52]. The results obtained with these two systems are good
and the resin not only has all the advantages desired but also is consider-
ably cheaper as a result of the use of vegetable tannins and of the halving of
the resorcinol content of the entire adhesive system [50, 52, 53].
The adhesive works in the following manner. Once the component A
glue mix is spread on one fingerjoint profile and component B is spread
on the other fingerjoint profile and the two profiles are joined under pres-
sure, the reaction of component B with the hardener of part A is very fast.
In 30 min at 25°C, fingerjoints prepared with these adhesives generally
reach the levels of strength that fingerjoints glued with more conventional
phenolic adhesives are able to reach only after 6 h at 40 to 50°C or in 16
to 24 h at 25°C [50, 53]. Clamping of laminated beams (glulam) bonded
with these fast set honeymoon adhesives is in average of only 3 h at ambi-
ent temperature compared with the 16 to 24 h necessary with traditional
PRF resins [52, 53]. These adhesives also present two other advantages,
namely, (i) they are capable to bond without any decrease of performance
RF Resins and Hydroxymethyl Resorcinol 161

at temperatures down to 5°C only and (ii) they are able to bond “green”
timber at high moisture content, a feat that is used in industrial glulam
bonding since their commercial introduction in 1981. Several variations
on the theme exist, such as the “Greenweld” system from New Zealand in
which component B is a solution just composed of ammonia as a strong
accelerator of the PRF + hardener of component A and of a thickener; this
system, however, suffers from the presence of the odor of ammonia, which
is unacceptable in some sophisticated markets.

6.9.2 Branched PRF Adhesives [11, 12, 54]


Recently, another step forward has been taken in the formulation of PRF
adhesives of lower resorcinol content. Liquid RF resin or PRF resins appear
to be mostly linear [55]. The original concept in “branching” erroneously
maintained that if a chemical molecule capable of extensively branching
(three or more effective reaction sites with an aldehyde) the PF and PRF
resins is used after, before, or during, but particularly during or after, the
preparation of the PF resin, the polymer in the branched PRF adhesive
has (1) higher molecular weight than in normal PRF adhesives where
branching is not present and (2) higher viscosity in water or water/solvent
solutions of the same composition and of the same resin solids content
(concentration). It also needs a much lower resorcinol amount on total
phenol to present the same performance of normal linear PRF adhesives
[11, 12, 54, 55]. This can be explained schematically as follows:

Resorcinol −CH2 phenol−CH2 n resorcinol

Resorcinol −CH2 phenol−CH2 n resorcinol

Resorcinol −CH2 phenol−CH2 n resorcinol

n in integer numbers

Resorcinol CH2 − phenol n Phenol − CH2 Resorcinol


CH2 CH2 n

Branching
molecule

CH2
Phenol
CH2
n
Resorcinol
162 Adhesives for Wood and Lignocellulosic Materials

with n > 1 and an integer number and comparable to, similar to, or equal to
n in the preceding scheme for the production of PRF resins.
When comparing linear and branched resins, for every n molecules of
phenol used, a minimum of two molecules of resorcinol are used in the case
of a normal, traditional linear PRF adhesive, whereas only one molecule of
resorcinol for n molecules of phenol is used in the case of a “branched” PRF
adhesive. The amount of resorcinol has then been halved or approximately
halved in the case of the branched PRF resin. A second effect caused by the
branching is a noticeable increase in the degree of polymerization of the
resin. This causes a considerable increase in the viscosity of the liquid adhe-
sive solution. Because PRF adhesives must be used within fairly narrow vis-
cosity limits, to return the viscosity of the liquid PRF adhesive within these
limits, the resin solids content in the adhesive must be lowered, consider-
ably, with a consequent further decrease in total liquid resin of the amount
of resorcinol and of the other materials, except solvents and water. This
decreases the cost of the resin further without decreasing its performance.
Thus, to conclude, the decrease of resorcinol by branching of the resin
is based on two effects:

1. A decrease of resorcinol percentage in the polymer itself,


hence in the resin solids, due to the decrease in the num-
ber of the PF terminal sites onto which resorcinol is grafted
during PRF manufacture.
2. An increase in molecular weight of the resin, which, by the
need to decrease the percentage resin solids content to a
workable viscosity, decreases the percentage of resorcinol on
liquid resin (not on resin solids).

It is clear that, in a certain sense, a branched PRF will behave as a more


advanced, almost precured phenolic resin. While the first effect described is a
definite advance on the road to better engineered PRF resins, the second effect
can also be obtained with more advanced (reactionwise) linear resins. The con-
tribution of the second effect to the decrease in resorcinol is not less marked
than that of the first effect. It is, however, the second effect that accounts for
the difference in behavior between branched and linear PRF adhesives.
Branching molecules that can be used could be resorcinol, melamine,
urea, and others [56]. Urea is the favorite, because it is much cheaper than
the others and needs to be added in only 1.5% to 2% of total resin. When
urea is used as a brancher, the adhesive assumes an intense and unusual
(for resorcinol resins) blue color, after a few days, hence its nickname, “blue
glue”. However, later work has shown that tridimensional branching has
RF Resins and Hydroxymethyl Resorcinol 163

very little to do with the improved performance of these low-resorcinol-


content adhesives, with tridimensionally branched molecules contributing,
at best, no more than 8% to 9% of the total strength [11, 12, 54, 56]. In real-
ity, addition of urea causes the reaction as foreseen, but not in three points
branching but rather only in two sites of the branching molecule. This is
equivalent to saying that most of the resin doubles linearly in molecular
weight and degree of polymerization, while the final effect, good perfor-
mance at half the resin resorcinol content, is maintained [11, 12, 54, 56].
This effect is based on the relative reactivity for phenolic methylols of urea
and of unreacted phenol sites and thus while the macro effect is as wanted,
at the molecular level, it is only a kinetic effect due to the different relative
reactivities of urea and phenol under the reaction conditions used. Thus,

resorcinol–CH2–[–phenol-CH2–]n–resorcinol
resorcinol–CH2–[–phenol-CH2–]n–resorcinol

But

resorcinol–CH2–[–phenol–CH2–]n – UREA – CH2–[–phenol–


CH2–]n –resorcinol

HaIving of the resorcinol content is still obtained, but between 90% and
100% of the polymers in the resin are still linear.
It is noticeable that the same degree of polymerization and “doubling”
effect cannot be obtained by lengthening the reaction time of a PF resin
without urea addition [11, 12, 54, 56]. These liquid resins then work at a
resorcinol content of only 9% to 11%, hence considerably lower than that
of traditional PRF resins. These resins can also be used with good results
for honeymoon fast-setting adhesives in PRF tannin systems, thus further
decreasing the total content of resorcinol in the total resin system to a level
as low as 5% to 6%. This concept was extended to URF cold-setting adhe-
sives, these too giving good results [12].

6.9.3 Cold-Setting PF Adhesives Containing No Resorcinol


As the cost of cold-setting, exterior grade adhesives based on resorcinol is
very high due to the high cost of resorcinol itself, the tendency to decrease
the amount of resorcinol while maintaining unaltered the performance of
the adhesive, when brought to its ultimate conclusion, leads to the concept
of exterior cold-setting phenolic adhesives of zero-level resorcinol. As alka-
line PF resins have not an ambient temperature rate of reaction, which is
164 Adhesives for Wood and Lignocellulosic Materials

even vaguely sufficient to set and harden to a sufficient level the adhesive,
some modifications need to be introduced to overcome in this regard the
lack of resorcinol. This can be done in several manners: (i) by using stan-
dard PF thermosetting resol resins and hardening them by increasing the
glue line temperature by radiofrequency in fingerjointing and glulam man-
ufacture. The system is expensive and needs considerably higher capital
outlay and more careful handling of both the equipment and of the joint,
for results that are certainly not particularly exciting. (ii) By using resins
in which the PF resol of adhesives of type 2 above is terminated by the
terminal grafting of a resorcinol substitute, for example, a natural polyfla-
vonoid tannin, this system being truly cold-setting and yielding relatively
good results but at best just on the inferior limit of the standard require-
ments [57]. (iii) By using self-neutralizing acid setting PF resols. The term
“acid-setting” when used in the presence of a lignocellulosic substrate
makes wood technologists shudder, conjuring visions of extensive acid-
induced substrate degradation and early exterior joint failure. And this is
indeed the case! In reality, some exterior aminoplastic resins do harden in
the moderately acid range without any major substrate degradation prob-
lems. PF resins, however, while hardening very rapidly under acid condi-
tions, do need very acid conditions to give a hardened strong network, and
this elevated acidity is not really acceptable as regards long-term durabil-
ity of the substrate. The damage due to the acid hydrolysis of cellulose and
other wood carbohydrates is particularly aggravated and compounded by
the long-term effect of the glue line remaining acid after resin hardening.
However, the main negative effect due to acid-induced degradation of the
substrate has been overcome by using acid-setting PF resins containing
no resorcinol but hardened by the use of a self-neutralizing catalyst [58].
According to this principle, the adhesive first becomes acid to allow the PF
resin to cure, and after hardening, the hardened glue line self-neutralizes in
a very short time [58]. The greater majority of the effects of substrate deg-
radation are then avoided and very strong and durable exterior wood joints
are produced [58]. The system works well in radiofrequency cured joints,
yielding much better results than the alkaline resols of point (i) above, and
can work well under purely cold-setting conditions [58].

References
1. Raff, R.A.V. and Silverman, B.M., Kinetics of the uncatalyzed reactions
between resorcinol and formaldehyde. Ind. Eng. Chem., 43, 1423–1427, 1951.
2. Pizzi, A., Advanced Wood Adhesives Technology, Marcel Dekker, New York, 1994.
RF Resins and Hydroxymethyl Resorcinol 165

3. Li, T., Cao, M., Liang, J., Xie, X., Du, G., Mechanism of base-catalyzed
resorcinol-formaldehyde and phenol-resorcinol-formaldehyde condensa-
tion reactions: A theoretical study. Polymers, 9, 426, 2017.
4. Pizzi, A. and Eaton, N.J.A., A conformational analysis approach to phenol-
formaldehyde resins adhesion to wood cellulose. J. Adhes. Sci. Technol., 1,
191–200, 1987.
5. Pizzi, A. and De Sousa, G., On the resolution of dihydroxydiphenylmethanes
on achiral crystalline Cellulose II—Correlation of experimental and calcu-
lated results. Chem. Phys., 164, 203–216, 1992.
6. Pizzi, A. and Maboka, S., Calculated values of the adhesion of phenol-
formaldehyde oligomers to crystalline Cellulose. J. Adhes. Sci. Technol., 7,
81–94, 1993.
7. Sedano-Mendoza, M., Lopez-Albarran, P., Pizzi, A., Natural lignans adhesion
to cellulose: Computational vs experimental results. J. Adhes. Sci. Technol.,
24, 1769–1786, 2010.
8. Lopez-Albarran, P., Pizzi, A., Navarro-Santos, P., Hernandes-Esparza, R.,
Garza, J., Oligolignols within lignin-adhesive formulations drive their
Young’s modulus: A ReaxFF-MD study. Int. J. Adhes. Adhes., 78, 227–233,
2017.
9. Li, T., Cao, M., Liang, J., Xie, X., Du, G., Theoretical confirmation of the quinone
methide hypothesis for the condensation reactions in phenol-formaldehyde
resin synthesis. Polymers, 9, 45, 2017.
10. Kim, M.G., Amos, W.L., Barnes, E.E., Investigation of a resorcinol-formaldehyde
resin by 13C-NMR spectroscopy and intrinsic viscosity measurement. J. Polym.
Sci., 31, 1871–1877, 1993.
11. Scopelitis, E. and Pizzi, A., The chemistry and development of branched PRF
wood adhesives of low resorcinol content. J. Appl. Polym. Sci., 47, 351–360,
1993.
12. Scopelitis, E. and Pizzi, A., Urea-resorcinol-formaldehyde adhesives of low
resorcinol content. J. Appl. Polym. Sci., 48, 2135–2146, 1993.
13. Christiansen, A.W., Resorcinol-formaldehyde reactions in dilute solution
observed by C13 NMR spectroscopy. J. Appl. Polym. Sci., 75, 1760–1768,
2000.
14. Christiansen, A.W., Chemical and mechanical aspects of HMR primer in
relationship to wood bonding. Forest Prod. J., 55, 73–78, 2005.
15. Son, J. and Gardner, D.J., Dimensional stability measurements of thin wood
veneers using the Wilhelmy plate technique. Wood Fiber Sci., 36, 98–106,
2004.
16. Sun, N. and Frazier, C.E., Probing the hydroxymethylated resorcinol cou-
pling mechanism with stress relaxation analysis. Wood Fiber Sci., 37, 673–
681, 2005.
17. Hosen, J.C., Fundamental Analysis of Wood Adhesion Primers, PhD
Dissertation, Virginia Polytechnic Institute and State University, Blacksburg,
VI, 2010.
166 Adhesives for Wood and Lignocellulosic Materials

18. Christiansen, A.W., Vick, C.B., Okkonen, E.A., A novolak-based hydroxymeth-


ylated resorcinol coupling agent for wood bonding, in: Proceedings, Wood
Adhesives 2000, Session 3B: Advances in Wood Adhesive Formulations, USDA
Forest Service, Forest Products Laboratory, Madison, WI, 2000.
19. ASTM D 2559 Standard Specification for Adhesives for Structural Laminated
Wood Products for Use Under Exterior (Wet Use) Exposure Conditions,
American Society for Testing and Materials, West Conshohocken,
Pennsylvania, USA 2004.
20. Vick, C.B., Coupling agent improves durability of PRF bonds to CCA-treated
southern pine. Forest Prod. J., 45, 3, 78–84, 1995.
21. Vick, C.B., Richter, K., River, B.H., Fried, A.R., Hydroxymethylated resorci-
nol coupling agent for enhanced durability of bisphenol-A epoxy bonds to
Sitka spruce. Wood Fiber Sci., 27, 1, 2–12, 1995.
22. Vick, C.B., Geimer, R.L., Wood, J.E., Jr., Flakeboards from recycled CCA-
treated southern pine lumber. Forest Prod. J., 46, 11/12, 89–91, 1996.
23. Vick, C.B., Hydroxymethylated resorcinol coupling agent for enhanced adhe-
sion of epoxy and other thermosetting adhesives to wood, in Proceedings of
Wood Adhesives 1995, A.W. Christiansen and A.H. Conner (Eds.), pp. 47–55,
Forest Products Society, Madison, WI, 1996.
24. Lopez-Anido, R., Gardner, D.J., Hensley, J.L., Adhesive bonding of eastern
hemlock glulam panels with E-glass/vinyl ester reinforcement. Forest Prod.
J., 50, 11/12, 43–47, 2000.
25. Christiansen, A.W. and Conner, A.H., Hydroxymethylated resorcinol cou-
pling agent for enhanced adhesion of epoxy and other thermosetting adhe-
sives to wood. Proceedings Nr. 7296 USDA Forest Service, Forest Products
Laboratory and Forest Products Society, pp. 47–55, 1995.
26. Vick, C.B. and Okkonen, E.A., Structurally durable epoxy bonds to aircraft
woods. Forest Prod. J., 47, 3, 71–77, 1997.
27. Vick, C.B., Christiansen, A.W., Okkonen, E.A., Reactivity of hydroxymeth-
ylated resorcinol coupling agent as it affects durability of epoxy bonds to
douglas-fir. Wood Fiber Sci., 30, 312–322, 1998.
28. Christiansen, A.W., Vick, C.B., Okkonen, E.A., Development of a novolak-
based hydroxymethylated resorcinol coupling agent for wood adhesives.
Forest Prod. J., 53, 2, 32–38, 2003.
29. Gardner, D.J., Frazier, C.E., Christiansen, A.W., Characteristics of the
wood adhesion bonding mechanism using Hydroxymethyl Resorcinol, in:
Wood Adhesives 2005, Frihart, C. (Ed.), pp. 93–97, 2005, Session 1B – Bond
Durability.
30. Son, J., Tze, W.T.Y., Gardner, D.J., Thermal behavior of hydroxymethylated
resorcinol (HMR)-treated wood. Wood Fiber Sci., 37, 220–231, 2005.
31. Gardner, D.J., Tze, W.T.Y., Shi, S.Q., Adhesive wettability of hydroxymethyl
resorcinol (HMR) treated wood, in: Proceedings Wood Adhesives 2000, pp.
321–327, Forest Products Society, Madison, WI, 2000.
RF Resins and Hydroxymethyl Resorcinol 167

32. Tze, W.T.Y., Bernhardt, G., Gardner, D.J., Christian, A.W., X-ray photo-
electron spectroscopy of wood treated with hydroxymethylated resorcinol
(HMR). Int. J. Adhes. Adhes., 26, 550–554, 2005.
33. Pizzi, A., Phenolic resins by reactions of coordinated metal ligands. J. Polym.
Sci., Polym. Lett., 17, 489, 1979.
34. Pizzi, A., Phenol and tannin-based adhesive resins by reactions of coordi-
nated metal ligands, Part 1: Phenolic chelates. J. Appl. Polym. Sci., 24, 1247–
1255, 1979.
35. Pizzi, A., Phenol and tannin-based adhesive resins by reactions of coordi-
nated metal ligands, Part II: Tannin adhesives preparation, characteristics
and application. J. Appl. Polym. Sci., 24, 1257–1268, 1979.
36. Vick, C.B. and Christiansen, A.W., Cure of phenol-fornaldehyde adhesive
in the presence of CCA-treated wood by differential scanning calorimetry.
Wood Fiber Sci., 25, 77–86, 1993.
37. AITC, American national standard for wood products- structural glued
laminated timber. ANSI/AITC A190.1-1992, American Institute of Timber
Construction, Vancouver, WA, 1992.
38. Lorenz, L.F. and Frihart, C.R., Adhesive bonding of wood treated with ACQ
and copper azole preservatives. Forest Prod. J., 56, 9, 90–93, 2006.
39. Dagher, H.J., Kimball, T.E., Shaler, S.M., Beckry, A.M., Effect of FRP rein-
forcement on low grade eastern hemlock glulams, pp. 207–214, National con-
ference on Wood Transportation Structures, Madison, WI, 1996.
40. Wipf, T.J., Ritter, M.A., Wood, D.L., Evaluation and testing of timber high-
way bridges, in: Pacific Timber Engineering Conference, G.B. Walford and D.J.
Gaunt (Eds.), pp. 333–340, Rotorua, New Zealand, 1999.
41. ASTM D 1101 97a, Standard Test Methods for Integrity of Adhesive Joints in
Structural Laminated Wood Products for Exterior Use, American Society for
Testing and Materials, West Conshohocken, Pennsylvania, USA, 1997.
42. Christansen, A.W. and Vick, C.B., Hydroxymethylated resorcinol coupling
agent for wood surfaces to produce exterior durable bonds, in: Silanes and
Other Coupling Agents, vol. 2, Mittal, K.L. (Ed.), pp. 193–208, 2000.
43. Eisenheld, L., Measuring the adhesive bond quality of vinyl ester-glass com-
posites on novolak HMR treated wood, MSc Dissertation, BOKU, University
of Agricultural Sciences, Vienna, Austria, 1997.
44. Eisenheld, L., Measuring the adhesive bond quality of vinyl ester–glass compos-
ites on novolak HMR treated wood, PhD Dissertation, University of Maine,
MN, Orono, Maine, 2003.
45. Christiansen, A.W., How overdrying wood reduces its bonding to PF adhe-
sives: A critical review of the literature Part 1 Physical responses. Wood Fiber
Sci., 22, 441–459, 1990.
46. Sernek, M., Comparative analysis of inactivated wood surfaces, PhD
Dissertation, Virginia Polytechnic Institute and State University, Blacksburg,
VI, 2002.
168 Adhesives for Wood and Lignocellulosic Materials

47. Kurt, R., Krause, A., Militz, H., Mai, C., Hydroxymethylated resorcinol
(HMR) priming agent for improved bondability of wax-treated wood. Holz.
Roh. Werkst., 66, 333–338, 2008.
48. Baxter, G.F. and Kreibich, R.E., Fast-curing phenolic adhesive system. Forest
Prod. J., 23, 1, 17–22, 1973.
49. Kreibich, R.E., High speed adhesives for the wood gluing industry. Adhes.
Age, 17, 1, 26–30, 1974.
50. Pizzi, A., Phenolic resins wood adhesives, in: Wood Adhesives Chemistry and
Technology, vol. 1, A. Pizzi (Ed.), pp. 105–178, Marcel Dekker, New York,
1983.
51. Pizzi, A., duT.Rossouw, D., Knuffel, W., Singmin, M., “Honeymoon” pheno-
lic and tannin-based fast setting adhesive systems for exterior grade finger-
joints. Holzforsch. Holzverwert., 32, 140–151, 1980.
52. Pizzi, A. and Cameron, F.A., Fast-set adhesives for glulam. Forest Prod. J., 34,
9, 61–65, 1984.
53. Pizzi, A. and Cameron, F.A., Fast setting adhesives for fingerjoints and glulam,
in: Wood Adhesives Chemistry and Technology, vol. 2, A. Pizzi (Ed.), pp. 229–
306, Marcel Dekker, New York, 1989.
54. Scopelitis, E., Synthetis, characteristics and applied aspects of cold-
setting urea-formaldehyde polymers, M.Sc. Dissertation, University of the
Witwatersrand, Johannesburg, South Africa, 1992.
55. Pizzi, A., Horak, R.M., Ferreira, D., Roux, D.G., Condensates of phenol,
resorcinol, phloroglucinol and pyrogallol, as flavonoids A- and B-rings
model compounds with formaldehyde, Part 1. J. Appl. Polym. Sci., 24, 1571–
1578, 1979.
56. Pizzi, A., Low resorcinol PRF cold set adhesives: The branching principle,
in: Wood Adhesives Chemistry and Technology, vol. 2, A. Pizzi (Ed.), pp. 190–
210, Marcel Dekker, New York, 1989.
57. Pizzi, A., Cameron, F.A., Orovan, E., Cold-set tannin-resorcinol-formaldehyde
adhesives of lower resorcinol content. Holz. Roh. Werkst., 46, 67–71, 1988.
58. Pizzi, A., Vosloo, R., Cameron, F.A., Orovan, E., Self-neutralizing acid-set
PF wood adhesives. Holz. Roh. Werkst., 44, 229–234, 1986.

Вам также может понравиться