Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

Tailoring of oxidized starch’s adhesion using crosslinker and adhesion

promotor for the recycling of fiberboards


Muhammad Adly Rahandi Lubis, Byung-Dae Park , Min-Kug Hong
Department of Wood and Paper Sciences, Kyungpook National University, Daegu 41566, Republic of Korea
Correspondence to: B.-D. Park (E-mail: byungdae@knu.ac.kr)

ABSTRACT: The growing interest in recycling waste medium density fiberboards (MDFs) is driving the development of new adhesives that
provide sufficient adhesion, and allow disintegration of the waste MDFs. Described in here is the preparation of adhesives based on oxi-
dized starch (OS) in combination with blocked-polymeric 4-4 diphenylmethane diisocyanate (B-pMDI) as a crosslinker and poly(vinyl
alcohol) (PVA) as an adhesion promotor for the recycling of waste MDFs. The COOH groups of OS were reacted with NCO groups
of B-pMDI to form amide linkages, and the CHO groups were reacted with OH groups of PVA through hydrogen bonding. Further,
when applied as an adhesive, the OS formed ester linkages with OH of MDF fibers. As the results, MDF bonded with 1% B-pMDI/15%
PVA/OS adhesive had an internal bonding strength of 0.13 MPa, 0.01 mg L−1 of formaldehyde emission (FE), and 12% of degree of fiber
disintegration. These results demonstrate that the B-pMDI/PVA/OS adhesive is a possible alternative to the current urea–formaldehyde
resins for the recycling of waste MDFs into recycled MDFs without FE. © 2019 Wiley Periodicals, Inc. J. Appl. Polym. Sci. 2019, 136, 47966.

KEYWORDS: adhesives; blends; cellulose and other wood products; manufacturing; recycling

Received 18 November 2018; accepted 28 April 2019


DOI: 10.1002/app.47966

INTRODUCTION expensive than these other renewable adhesives. Starch has been
As one of the fastest growing wood products, medium density used for bonding wood since the time of the ancient Egyptians.11
fiberboards (MDFs) are manufactured by bonding wood fibers Previous studies have reported that MDFs were fabricated by a
with urea–formaldehyde (UF) resins and are commonly used for partial substitution of UF resins with starch,12 a mixture of natu-
furniture and flooring applications.1 However, UF resins adhesive ral rubber latex and starch,13 and only native starch.14
emits formaldehyde,2 which is known to cause significant health Starch is a thermoplastic polymer with a linear structure of amy-
risks3,4 In addition, the poor recyclability of waste MDFs is lose and branched structure of amylopectin, which has three
another important challenge being investigated in the field. It was hydroxyl groups at C-2, C-3, and C-6 positions in each glucose
reported that the cured UF resins in waste MDFs was difficult to unit with the ability to form hydrogen bonds.15,16 The affinity of
remove5; hence, the recycling of waste MDFs is challenging. As a starch to water leads to poor water resistance and slow setting of
result, waste MDFs are usually sent to boilers or landfill in most starch-based adhesives.17 Hence, it is necessary to modify the starch
MDFs mills. Although the industry has established flue-gas in order to improve its performance as an adhesive. Several modifi-
cleaning systems, burning is not an environmentally friendly cations, such as esterification,18 crosslinking,19 and grafting20 have
solution as it releases dioxins, carbon dioxide, methane, and other been reported to improve the water resistance of starch-based adhe-
pollutants which can harm the environment.6,7 Therefore, the sives. However, the viscosity of starch-based adhesives modified
recycling of waste MDFs is more benign than burning or land- using those methods reaches 10,000 MPa s−1, which limits their
filling does.8 Growing interest in recycling waste MDFs is driving use for MDF production. One way to reduce the viscosity to an
the development of new adhesives that provide good adhesion, applicable range is to convert starch macromolecules into small
and allowing easy disintegration of MDFs. Consequently, it is molecules by oxidation processes. The oxidation of starch causes
necessary to develop alternative adhesives to replace UF resins, depolymerization and introduces aldehyde and carboxyl groups,
which are suitable for producing high-quality MDF products with which minimize the retrogradation of amylose and eventually
favorable mechanical properties, and more efficient recycling. improves the stability of the oxidized starch (OS).21 During the oxi-
Although MDFs have been prepared using renewable adhesives, dation, hydroxyl groups at the C-2, C-3, and C-6 positions of the
such as lignin and chitosan,9,10 starch has good potential as an starch are first oxidized to aldehyde groups, and further oxidation
adhesive for manufacturing MDFs because it is generally less converts free hydroxyl or aldehyde groups into carboxyl groups.22

© 2019 Wiley Periodicals, Inc.

47966 (1 of 10) J. APPL. POLYM. SCI. 2019, DOI: 10.1002/APP.47966


ARTICLE WILEYONLINELIBRARY.COM/APP

OS has been used in film formation,23 coatings,24 binders,25 ther- The temperature was adjusted to 50  C, and subsequently, 0.5%
moplastic composites,26 flame retardants,27 and wood adhesives28; copper sulfate, 0.5% ammonium persulfate, and 0.25% sodium
hence, it is expected to be a potential adhesive for MDF. dedocyl sulfate based on the weight of starch were added to the
gelatinized starch. Subsequently, 127.5 mL H2O2 was added to
The hydrophilicity of starch should make the OS adhesive easy to
obtain OS with a H2O2/starch mole ratio of 1.0. The oxidation reac-
recycle. In addition, no formaldehyde is emitted from the OS-
tion was performed by maintaining the temperature at 50  C and
bonded panel. However, to date, very limited studies have investi-
pH at 3 for 5 h under continuous stirring at 500 rpm, and the reac-
gated the use of OS on the manufacture of MDFs for their
tion was completed when the pH was stable at 3.
recycling. Therefore, it is necessary to develop an OS adhesive
that provides an appropriate balance between good adhesion and
Preparation of OS Adhesives
easy disintegration of MDFs for recycling. Moreover, the OS
Two series of OS adhesives were prepared for MDF manufacture
adhesive should have low viscosity to allow it to be sprayed onto
and recycling. The first one was prepared by mixing OS with
wood fibers for bonding.11 Available literature reports that the
B-pMDI levels of 0, 1, 3, and 5% based on the OS solids content.
modification of OS with a crosslinker can promote chemical
Approximately 0.7, 2.2, an© 2019 Wiley Periodicals, Inc.d 3.6 g of
crosslinking by forming ester linkages.29 Blocked-polymeric 4-4
B-pMDI were added into 200 g of OS to prepared 1, 3, and 5% B-
diphenylmethane diisocyanate (B-pMDI) can be used as a
pMDI/OS, respectively. The mixture was stirred at 80  C and
crosslinker which releases NCO groups after deblocking, and pro-
300 rpm for 30 min. The B-pMDI with a HSO3/NCO mole ratio of
motes reactions with aldehyde and carboxyl groups of OS.30 In
1.5 had a deblocking temperature of 78.3  C was prepared
addition, poly(vinyl alcohol) (PVA) can be used as a promotor to
according to the published method.31 The second series of adhe-
improve the adhesion of OS through the formation of hydrogen
sives was prepared by adding different PVA contents (5, 10, and
bonds between OS and PVA.19
15% based on the OS solids content) into the OS. Around 3.6, 7.2,
Hence, the present work aims to prepare a novel MDF adhesive and 10.8 g of PVA were added into 200 g of OS to prepared 5, 10,
based on OS synthesized via oxidation using hydrogen peroxide. and 15% PVA/OS, respectively. The mixture was stirred at 80  C
OS’s adhesion performance was tuned by mixing OS with B- and 500 rpm for 2 h. Furthermore, around 0.7 g of B-pMDI (1%
pMDI as a crosslinker and PVA as an adhesion promoting agent. based on the OS solids content) was added to the PVA/OS and
The functional groups and thermal properties of neat OS, B- stirred for 30 min. This optimum level was obtained by comparing
pMDI/OS, and B-pMDI/PVA/OS adhesives were investigated the mean values of density, internal bonding (IB) strength, modulus
using Fourier transform infrared (FTIR) spectroscopy and differ- of rupture (MOR), modulus of elasticity (MOE), thickness swelling
ential scanning calorimetry (DSC) analysis. This article, for the (TS), water absorption (WA), and formaldehyde emission (FE); the
first time, reports a method of tailoring the adhesion of OS adhe- B-pMDI/OS-bonded MDF using statistical analysis. As a control,
sives by combining with crosslinker and adhesion promoter for UF resins with formaldehyde/urea mole ratio of 1.0 were synthe-
the recycling of MDFs. It is expected that the control of levels of sized using an alkaline–acid two-step reaction, as reported in the
B-pMDI and PVA in the OS adhesives could provide a good bal- literature.32,33
ance between adhesion and easy disintegration with MDFs for
their recycling. MDF Manufacture
Laboratory MDFs were manufactured using three types of adhe-
EXPERIMENTAL sives: UF resins, B-pMDI/OS, and B-pMDI/PVA/OS. As a control,
Materials two MDFs with 700 kg m−3 density were manufactured with UF
Corn starch (28.5% amylose, 18% moisture content), hydrogen per- resin according to a published method.34 Briefly, a mixture of 12%
oxide (30 wt %), copper sulfate (99 wt %), sodium dedocyl sulfate UF resins and 1% emulsion wax (based on the dried fiber mass)
(98 wt %), ammonium persulfate (99 wt %), and PVA (molecular was mixed with 3% ammonium chloride (based on the UF resins
weight of 22,000 g mol−1) were used to prepare the OS. Liquid pMDI solids content). The adhesives were sprayed onto the fibers in a
(31% NCO, Lupranate M20S, BASF, Ludwigshafen, Germany), drum-type mechanical blender using an atomization nozzle with
sodium bisulfate (58.5 wt %), and acetone (99.8 wt %) were used to 1 mm diameter. The blended fibers were then used to prepare the
prepare the B-pMDI. Urea granules (99 wt %), formalin (37 wt %), fiber mat using a centrifugal air blower. The fiber mat was pre-
formic acid (20 wt %), sodium hydroxide (20 wt %), and ammonium pressed for 30 s and then hot-pressed at 180  C using a single-phase
chloride (20 wt %) were used to synthesize UF resin. These chemicals pressing schedule of 2.45 MPa for 4 min to produce MDFs
were purchased from Duksan or Daejung Chemical, Seoul, Korea. (300 × 400 × 15 mm3). For the evaluation of the OS adhesives,
Virgin wood fibers made of Red pine (Pinus densiflora Siebold & eight MDFs with 700 kg m−3 density were prepared using
Zucc., 3% moisture content) were supplied from a commercial MDF B-pMDI/OS adhesives with various B-pMDI contents. Approxi-
mill (Hansol Home Deco, Iksan, Korea). mately 12% of B-pMDI/OS based on the fiber dry mass was sprayed
onto the fibers using the same conditions as UF resins-bonded
Oxidation of Starch MDF. The fibers sprayed with OS adhesives were then used to pre-
OS was prepared according to previously published methods with pare the fiber mat using a centrifugal air blower. The fiber mats
some modifications.28,29 A starch solution was prepared by mixing were prepressed for 30 s and then hot-pressed at 180  C using a
200 g corn starch and 250 mL of distilled water. The solution was single-phase pressing schedule of 2.45 MPa for 10 min to fabricate
sonicated for 30 min at 25  C and 20 kHz. The solution then was MDFs. Using the same method as for the B-pMDI/OS samples, six
gelatinized at 80  C for 1 h under continuous stirring at 100 rpm. MDFs were prepared using B-pMDI/PVA/OS adhesives with

47966 (2 of 10) J. APPL. POLYM. SCI. 2019, DOI: 10.1002/APP.47966


ARTICLE WILEYONLINELIBRARY.COM/APP

different PVA contents. All MDFs were conditioned for 2 days at


28  2  C prior to cutting and testing.

Characterization of OS Adhesives
Some basic properties of OS adhesives, such as nonvolatile solids
content, viscosity, and gelation time were determined. The nonvola-
tile solids content of OS adhesives for various B-pMDI and PVA
contents was obtained by drying 2 g of the sample in an oven at
105  C for 3 h and dividing the oven-dried weight by the initial
weight. The viscosity of OS adhesives was determined using a cone-
plate viscometer (DV-II+, Brookfield, Middleboro, MA) with
spindle no. 2 at 27  2  C and 60 rpm. The gelation time of OS
adhesives was measured using a gel time meter at 100  C (Davis
Inotek Instruments). All experiments were replicated three times.
FTIR spectroscopy (Frontier; PerkinElmer) was performed to quali-
tatively investigate the functional groups of the OS adhesives with
different B-pMDI and PVA contents. The powdered adhesives were Figure 1. Typical FTIR spectra of native starch and OS. [Color figure can
pressed into pellets with KBr as a reference, and were scanned from be viewed at wileyonlinelibrary.com]
400 to 4000 cm−1 with a resolution of 4 cm−1 and 32 scans per
sample. Semiquantitative analysis was done by normalizing the each (25  C for 6 h and 80  C for 1 h) as the previously reported proce-
spectrum using the min–max normalization method. It was done dures with three replications for each sample.5,37 The mass loss, the
by subtracting each spectrum with its minimum value and divided degree of fiber disintegration, pH of the extract solution, and func-
with its range value (maximum value subtracts minimum value).35 tional groups of the MDFs after hydrolysis were determined to
The change in normalized intensity of functional groups was compare the recyclability of the MDF. The mass loss was deter-
observed as a function of B-pMDI and PVA. A DSC system mined by measuring the weight of MDF cubes before and after
(Discovery 25; TA Instruments) was used to investigate the exother- hydrolysis, while the pH was determined using a digital pH meter.5
mic peak temperature of B-pMDI/PVA/OS adhesives at different Furthermore, the degree of fiber disintegration was calculated by
levels of B-pMDI and PVA. Approximately 5 mg of liquid dividing the dry weight of MDF cubes, in the panel form, after
B-pMDI/PVA/OS mixed was sealed in a high-pressure capsule pan hydrolysis with the dry weight of MDF cubes before hydrolysis.
and then heated from 30 to 300  C at 2.5, 5, and 10  C min−1 of
heating rate, and under a nitrogen flow of 50 mL min−1. Statistical Analysis
Physical and Mechanical Properties of MDF Multivariate analysis was conducted to compare the mean values
Some physical and mechanical properties of MDFs were deter- of MDF properties, while Duncan multiple range test at α = 0.05
mined using a standard procedure.36 Around 10 samples was used to determine the optimum B-pMDI and PVA content
(50 × 50 × 15 mm3) were prepared to determine IB strength, and in the OS adhesive. Statistical analysis was undertaken using the
four samples (200 × 50 × 15 mm3) were prepared for MOR and SPPS 17 software (SPSS Inc.).
MOE. IB strength, MOE, and MOR were determined using a uni-
versal testing machine (H50KS; Tinius Olsen Ltd., Redhill, UK) at RESULTS AND DISCUSSION
loading speeds of 2, 10, and 10 mm min−1, respectively. Further, Adhesive Properties
10 samples for TS and WA tests were soaked for 24 h in water, FTIR spectra showed the formation of C O bonds (1715 cm−1)
and then the thickness and weight were measured after removing in the aldehyde and carboxyl groups in the OS (Figure 1), indi-
surface water. FE was determined by placing 10 MDF specimens cating that starch was successfully oxidized by converting the
in a desiccator at 20  2  C for 24 h. Further, approximately OH groups on the C-2, C-3, and C-6 positions of the glucose
25-mL aliquot solutions were pipetted into 100-mL flasks, and unit to aldehyde and carboxyl groups using hydrogen peroxide.22
25 mL acetyl acetone–acetic acid ammonium solutions were sub- In addition, the peak at 1630 cm−1 in both native starch and OS
sequently added to the flask. The solutions were then heated for was attributed to intramolecular O H bonds.23
10 min at 65  2  C in a water bath. Finally, the solutions were
cooled to room temperature (25  2  C) before further analysis. After the addition of B-pMDI into OS, there were several new peaks
FE was analyzed using ultraviolet–visible spectrophotometry detected, such as C H of pMDI at 2723 cm−1, C O of amide link-
(Optizen 3220 UV; Mecasys, Korea) at a wavelength of 412 nm to age at 1586 cm−1, and C N of amide at 1348 cm−1 [Figure 2(a)].
determine the FE values. In addition, the peak at 3300 cm−1, which belonged to O H in OS,
was replaced with the N H of pMDI at 3258 cm−1. No significant
Hydrolysis of MDF changes were observed for other functional groups of B-pMDI/OS
MDF panels were cut into cube specimens (20 × 20 × 15 mm3). A as the B-pMDI content increased. Furthermore, FTIR spectroscopy
total of 30 MDF cubes were prepared from five different MDF sam- was also conducted to evaluate whether additional chemical reac-
ples: UF resin, 1% B-pMDI/OS, and 1% B-pMDI/PVA/OS at three tions occurred in 1% B-pMDI/PVA/OS with different PVA con-
different PVA contents (5, 10, and 15% based on the OS solids con- tents [Figure 2(b)]. The peak at 3300 cm−1 was assigned to O H of
tent). The MDF cubes were hydrolyzed under two conditions OS and intermolecular hydrogen bonds in PVA/OS. This peak did

47966 (3 of 10) J. APPL. POLYM. SCI. 2019, DOI: 10.1002/APP.47966


ARTICLE WILEYONLINELIBRARY.COM/APP

Figure 2. FTIR spectra of (a) B-pMDI/OS with different B-pMDI contents, (b) 1% B-pMDI/PVA/OS with different PVA contents, (c) normalized intensity
of B-pMDI/OS with different B-pMDI contents, (d) normalized intensity of 1% B-pMDI/PVA/OS with different PVA contents (e) normalized intensity of
specific functional groups in B-pMDI/OS, and (f) normalized intensity of specific functional groups in 1% B-pMDI/PVA/OS. [Color figure can be viewed at
wileyonlinelibrary.com]

not change to that for N H (3258 cm−1) due to the higher portion normalized intensity, and new peak was detected at 1600 cm−1,
of PVA than B-pMDI. The intensity of this peak increased with which was attributed to the C O stretching vibration of acid
increasing PVA content in the OS. Other peaks at 2925, 1715, 1630, hydrolyzed PVA [Figure 2(e,f)].39
and 1348 cm−1 were similar to those observed in the B-pMDI/OS
DSC analysis was performed to evaluated thermal behavior of OS
spectra.
adhesives. As expected, the Tp values of OS adhesives increased as a
Semiquantitative analysis showed that the normalized intensity of function of heating rate [Figure 3(a)], and the Tp values decreased
C O of carboxyl groups at 1715 cm−1 decreased from 0.21 to an as the PVA content increased [Figure 3(b)]. Further, DSC thermo-
average 0.13 after the addition of B-pMDI, meaning that the car- grams depicted two exothermic peaks (Tp) in B-pMDI/OS and B-
boxyl groups reacted with the NCO group of B-pMDI to form pMDI/PVA/OS adhesives. The first peak, Tp1, (dot) at 134–155  C
amide linkages, as indicated by the peak at 1586 and 1348 cm−1 only appeared when B-pMDI was added, while no peak was
[Figure 2(c,d)].38 In case of 1% B-pMDI/PVA/OS adhesives, it observed in pure OS, meaning that the Tp1 was attributed to the for-
was also found that peak of carboxyl groups at 1715 decreased in mation of amide linkages from the reaction of carboxyl groups of

47966 (4 of 10) J. APPL. POLYM. SCI. 2019, DOI: 10.1002/APP.47966


ARTICLE WILEYONLINELIBRARY.COM/APP

Figure 3. Typical DSC thermogram of (a) 1% B-pMDI/OS at different heating rates, and (b) 1% B-pMDI/PVA/OS with different levels of PVA at
2.5  C min−1. [Color figure can be viewed at wileyonlinelibrary.com]

OS and NCO groups of B-pMDI. It was possible that the NCO of reduced by 12% with the addition of 1% B-pMDI and 15% PVA,
B-pMDI could have reacted with COOH of OS to form amide indicating that B-pMDI/PVA/OS had higher reactivity than the
(R NCO R) linkages.38 Second peak, Tp2, (asterisk) around pure OS. This was probably due to the chemical crosslinking in the
210  C could be due to the degradation of OS adhesives, which also OS adhesives, which was supported by the FTIR spectra (Figure 2)
happened at exothermic.40 This peak decreased with increasing in and DSC of OS adhesives (Figure 3). FTIR spectra and DSC ther-
PVA level in B-pMDI/PVA/OS adhesive. mograms showed the formation of amide linkage (1586 and
The nonvolatile solids content, viscosity, and gelation time of the 1348 cm−1) for crosslinking the OS adhesives at 134  C when 1%
OS adhesives are summarized in Table I. Pure OS had quite low B-pMDI and 15% PVA were added into the OS. However, no peak
solids content, low viscosity, and long gelation time, which was in was observed in pure OS. In addition, an increase in viscosity after
agreement with the previous study.29,30 No significant changes addition of 15% PVA probably resulted in more close-packed OS
were observed after the addition of B-pMDI. However, the solids molecules, which could reduce the gelation time. However, mate-
content and viscosity of OS increased gradually with the addition rials with such a long gelation time would require longer hot-
of B-pMDI and PVA together. The gelation time of the OS was pressing time to manufacture the MDF panels.

Table I. Properties of the Adhesives

1% B-pMDI/5% 1% B-pMDI/10% 1% B-pMDI/15%


Property UF resin OS B-pMDI PVA/OS PVA/OS PVA/OS
Solids content (%) 62.1 36.3 100 37.7 42.7 48.3
−1 a
Viscosity (MPa s ) 272 124.0 151.6 171.3 191.7
a
Gelation time (s) 187 625 597 563 549
a
The values are not available in the solid form.

Table II. Properties of MDFs Bonded with B-pMDI/OS Adhesives at Different B-pMDI Contents

B-pMDI content (wt %) Density (kg m−3) IB strength (MPa) MOR (MPa) MOE (MPa) TS (%) WA (%) FE (mg L−1)
Controla 728 0.13 7.8 2661.4 54.1 78.1 0.89
(8.93)b (0.01) (0.50) (169.0) (5.07) (10.55) (0.02)
0 731 0.06 4.5 1530.1 150.8 431.9 0.01
(18.47) (0.01) (0.53) (177.0) (19.35) (15.44) (0.00)
1 736 0.08 5.1 1835.4 60.3 123.9 0.01
(16.03) (0.01) (0.62) (208.0) (10.39) (17.90) (0.00)
3 735 0.07 4.8 1493.5 85.3 178.3 0.01
(18.56) (0.01) (0.95) (317.0) (14.77) (19.58) (0.00)
5 730 0.03 3.9 1292.6 100.6 261.6 0.01
(22.59) (0.01) (0.71) (239.0) (10.54) (29.11) (0.00)
a
The control is MDFs bonded with UF resins.
b
The values in parentheses are the standard deviation.

47966 (5 of 10) J. APPL. POLYM. SCI. 2019, DOI: 10.1002/APP.47966


ARTICLE WILEYONLINELIBRARY.COM/APP

Table III. Duncan Multiple Range Test Results for the Properties of MDFs panel density and FE values. IB strength, MOR, and MOE values
Bonded with OS Adhesive with Different B-pMDI Contents of MDFs bonded with B-pMDI/OS showed similar trends, show-
ing that the values increased up to 1% B-pMDI addition and
B-pMDI then decreased after 3 and 5% B-pMDI addition. It was expected
content that higher B-pMDI content would improve the performance of
(wt %) Density IB strength MOR MOE TS WA FE MDF; however, it was not the case. This may have been due to
0 a* b a b c d a the B-pMDI agglomerating at high concentration in the OS,
resulting in dark spots on the MDF surface.41
1 a a b a a a a
3 a ab bc ab a b a The B-pMDI/OS-bonded MDFs had an average of 0.01 mg L−1
5 a c c ab b c a
of FE while the UF-bonded MDFs emitted nearly 90 times of this
level (0.89 mg L−1). It was clear that the FE of OS adhesives was
*The different letters are significantly different at p value of 0.05. in the range of a detection limit. However, the adhesion perfor-
mance of B-pMDI/OS was not comparable to that of the UF
Properties of MDF resins. This is probably because the OS adhesives had lower solids
The properties of MDFs bonded with OS at different B-pMDI content, lower viscosity, and longer gelation time than those of
contents are shown in Table II. The B-pMDI/OS bonded MDF the UF resins (Table I), resulted in lower reactivity and lower
panel generally had poorer physical and mechanical properties crosslinking density than that of UF resins.42 The TS and WA
than those of MDFs bonded with neat UF resins, except the values of MDFs bonded with OS adhesives were six times higher

Figure 4. Properties of MDFs bonded with 1% B-pMDI/PVA/OS with different PVA contents: (a) IB strength; (b) FE; (c) MOR; (d) MOE; (e) TS; and (f) WA.

47966 (6 of 10) J. APPL. POLYM. SCI. 2019, DOI: 10.1002/APP.47966


ARTICLE WILEYONLINELIBRARY.COM/APP

preparation of OS adhesives, the FE values were close to a detec-


tion limit. However, the small amount of formaldehyde from
MDFs bonded with OS adhesives was originated from the alde-
hyde groups of OS, or from biogenic formaldehyde.43
Both MOR and MOE values showed a similar trends to the IB
strength, which increased with increasing PVA content. The sam-
ple with 15% PVA in B-pMDI/OS showed comparable values to
the UF resins-bonded MDFs [Figure 4(c,d)]. MDFs bonded with
1% B-pMDI/15% PVA/OS had smooth, and dense surfaces due
to the longer hot-pressing time, resulting in similar MOR and
MOE values to those of the UF resins-bonded MDFs. The TS
and WA values of MDFs bonded with 1% B-pMDI/PVA/OS were
comparable to those of the UF-bonded MDFs [Figure 4(e,f)]. The
presence of ester linkages at 1730 cm−1 and hydrogen bonds in
MDFs bonded with 1% B-pMDI/PVA/OS (Figure 5) could have
Figure 5. Typical FTIR spectra of MDFs bonded with 1% B-pMDI/PVA/OS enhanced both adhesion and water resistance of the MDFs. The
adhesives with different PVA contents. [Color figure can be viewed at statistical analysis showed that MDFs with good mechanical
wileyonlinelibrary.com] properties, proper dimensional stability, and almost zero formal-
than those of UF-bonded MDFs. However, introducing B-pMDI dehyde emission can be produced by adding 15% PVA to the 1%
into the OS could improve the water resistance of MDFs, where B-pMDI/OS adhesives (Table IV).
1% B-pMDI/OS showed 60.3% TS and 123.9% WA values, while
the UF-bonded MDFs showed 54.1% TS and 78.1% WA. The sta-
tistical analysis suggested that the addition of 1% B-pMDI in the Recyclability of MDFs
OS was an optimum level to obtain the best MDF properties Developing OS adhesives with good adhesion performance and
(Table III). high recyclability is necessary for developing recyclable MDFs.
To obtain OS adhesives with acceptable physical and mechanical Herein, the recyclability of MDFs bonded with 1% B-
properties of MDF, as well as good recyclability, PVA was added pMDI/PVA/OS with various PVA contents was investigated at dif-
into the OS prior to the addition of 1% B-pMDI. Physical proper- ferent hydrolysis conditions. Figure 6 shows pictures of MDFs
ties, mechanical properties, and FE values of MDFs bonded with bonded with UF resin and 1% B-pMDI/15% PVA/OS after hydro-
1% B-pMDI/PVA/OS as a function of PVA content are shown in lysis at 25  C for 6 h. It can be seen that the UF-bonded MDF
Figure 4. The IB strength of MDFs bonded with 1% solution did not change color after hydrolysis for 6 h [Figure 6(a)].
B-pMDI/PVA/OS increased gradually with increasing the PVA In contrast, the solution of MDFs bonded with 1% B-pMDI/15%
content. The addition of 15% PVA in the OS resulted in an IB PVA/OS started to change from a clear solution to light brown
strength of 0.13 MPa, which was comparable to that of UF resins- after 2 h of hydrolysis, and became dark brown after 6 h [Figure 6
bonded MDFs [Figure 4(a)]. This could be due to the presence of (b)]. The brown color is thought to be due to the dissolution of
ester linkages in MDFs bonded with 1% B-pMDI/PVA/OS. The OS adhesives and wood extractives in the solvent. In the case of
carboxyl groups of OS adhesives can react with OH groups of MDFs bonded with 1% B-pMDI/15% PVA/OS, the OS adhesives
the wood, forming C O groups as ester linkage, as shown by the were much easier to be removed than the UF resins did. As shown
FTIR peak at 1730 cm−1 in Figure 5. The addition of PVA greatly in Figure 7, the solution of both 1% B-pMDI/15% PVA/OS- and
increased the solids content and viscosity, and decreased the gela- UF resins-bonded MDFs hydrolysis at 80  C for 1 h changed to
tion time of the B-pMDI/PVA/OS adhesives via hydrogen bonding light brown after 10 and 20 min. A similar observation was
between OS and PVA (Table I). Similar to B-pMDI/OS-bonded reported for the removal of cured UF resin from MDFs.5,37
MDFs, the FE values of MDFs bonded with 1% B-pMDI/PVA/OS The mass loss of MDFs after hydrolysis is shown in Figure 8(a).
were 0.01 mg L−1, while UF-bonded MDFs showed a value of Regardless of the type of adhesive, hydrolysis at 80  C for 1 h
0.89 mg L−1 [Figure 4(b)]. Since formaldehyde was not used in the resulted in higher mass loss than those of the hydrolysis at 25  C
for 6 h. This mass loss was responsible for the removal of OS
Table IV. Statistical Analysis of the Properties of MDFs Bonded with 1% adhesives and wood extractives, and was reduced by adding 15%
B-pMDI/OS Adhesives at Different PVA Contents PVA, but still maintained at higher value than that of UF resins.
This indicated that 1% B-pMDI/15% PVA/OS could be used to
Concentration manufacture MDFs with a proper adhesion performance and high
of PVA (wt %) IB strength MOR MOE TS WA FE recyclability compared to UF resins-bonded MDFs. The degree of
0 c* b b c d a fiber disintegration was also investigated to evaluate the recyclabil-
5 b ab ab b c a ity of MDFs [Figure 8(b)]. More fibers were disintegrated after
10 a ab ab ab b a hydrolysis at 80  C for 1 h compared to treatment at 25  C for
6 h. Similar to the mass loss, the addition of 15% PVA to the 1%
15 a a a a a a
B-pMDI/OS decreased the degree of fiber disintegration, while the
*The different letters are significantly different at p value of 0.05. value was still higher than that of UF resins.

47966 (7 of 10) J. APPL. POLYM. SCI. 2019, DOI: 10.1002/APP.47966


ARTICLE WILEYONLINELIBRARY.COM/APP

Figure 6. Photographs of the hydrolysis of MDFs bonded with (a) UF resins and (b) 1% B-MDI/15% PVA/OS at 25  C for 6 h. [Color figure can be viewed
at wileyonlinelibrary.com]

Figure 7. Photographs of the hydrolysis of MDFs bonded with (a) UF resins and (b) 1% B-pMDI/15% PVA/OS at 80  C for 1 h. [Color figure can be viewed
at wileyonlinelibrary.com]

Measurements of the pH of the extract solutions after the hydrolysis 3. Figure 9(a) shows the pH of MDF extract solutions after hydroly-
of MDF cubes were performed to verify whether the OS was sis at 25  C for 6 h. The pH of all samples decreased gradually as a
extracted by this process. The original pH of the OS adhesives was function of hydrolysis time. The pH of the extract solutions of UF

Figure 8. (a) Mass loss and (b) fiber disintegration of MDFs bonded with 1% B-pMDI/OS adhesives with different PVA contents after hydrolysis under dif-
ferent conditions.

47966 (8 of 10) J. APPL. POLYM. SCI. 2019, DOI: 10.1002/APP.47966


ARTICLE WILEYONLINELIBRARY.COM/APP

Figure 9. pH values of the extract solutions from MDFs bonded with 1% B-pMDI/OS adhesives with different PVA contents after hydrolysis at (a) 25  C
for 6 h and (b) 80  C for 1 h. [Color figure can be viewed at wileyonlinelibrary.com]

resins-bonded MDFs was in the range of 4.9–5.1, while those of ACKNOWLEDGMENTS


MDFs bonded with pure OS were in the range 3.8–4.4. The pH of This study was conducted with the support from the Korea Forest
virgin Red pine fiber was reported to be around 4.8–5.5.44–47 There- Science Technology (Project grant no. 2015077B10-1718-AA01)
fore, it was clear that most of the OS adhesive was extracted after provided by the Korea Forest Service (Korea Forestry Promotion
hydrolysis of OS-bonded MDFs at 25  C for 6 h, while only wood Institute).
extractives were extracted from UF resins-bonded MDFs. However,
it should be noted that the addition of B-pMDI and PVA to the OS
increased the pH of extract solutions from 3.8 to 4.8, indicating the REFERENCES
presence of ester linkages and hydrogen bonds (Figure 5) changed 1. Mantanis, G. I.; Athanassiadou, E. T.; Barbu, M. C.;
the pH of pure OS. By contrast, the hydrolysis of MDFs bonded Wijnendaele, K. Wood Mater. Sci. Eng. 2018, 13, 104.
with OS adhesives at 80  C for 1 h had the same pH values as in
the case of hydrolysis at 25  C for 6 h [Figure 9(b)]. Hence, high 2. Abdullah, Z. A.; Park, B. D. J. Appl. Polym. Sci. 2009, 114,
temperature hydrolysis extracted a similar amount of OS adhesive 1011.
compared to the low-temperature treatment, but in a shorter time. 3. Hauptmann, M.; Jay, H.; Lubin, P. A.; Stewart, R. B.;
In particular, the pH of the extract solution from UF resins-bonded Hayes; Aaron, B. J. Natl. Cancer Inst. 2003, 95, 1615.
MDFs increased with increasing hydrolysis time, which was clearly 4. Cogliano, V.; Yann, G.; Robert, B.; Kurt, S.; Béatrice, S.;
due to the formation of free ammonia from the degradation of UF Fatiha, E. G. Lancet Onco. 2004, 5, 528.
resins at high temperature.48,49
5. Lubis, M. A. R.; Hong, M. K.; Park, B. D. J. Wood Chem.
Technol. 2018, 38, 1.
CONCLUSIONS 6. Karazipour, A.; Kües, U. Wood Production, Wood Tech-
OS adhesives were successfully prepared by mixing OS with vari- nology, and Biotechnological Impacts; Universitätsverlag:
ous B-pMDI and PVA contents, which were then used to pro- Göttingen, Germany, 2007. p. 509.
duce MDFs. MDFs bonded with pure OS had low IB strength 7. Tatàno, F.; Barbadoro, L.; Mangani, G.; Pretelli, S.;
and very poor water resistance. Introducing B-pMDI into the OS Tombari, L.; Mangani, F. Waste Manag. 2009, 29, 2656.
greatly enhanced the physical and mechanical properties of 8. Morris, J. J. Ind. Ecol. 2017, 21, 844.
MDFs panel, but the properties were still inferior to those of UF
resins-bonded MDFs. However, the MDF properties could be 9. Tupciasukas, R.; Gravitis, J.; Abolins, J.; Veveris, A.;
improved by adding PVA to 1% B-pMDI/OS, which had physical Andzs, M.; Liitia, T.; Tamminen, T. Holzforschung. 2017,
and mechanical properties comparable to those of the UF resins 71, 555.
system. The mass loss and degree of fiber disintegration after 10. Ji, X.; Dong, Y.; Yu, R.; Du, W.; Gu, X.; Guo, M.
hydrolysis of MDFs bonded with OS adhesives were higher than Holzforschung. 2018, 72, 275.
those of UF resins-bonded MDFs, indicating enhanced recyclabil- 11. Onusseit, H. Ind. Crop Prod. 1993, 1, 141.
ity. In addition, the pH of the extract solutions after hydrolysis of
12. Jarusombuti, S.; Bauchongkol, P.; Hiziroglu, S.;
MDFs bonded with OS adhesives was more acidic than those of
Fueangvivat, V. Forest Prod. J. 2012, 62, 58.
UF resins-bonded MDFs, showing that the OS adhesives were
much easier to extract. We proposed that the OS adhesive with a 13. Akbari, S.; Gupta, A.; Khan, T. A.; Jamari, S. S.;
composition of 1% B-pMDI and 15% PVA was found as a proper Ani, N. B. C.; Poddar, P. J. Indian Acad. Wood Sci. 2014,
combination for manufacturing MDFs panel with balanced adhe- 11, 109.
sion and disintegration properties to produce high-quality MDF 14. Abbott, A. P.; Conde, J. P.; Davis, S. J.; Wise, W. R. Green
panels that can be easily recycled. Chem. 2012, 14, 3067.

47966 (9 of 10) J. APPL. POLYM. SCI. 2019, DOI: 10.1002/APP.47966


ARTICLE WILEYONLINELIBRARY.COM/APP

15. Buléon, A.; Colonna, P.; Planchot, V.; Ball, S. Int. J. Biol. 32. Park, B. D.; Jeong, H. W.; Lee, S. M. J. Appl. Polym. Sci.
Macromol. 1998, 23, 85. 2011, 120, 1475.
16. Tester, R. F.; Karkalas, J.; Qi, X. J. Cereal Sci. 2004, 39, 151. 33. Lubis, M. A. R.; Park, B. D. Holzforschung. 2018, 72, 759.
17. Hemmilä, V.; Adamopoulos, S.; Karlsson, C.; Kumar, A. 34. Hong, M. K.; Lubis, M. A. R.; Park, B. D. J. Korean Wood
RSC Adv. 2017, 7, 38604. Sci. Technol. 2017, 45, 444.
18. Din, Z. U.; Xiong, H.; Wang, Z.; Fei, P.; Ullah, I.; 35. Korean Standard F 3200: Fiberboards, Korean Standard
Javaid, A. B.; Wang, Y.; Jin, W.; Chen, L. Int. J. Biol. Association, Seoul, 2006.
Macromol. 2017, 104, 846.
36. Wilkox, K. R. Thesis, University of Manchester, 2014.
19. Sridach, W.; Jonjankiat, S.; Wittaya, T. J. Adhes. Sci.
Technol. 2013, 27, 1727. 37. Grigsby, W. J.; Carpenter, J. E. P.; Sargent, R. J. Wood
Chem. Technol. 2014, 34, 225.
20. Wang, Z.; Gu, Z.; Hong, Y.; Cheng, L.; Li, Z. Carbohydr.
Polym. 2011, 86, 72. 38. Blagbrough, I. S.; Mackenzie, N. E.; Ortiz, C.; Scott, A. I.
21. Rutenberg, M. W.; Solarek, D. In Starch: Chemistry and Tetrahedron Lett. 1986, 27, 1251.
Technology, 2nd Ed.; Whistler, R. L., Bemiller, J. N., 39. Mansur, H.; Sadahira, C. M.; Souza, A. N.;
Paschall, E. F., Eds.; Academic Press: Atlanta, GA, USA, Mansur, A. A. P. Mater. Sci. Eng. C. 2008, 28, 539.
1984, p. 311. 40. Soliman, A. A. A.; El-Shinnawy, N. A.; Mobarak, F.
22. Patel, K. F.; Mehta, H. U.; Srivastava, H. C. J. Appl. Polym. Thermochim. Acta. 1997, 296, 149.
Sci. 1974, 18, 389.
41. Cooper, M. D. Thesis, University of Canterbury, 1992.
23. Xu, H.; Canisag, H.; Mu, B.; Yang, Y. ACS Sustain. Chem.
42. Park, B. D.; Kim, J. W. J. Appl. Polym. Sci. 2008, 108, 2045.
Eng. 2015, 3, 2631.
24. Caźon, P.; Velazquez, G.; Ramírez, J. A.; Vázquez, M. Food 43. Tasooji, M.; Wan, G.; Lewis, G.; Wise, H.; Frazier, C. E.
Hydrocoll. 2017, 68, 136. ACS Sustain. Chem. Eng. 2017, 5, 4243.
25. Bie, Y.; Yang, J.; Nuli, Y.; Wang, J. RSC Adv. 2016, 6, 97084. 44. Xing, C.; Zhang, S. Y.; Deng, J. Holzforschung. 2004, 58, 408.
26. Wu, H.; Liu, C.; Chen, J.; Chen, Y.; Anderson, D. P.; 45. Xing, C.; Zhang, S. Y.; Deng, J.; Riedl, B.; Cloutier, A. Wood
Chang, P. R. J. Appl. Polym. Sci. 2010, 118, 3082. Sci. Technol. 2006, 40, 637.
27. Zhang, S.; Liu, F.; Peng, H.; Peng, X.; Jiang, S.; Wang, J. 46. Lubis, M. A. R.; Hong, M. K.; Park, B. D.; Lee, S. M. Eur.
Ind. Eng. Chem. Res. 2015, 54, 11944. J. Wood Wood Prod. 2018, 76, 1515.
28. Zhang, Y.; Ding, L.; Gu, J.; Tan, H.; Zhu, L. Carbohydr. 47. Hong, M. K.; Lubis, M. A. R.; Park, B. D.; Sohn, C. H.;
Polym. 2015, 115, 32. Roh, J. K. Wood Mater. Sci. Eng. 2018, 1. https://doi.org/
29. Lubis, M. A. R.; Park, B. D. J. Adhes. Sci. Technol. 2018, 32, 10.1080/17480272.2018.1528479.
2667. 48. Roffael, E.; Hüster, H. G. Eur. J. Wood Wood Prod. 2012,
30. Wicks, D. A.; Wicks, Z. W. Prog. Org. Coat. 2001, 41, 1. 70, 401.
31. Lubis, M. A. R.; Park, B. D.; Lee, S. M. Int. J. Adhes. Adhes. 49. Wan, H.; Wang, X. M.; Barry, A.; Shen, J. BioResources.
2017, 73, 118. 2014, 9, 7554.

47966 (10 of 10) J. APPL. POLYM. SCI. 2019, DOI: 10.1002/APP.47966

You might also like