Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

Chemical Engineering Journal 370 (2019) 420–431

Contents lists available at ScienceDirect

Chemical Engineering Journal


journal homepage: www.elsevier.com/locate/cej

Depolymerization and characterization of Acacia mangium tannin for the T


preparation of mussel-inspired fast-curing tannin-based phenolic resins
Jiongjiong Lia,b, Wenjie Zhua,b, Shifeng Zhanga,b, Qiang Gaoa,b, Changlei Xiac, Wei Zhanga,b, ,

Jianzhang Lia,b,

a
Ministry of Education Key Laboratory of Wood Materials Science and Utilization, Beijing Forestry University, Beijing 100083, China
b
Beijing Key Laboratory of Wood Science and Engineering, Beijing Forestry University, Beijing 100083, China
c
Department of Mechanical and Energy Engineering, University of North Texas, Denton, TX 76203, USA

HIGHLIGHTS GRAPHICAL ABSTRACT

• AMT was depolymerized under acidic


condition using 2-methylfuran to prepare
DAMT.
• DAMT has a similar catechol with
mussel protein and high reactivity.
• DAMT and PEI were used to prepare
phenolic resins with high crosslinking
density and good water resistance.
• The gel time decreased by 67.6% and
the bonding strength increased by
63.6% after the addition of DAMT.
• The developed adhesives exhibited
low formaldehyde emission and high
thermal stability.

ARTICLE INFO ABSTRACT

Keywords: Similar to the 3,4-dihydroxy-L-phenylalanine (DOPA) structures in mussel adhesive proteins (MAPs), tannin
Tannin-based phenolic resins contains abundant similar catechol structures in the flavonoid units, which might be developed into high-per-
Condensed tannin formance adhesives. In this study, Acacia mangium tannin (AMT) was depolymerized under acidic conditions
Depolymerization reactions with 2-methylfuran as a nucleophilic reagent. The depolymerized tannin (DAMT) was combined with poly-
DOPA structures
ethyleneimine (PEI) to prepare tannin-based phenolic resins to mimic the MAP systems. The investigation of the
Adhesion properties
chemical structures of tannin indicated the successful reaction of 2-methylfuran onto the flavonoid units to
increase the chemical reactivity, as evidenced by the MALDI-TOF-MS, GPC, 1H NMR, and FT-IR analyses. The as-
prepared PEI and DAMT-modified phenolic (DTPF-PEI) resin showed a high cross-linking and three-dimensional
network, resulting from the high reactivity of DAMT. Complex reactions were observed during the curing of the
as-prepared resin, including Michael addition and/or Schiff base formation reactions between the amino groups
of PEI and the ortho-quinone of tannins, as well as the condensation reactions between PEI and the phenolic
resins. Compared with tannin-phenol-formaldehyde (TPF) resin, the gel time of the DTPF-PEI resin decreased
from 630 s to 204 s (decreased by 67.6%), the mass loss after hydrolysis decreased by 26.3%, the bonding
strength increased by 63.6% when cured at 135 °C, and the formaldehyde emission decreased by 64.4%. Such
high performance might be contributed by the formation of highly cross-linked networks, as well as the hy-
drogen bonding and pi-cation interactions between the catechol structures and the adhered surfaces. With its


Corresponding authors at: Ministry of Education Key Laboratory of Wood Materials Science and Utilization, Beijing Forestry University, Beijing 100083, China.
E-mail addresses: zhangwei@bjfu.edu.cn (W. Zhang), lijzh@bjfu.edu.cn (J. Li).

https://doi.org/10.1016/j.cej.2019.03.211
Received 13 December 2018; Received in revised form 22 March 2019; Accepted 23 March 2019
Available online 25 March 2019
1385-8947/ © 2019 Published by Elsevier B.V.
J. Li, et al. Chemical Engineering Journal 370 (2019) 420–431

outstanding performance, including fast curing rate, excellent adhesion properties, low formaldehyde emission,
and high thermal stability, this newly developed adhesive has the potential to be used as an alternative of
conventional phenol-formaldehyde resins.

1. Introduction structures similar to DOPA. Considering the inherent structure features,


CTs might be used as adhesive materials to mimic the adhesive systems
Mussels, as typical promiscuous fouling marine organisms, can of MAPs.
firmly affix themselves to almost any kinds of underwater surfaces via Phenol-formaldehyde (PF) resins, with excellent mechanical prop-
byssus [1]. The attachment system is composed of several proteins, erties, good moisture and chemical corrosion resistance, have been used
commonly referred to as mussel adhesive proteins (MAPs) [2]. Detailed in many fields, including wood-based panel industry, aerospace, mili-
studies of the MAPs revealed that their key feature is the high content of tary industry, electronics, etc. [10]. Condensed tannins containing
the amino acid 3,4-dihydroxy-L-phenylalanine (DOPA, Scheme 1). It abundant phenols have been developed into tannin-based phenolic re-
was confirmed that the catecholic functionality of DOPA played a cri- sins [8,11]. Crude CTs with high molecular weight, great steric hin-
tical role in the robust adhesion [3]. The catecholic moiety in DOPA has drance, and far apart reactive sites [7], dramatically increase the
the capacity of establishing hydrogen bonds and π-π interactions to the viscosity of the synthesized phenolic resins, which reduce the adhesive
hydrophilic and/or organic surfaces [3]. The phenolic hydroxyl groups performance. Thus, depolymerization of CTs is suggested before being
in DOPA can undergo various oxidation reactions at an elevated pH, fabricated into resin [12]. Currently, acid-assisted depolymerization in
resulting in the formation of dopaquinone. The dopaquinone can react the presence of nucleophilic reagents is considered to be a promising
with catechol via reverse dismutation reactions, leading to the forma- method, in which the A-rings are protonated and the interflavanoid
tion of a cross-linked product [4]. In addition, the amino groups in bonds are cracked, resulting in the formation of terminal flavonoid
DOPA structures can also react with the dopaquinone through Michael units and extension units (carbocation intermediates). The carbocation
addition or Schiff base reactions, resulting in the formation of poly- intermediates subsequently react with the nucleophilic reagents to form
dopamine, which is able to affix to various underwater substrates [5]. flavonoid derivatives [13,14]. Two kinds of nucleophilic reagents have
Inspired by these findings, DOPA has been introduced into many ad- been explored to trap the carbocation intermediates, namely, mercap-
hesive materials to enhance adhesion properties [1]. However, the re- tans [15,16] and activated aromatic compounds [13]. For the mer-
latively high cost of DOPA has seriously limited its industrial applica- captans, the flavanyl-mercaptyl bonds of the thioetherified flavonoids
tions. Therefore, it is necessary to find an alternative that has similar are prone to be cleaved when the pH values are higher than 8.5 [17].
catecholic moiety structures but is less expensive. Although the CeC bonds between the carbocation intermediates and
Condensed tannins (CTs) are cost-efficient, natural environmentally the phenolic compounds (phloroglucinol [18] or pyrogallol [19]) are
friendly materials that contain 3 to 8 flavanoid repetition units [7], stable, such nucleophilic reagents are unsuitable for tannin depoly-
which show great potential as a structural mimic of DOPA due to their merization because they can be used as building blocks themselves.
catechol-laden structural features (Scheme 1). The most important re- Recently, an eco-friendly and cost-efficient nucleophile, 2-methylfuran,
actions associated with the A-rings in CTs are the polycondensation has been proved to be capable of depolymerizing CTs [12]. The depo-
reactions with aldehydes, based on which, aldehyde-based adhesives lymerized, fully biobased products are stable under alkaline conditions,
derived from CTs have been explored [8,9]. The B-rings are generally which makes them suitable building blocks for the synthesis of tannin-
composed of pyrogallol rings and catechol rings, which have chemical based adhesives. In addition, the furylated flavonoid derivatives have

Scheme 1. Flavonoid units in condensed tannins and the chemical structure of DOPA (※[6]).

421
J. Li, et al. Chemical Engineering Journal 370 (2019) 420–431

phloroglucinolic and/or resorcinolic A-rings, low molecular weight, Beijing, China) for 48 h. The obtained products were denoted as DAMT-
high reactivity, decreased steric hindrance, and catecholic moieties, 1, DAMT-2, DAMT-3, DAMT-4, DAMT-5, and DAMT-6, respectively.
which mimic the structures of DOPA. Prior to the freeze drying, DAMT-2 was dialyzed using dialysis bag (Mw:
Previous work reported that wood adhesives prepared by mixing 3000 Da) for 4 days, and the obtained products was denoted as DAMT-
polyethyleneimine (PEI) and demethylated lignin [20] or crude tannins d.
[21] can be used to mimic the adhesive systems of MAPs. However,
their wet shear strength is limited. The reason may be the high mole- 2.3. Synthesis of neat PF resin and tannin-based phenolic resins
cular weight and low reactivity of polyphenols, as well as the in-
effective use of the active sites in lignin and the active A-rings in CTs. The neat PF resin was synthesized using a molar ratio of 1:2.0 of
The B-rings in depolymerized tannins can be grafted onto the molecular phenol to formaldehyde based on our previous work [22,23]. First,
chains of phenolic resins via reactions between the active A-rings and 109 g phenol, 62.7 g formaldehyde aqueous solution and 11 g NaOH
formaldehyde to mimic the catechol moiety in DOPA, which should solution (40%) were mixed in a flask and heated at 85 °C for 50 min.
improve the performance of the synthesized tannin-based phenolic re- Next, 62.7 g formaldehyde aqueous solution and 11 g NaOH solution
sins. Moreover, the combination of PEI and depolymerized tannin- (40%) were added to the flask at 85 °C for 60 min. Then, 62.7 g for-
based phenolic resins can be used to simulate the adhesive systems of maldehyde aqueous solution and 30 g NaOH solution (40%) were added
MAPs; it has great potential for eliminating the disadvantages of using to the same flask at 85 °C for 50 min. Finally, the reaction solution was
high-molecule polyphenols, therefore yielding high-performance rapidly cooled to 40 °C to yield the neat PF resin.
tannin-based phenolic resins. For the synthesis of tannin-based phenolic resins, the substitution
In this study, Acacia mangium tannin (AMT, condensed tannin) was rate of phenol by AMT, DAMT-2, or DAMT-d was 40% by weight, which
acid depolymerized in the presence of the nucleophilic reagent 2-me- corresponded to the TPF, DTPF, or DTPF-d resins. The tannin-based
thylfuran. Then, the depolymerized AMT (DAMT) and PEI were used to phenolic resins were prepared using a molar ratio of 1:2.0 of phenol
prepare tannin-based phenolic resins to mimic the MAP adhesive sys- together with tannin to formaldehyde. First, 40 g tannin was added into
tems. Matrix-assisted laser desorption/ionization-time-of-flight mass 100 g distilled water, and the pH value of the mixture was adjusted to
spectrometry (MALDI-TOF-MS), gel permeation chromatography 7.0 by using 40% NaOH solution. Next, 60 g phenol, 95 g formaldehyde
(GPC), Fourier transform infrared (FT-IR), thermogravimetric analysis aqueous solution, 30 g distilled water, and 10 g 40% NaOH solution
(TGA), and 1H-nuclear magnetic resonance spectroscopy (1H NMR) were added to the flask at 90 °C for 10 min. Then, 70 g tannin solution
were used to study the depolymerization mechanisms. The synthesized was added to the flask at 90 °C for 30 min. Subsequently, 70 g tannin
tannin-based phenolic resins were investigated and evaluated, in terms solution, 10 g 40% NaOH solution, and 22.5 g formaldehyde aqueous
of their chemical structures, gelation time, thermal behavior, bonding solution were added to the flask at 90 °C for 30 min. After that, 20 g
strength, and formaldehyde emissions. formaldehyde aqueous solution and 30 g 40% NaOH solution were
added to the flask at 90 °C for 30 min. Finally, the reaction mixture was
2. Materials and methods rapidly cooled to 40 °C to yield the tannin-based phenolic resins.
TPF-PEI and DTPF-PEI resins were synthesized by mixing of PEI and
2.1. Materials TPF or DTPF resins using mass ratio of 1:2.0 of PEI to tannins. After
stirring (1000 rpm) at ambient temperature for 30 min, PEI modified
AMT was purchased from Guangxi Baise Forest Chemicals General tannin-based phenolic resins were prepared.
Plant (Guangxi, China). The moisture content of AMT is 10.1%, the
total extract content is 89.9%, the tannin content is 67.1%, the non- 2.4. Characterization of tannins and phenolic resins
tannin content is 27.3%, the insoluble content is 3.2%, and the pH value
is ca. 5.13. Solid phenol, 37 wt% formaldehyde aqueous solution, so- The MALDI-TOF-MS spectra of tannins were acquired on a MALDI-
dium carbonate, 2-methylfuran, hydrochloric acid (HCl, 37%), absolute TOF instrument (AXIMA Performance, Shimadzu, Japan). A pulsed ni-
ethyl alcohol, and sodium hydroxide were all analytical grade reagents trogen laser with a wavelength of 337 nm was used as the irradiation
provided by Beijing Chemical Reagents Co., Ltd. (China). PEI, a par- source. The duration of one laser pulse was 3 ns. The GPC analysis of
tially branched polyamine with an average molecular weight of 10, tannins was performed on PL-GPC 50 (Varian Inc., Palo Alto, USA). The
1
000, was purchased from Aladdin (Shanghai, China). All chemicals H NMR spectra of freeze-dried tannins were collected on a JEOL ECS-
were used as received without any further purification. 400 NMR spectrometer (JEOL, Tokyo, Japan) at a frequency of
300 MHz with an acquisition time of 3.0 s. D2O was used as the solvent.
2.2. Optimization of the AMT depolymerization reactions The FT-IR spectra were performed on a Nicolet 6700 spectrometer
(Nicolet Instrument Corporation, Madison, WI, USA) at a resolution of
The depolymerization reactions were conducted under acidic con- 4 cm−1. All the spectra were recorded in the frequency range of
ditions in a nitrogen atmosphere. In a typical depolymerization process, 500–4000 cm−1 with 32 scans. The TGA spectra were acquired on a
4 g AMT was added into 100 mL 2-methylfuran/absolute ethyl alcohol Q50 TGA analyzer (TA Instruments, Newcastle, DE, USA) in a nitrogen
solutions (2-methylfuran/absolute ethyl alcohol = 1/3, V/V). Then, 1 g atmosphere (60 mL·min−1) at a heating rate of 10 °C·min−1. The yield
37% HCl (i.e., 0.1 mol·L-1) was added. The reaction temperature was set of phenolic resins was calculated according to the weight ratio of the
to 50 °C. To investigate the effect of reaction time on the depolymer- phenolic resins and the raw materials used during the preparation
ization of tannins, the depolymerization reactions were conducted for process. The gelation time determination of the phenolic resins was
30 min, 60 min, 90 min, 120 min, 150 min, or 180 min under magnetic conducted according to the Chinese National Standard (GB/T 14074.3-
stirring at 400 rpm. To stop the depolymerization reaction, the reaction 2006). For this test, 5.0 g phenolic resins were added into a test tube,
solution was cooled to 0 °C in an ice bath, and then 75 mL sodium and the time period from the immersion of the test tube into an oil bath
carbonate aqueous solution (10.6 g·L-1) was added to neutralize the acid (135 °C) to the start of the gelation of the adhesive was defined as the
in the reaction solution. After the addition of sodium carbonate, the pH gelation time. The mass loss of cured phenolic resins after hydrolysis
of the reaction solution was adjusted to be weakly acidic (pH = ca. was determined according to previous work [24]. The phenolic resin
6.5). Ethanol and 2-methylfuran were evaporated under vacuum at samples were oven-dried at 120 °C and then ground into 200 mesh
room temperature, resulting in the purification of the obtained pro- powder. After that, the cured phenolic resins were wrapped using a
ducts. Finally, the depolymerized products were freeze-dried by placing qualitative filter paper and immersed in distilled water at 60 °C for 6 h.
them in a vacuum freeze-drier (Sihuan LGJ-25C, Sihuan Co., Ltd, Finally, the samples were dried at 105 °C for 3 h and then weighed. The

422
J. Li, et al. Chemical Engineering Journal 370 (2019) 420–431

weights change before and after hydrolysis was defined as the mass
loss. The accelerated aging test of the cured TPF-PEI resin was per-
formed using a QUV accelerated weathering tester (QUV/Spray, Q-Lab
Co., USA) for a total duration of 72 h, 144 h, and 216 h, respectively.
Each weathering cycle (12 h) consisted of 8 h of UV exposure at 60 °C
and 4 h condensation at 50 °C. The obtained products were denoted as
TPF-PEI-1, TPF-PEI-2, and TPF-PEI-3, respectively. The XPS measure-
ment was performed on a thermo scientific ESCALAB 250 (Thermo
Fisher Scientific, UK) equipped with monochromatic Al Kα radiation.

2.5. Preparation and characterization of plywood

Three-layer plywood with the dimension of 400 mm × 400 mm


× 4.5 mm was prepared using poplar veneers. First, each side of the
middle veneer was coated with (125–150) g·m−2 phenolic resins. The
glued plywood was then hot-pressed at 135 °C or 115 °C under 1.0 MPa Fig. 2. Typical polymer structures of flavonoid repeating units in AMT: (A)
for 5 min. Finally, the prepared plywood was stored at room tempera- fisetinidin, (B) catechin and robinetinidin, and (C) delphinidin.
ture for 2 days prior to use.
The shear strength of plywood (100 mm × 25 mm × 4.5 mm) was flavonoid unit [26]. These peaks revealed that catechin and robineti-
measured as per ASTM D906-98 A. The emission of formaldehyde from nidin are the dominant repeating units in the ATM, as evidenced by
plywood was measured by the method of acetylacetone according to their high intensity. A series of peaks at 1174.7 Da, 1478.5 Da,
previous work [25]. First, a glass dish containing 300 mL distilled water 1782.2 Da, and 2070.1 Da separated by ca. 304 Da were related to the
was placed in the bottom of a 10-L desiccator. Ten plywood samples occurrence of C-type flavonoid units [26]. The mass increment of ca.
(15 cm × 5 cm) were then placed in the same desiccator, which was 16 Da between adjacent peaks could be explained by different numbers
kept at 20 °C for 24 h. Finally, the formaldehyde concentration in the of hydroxyl groups attached to the flavonoid unit [27].
distilled water was determined using an ultraviolet spectroscope (UV) The number of repeating units present in the oligomers can be
at a wavelength of 412 nm. calculated using the following expression: M = 274.3A + 289.3B
+ 304.3C + 2.0 (endgroups, 2 × H), where A, B and C represent the
relevant flavonoid monomers. For instance, the peak at 1190.8 Da is a
3. Results and discussion
flavonoid tetramer according to two possible combinations of repeating
units such as (I) one A-type unit and three C-type units plus 2H end-
3.1. Depolymerization and characterization of tannins
groups; (II) two B-type units and two C-type units plus 2H endgroups.
The possible combinations of the three repeating flavonoid units and
3.1.1. Molecular mass measurements of tannins
the polymerization degrees are shown in Table S1. Analysis of the AMT
The MALDI-TOF-MS spectra of AMT are shown in Fig. 1. The re-
revealed that these tannins have oligomers of up to 9 repeating flavo-
petitive pattern of peaks observed in the MALDI-TOF-MS spectra was
noid units. These abundant aromatic building blocks can be used to
used to identify a specific oligomer series in AMT, including three types
prepare phenolic resins. Owing to the presence of fisetinidin and ca-
of flavonoid units, namely, fisetinidin (A, 274.3 Da), catechin and ro-
techin repeating units that contain catechol structures, AMTs have
binetinidin (B, 289.3 Da), and delphinidin (C, 304.3 Da) (Fig. 2). A
promising potentials to mimic the DOPA structures.
series of peaks at 903.1 Da, 1174.7 Da, 1446.3 Da, 2070.1 Da,
The MALDI-TOF-MS spectra of the depolymerized tannins are shown
2342.3 Da, and 2614.6 Da, which are separated by ca. 273 Da, indicated
in Fig. 3. Most of the peaks present in the spectra of DAMT differ from
the presence of A-type flavonoid units [26]. The low intensity of these
those in the spectra of AMT (Fig. 1). After depolymerization reactions, the
peaks suggested that small quantities of profisetinidin repeating units
dominate peaks decreased and many low molecular weight peaks
were present in the AMT. The peaks at 903.1 Da, 1190.8 Da, 1478.5 Da,
emerged. These observations indicated successful depolymerization of the
1766.1 Da, 2054.4 Da, 2342.3 Da, and 2629.9 Da, which were clearly
AMTs, which could also be confirmed by GPC and TGA measurements
separated by ca. 289 Da mass, indicated the presence of B-type

Fig. 1. MALDI-TOF-MS spectra of AMT. (a) 500 Da – 1750 Da and (b) 1750 Da – 2800 Da.

423
J. Li, et al. Chemical Engineering Journal 370 (2019) 420–431

Fig. 3. MALDI-TOF-MS spectra of DAMT: (a) DAMT-1, DAMT-2, and DAMT-3 in the range of 500 Da – 2800 Da; (b) DAMT-4, DAMT-5, and DAMT-6 in the range of
500 Da – 2800 Da; (c) DAMT-1, DAMT-2, and DAMT-3 in the range of 250 Da – 600 Da; (d) DAMT-4, DAMT-5, and DAMT-6 in the range of 250 Da – 600 Da.

(Fig. S1–S2 and Table S2). As shown in Fig. 3a and Fig. 3b, a series of series peaks at 354.1 Da – 356.0 Da, 363.6 Da – 364.3 Da, and 394.0 Da –
peaks with a repeating interval of ca. 81 Da were noted, which may be 394.3 Da could be associated with the furylated A-type, furylated B-type,
related to a 2-methylfuran unit (82.1 Da − 1H = 81.1 Da). It was con- and furylated C-type flavonoid monomer units, respectively. The in-
firmed that 2-methylfuran reacted with the carbocation intermediates to tegrated intensity of peaks associated with the flavonoid monomers and
form the nucleophile adducts. Peaks at 272.3 Da – 275.4 Da corresponded the furylated flavonoid monomers initially increased and then deceased as
to the occurrence of A-type flavonoid monomer units (fisetinidin), peaks at the reaction time increased. Series of peaks in the spectrum of DAMT-2
282.1 Da – 285.1 Da were related to the B-type flavonoid monomer units exhibited the highest integrated intensity of the depolymerized tannins,
(catechin and robinetinidin), and peaks at 311.8 Da – 312.5 Da indicated indicating that more flavonoid monomer units were obtained after reac-
the presence of the C-type flavonoid monomer units (delphinidin). The tion for 1.0 h.

Fig. 4. (a) FT-IR spectra of AMT and DAMT-2, (b) 1H NMR spectra of AMT and DAMT-2.

424
J. Li, et al. Chemical Engineering Journal 370 (2019) 420–431

As for the spectral range of 500 Da – 2800 Da (Fig. 3c and d), the phenomenon can be explained by the successful depolymerization of AMT.
intense peaks of the DAMT were clearly different, and they were dif- It can be clearly observed that the integrated intensity of band at ap-
ferent from those of the AMTs, which may be explained by the con- proximately 2925 cm−1 increased after the depolymerization reactions. A
densation and rearrangement of molecular fragments during the de- new peak at 1382 cm−1 associated with the methyl groups was present in
polymerization process [28]. The spectra clearly showed dominant the spectrum of DAMT-2, but absent in the spectrum of AMT. In addition,
fragment peaks at 533.6 Da – 595.9 Da, and exhibited oligomer series the absorption bands in the range of 1204–1030 cm−1 due to the CeOeC
with a polymerization degree of up to 3 – 4 repeating units. The linkages became stronger. These observations confirmed that 2-methyl-
polymerization degree decreased initially and then increased as the furan has been successfully added to the flavonoid units, indicating a
reaction time increased. Through the spectral analysis, the reaction positive effect of the nucleophilic reagent on the depolymerization of AMT.
time of 1.0 h seems to be the most effective reaction condition for AMT The 1H NMR spectra of AMT and DAMT-2 are shown in Fig. 4b. As
depolymerization, as evidenced by their low polymerization degree and seen in the spectrum of AMT, the protons linked to the C3 position and
high integrated intensity of dominant fragment peaks at 533.6 Da – protons in the residual carbohydrates can be observed between
595.9 Da. 3.45 ppm and 3.92 ppm. The protons at the C4 position when R1 = eH
Based on the abovementioned results, it can be concluded that the can be observed at 3.99 ppm [33]. The proton signals at 4.95 ppm in-
flavonoid monomer units, oligomers, and corresponding flavonoid de- dicated the presence of protons at the C4 position when R1 = another
rivatives coexisted in the depolymerized tannins. The as-prepared flavonoid unit [33]. The proton signals between 4.25 ppm and
DAMTs with low molecular weight, low steric hindrance, and catechol 4.35 ppm were attributed to the protons at the C2 position. The protons
structures might be fabricated into a high-performance phenolic resin. at the C6 position resonated at 5.82 ppm, and the proton signals at
5.99 ppm were attributed to the protons at the C8 position [34]. The
proton signals in the range of 6.25–6.82 ppm were assigned to the
3.1.2. FT-IR and 1H NMR measurements of tannins
aromatic protons in the flavonoid units. The protons in the hydroxyl
Motivated by the above results, the optimum reaction conditions for
groups can be observed at 8.31 ppm [33,34].
efficient depolymerization of AMT into flavonoid monomers and/or
Compared with the AMT, some clear differences were observed after
oligomers were the use of 2-methylfuran in 0.1 mol·L-1 HCl solution at
depolymerization. For instance, the intensity of the proton signal at
50 °C for 60 min. The chemical structures of AMT and DAMT-2 were
8.31 ppm increased in the DAMT-2, indicating an increased content of
analyzed by FT-IR (Fig. 4a) and 1H NMR (Fig. 4b). As seen in Fig. 4a, hydroxyl groups. The intensity of the proton signal at 4.95 ppm de-
the broad peak at approximately 3303 cm−1 was related to the eOH
creased after the depolymerization reactions, and the DAMT-2 showed
stretching vibrations. The absorption band at 2939 cm−1 was asso- higher intensity of the proton signal at 3.99 ppm than the AMT. These
ciated with the CeH stretching vibration of the eCH and eCH2 groups
results indicated that the interflavanoid bonds connecting the flavonoid
of the aliphatic hydrocarbon [29]. The absorption bands at 1616 cm−1, units are cleaved, resulting in the formation of flavonoid monomer
1512 cm−1, and 1453 cm−1 were the characteristic bands of aromatic
units and oligomers. It can be clearly observed that new proton signals
ring stretching vibration. The high integrated intensity of the band at associated with protons in the methyl groups and at the C3′' positions
1616 cm−1 associated with the CeC stretching vibrations revealed that
are present in the spectrum of the DAMT-2. These results confirmed
the C4eC8 bond was the dominated interflavonoid linkage in AMT
that the 2-methylfuran units have been successfully added to the fla-
[30]. The single absorption band at 1512 cm−1 indicated that the AMT
vonoid units of tannins, which is consistent with the above results. The
was mainly composed of procyanidin, which was consistent with the
depolymerization mechanism is illustrated in Scheme 2. Under acidic
MALDI-TOF-MS data [31]. The peaks in the region of 1204–1030 cm−1
conditions, the interflavonoid bonds between the subunits are cleaved
indicated the CeOeC linkages, including aromatic CeO and aliphatic
through a heterolytic process, resulting from the protonation of A-rings.
CeO present in the AMT [28]. The absorption bands in the range of
The terminal flavonoid and extension subunits are released as neutral
841–616 cm−1 were related to the CeH of benzene rings and the OeH
monomeric flavan-3-ols and carbocation intermediates, respectively.
of alcohol vibration [32].
Subsequently, the carbocations are trapped by 2-methylfuran, resulting
After the depolymerization reactions, some apparent differences were
in the formation of flavonoid derivatives.
observed. As seen in the spectrum of DAMT-2, the bands associated with
Based on the abovementioned results, it can be concluded that
the eOH stretching vibrations became broader and shifted from
DAMT-2 had low molecular weight, low steric hindrance, high re-
3303 cm−1 to 3372 cm−1, suggesting an increased hydrogen bonding in-
activity, and good thermal stability (Fig. S2), and it can be used as
teractions and an increase in the content of hydroxyl groups. This

Scheme 2. The depolymerization mechanism of AMT using 2-methylfuran as a nucleophilic reagent under acidic conditions.

425
J. Li, et al. Chemical Engineering Journal 370 (2019) 420–431

promising building block to prepare low-cost tannin-based phenolic between 3346 cm−1 and 3278 cm−1, CeH stretching vibrations in
resins. The highly reactive phloroglucinolic and resorcinolic A-rings can eCH2 groups between 2935 cm−1 and 2814 cm−1, NeH bending vi-
react with formaldehyde. Hence, the catechol units (B-rings) in DAMT-2 brations in eNH2 groups at 1595 cm−1, CeN stretching vibration at
can be grafted onto the molecular chains of the phenolic resin, and the 1457 cm−1, and NeH bending vibrations in eNH groups at 1119 cm−1
high content of catechol rings can be used to mimic the DOPA struc- [39]. Amino groups of PEI impart the ability to react with ortho-qui-
tures. To increase their similarity to the MAP adhesive systems, poly- none structures via Michael addition and Schiff base formation reac-
amines can be added to the tannin-based phenolic resins. The catechol tions, which have been used to prepare wood adhesives by combining
units are readily oxidized to form ortho-quinone structures under al- with poly (N-acryloyldopamine) [40] or demethylated lignins [20]. In
kaline conditions [35]. Subsequently, the ortho-quinone structures can addition, the combination of PEI and Douglas-fir condensed tannins to
react with polyamine via Michael addition and/or Schiff base formation manufacture wood adhesives were also explored [21]. In this approach,
reactions [36]. These reactions can result in the formation of a cross- the depolymerized AMT was used to prepare tannin-based phenolic
linked and three-dimensional network, which are believed to enhance resins, and the combination of PEI and such phenolic resins was used to
the properties of the synthesized phenolic resins. These reaction me- increase their similarity to the MAP adhesive systems and enhance
chanisms are depicted in Scheme 3. adhesion properties. After the addition of PEI, the amino groups in PEI
reacted with the ortho-quinone in the tannin framework, resulting in
the formation of C]N and/or CeN groups, as shown in Scheme 3. The
3.2. Chemical structure measurement of the tannin-based phenolic resins amino groups can also react with the methylol groups in the phenolic
resins, leading to the formation of CeN groups [41]. However, no ab-
The AMT and DAMT-2 were used to prepare tannin-based phenolic sorption bands associated with such groups were observed in the
resins, and the obtained resins were denoted as TPF and DTPF, re- spectra of TPF-PEI and DTPF-PEI resins, expect for the enhanced peak
spectively. To mimic the MAP adhesive systems, PEI, a polyamine, was at 2830 cm−1. The reason may be the overlaps of bands between the
added to the synthesized tannin-based phenolic resins. In addition, the C]N (ca. 1665 cm−1), CeN (ca. 1460) and the aromatic rings
DAMT-d was also used to prepare tannin-based phenolic resin, and the (1700–1400 cm−1) in the phenolic resins [42,43]. The absorption
obtained resin was denoted as DTPF-d. The FT-IR spectra of the cured bands related to eNH2 groups (3346–3278 cm−1) in the FT-IR spectrum
phenolic resins and PEI are shown in Fig. 5. As shown in Fig. 5a, all the of PEI were sharp, and the TPF-PEI and DTPF-PEI resins exhibited wide
tannin-based phenolic resins exhibited similar FT-IR absorption profiles and flat absorption bands in the range of 3700–3300 cm−1, which may
compared to the PF resin, because of their similar chemical structures. be because there were no or very little unreacted PEI in the phenolic
In addition, DTPF and DTPF-d resins had similar chemical structures resin systems.
based on the FT-IR analysis. The absorption bands in the range of
3700–3300 cm−1 and 2950–2800 cm−1 were related to the hydroxyl
groups and the CeH stretching vibrations in the eCH2 groups, re- 3.3. Thermal properties of the tannin-based phenolic resins
spectively. All the phenolic resins exhibited typical aromatic rings in
the region 1700–1400 cm−1, phenolic CeO stretching at 1208 cm−1, The TGA and DTG curves of the phenolic resins are shown in Fig. 6. For
and CeO stretching at 1010 cm−1 [37]. The presence of aromatic CeH the PF, TPF, DTPF, and DTPF-d resins, three main mass loss events, i.e.,
was observed at 878 cm−1, 775 cm−1, and 708 cm−1 [38]. Compared below 250 °C, (250–450) °C, and (450–600) °C were observed. The con-
with the spectrum of PF resins, the integral intensity of bands at densation reaction between methylol groups caused the mass loss in the
1464 cm−1, 1438 cm−1, and 1208 cm−1 became weaker in the spectra first thermal event [44]. The condensation reactions of methylol and
of TPF, DTPF, and DTPF-d resins. These observations were accounted phenolic hydrogen, as well as between two hydroxyl functional groups
for the high content of nontannin materials in the tannin samples. resulted in the mass loss in the second thermal event [44]. Degradation of
As shown in Fig. 5b, spectrum of PEI showed typical NeH bands methylene groups led to the mass loss in the third thermal event [23].

Scheme 3. Reaction mechanisms of the depolymerized tannins.

426
J. Li, et al. Chemical Engineering Journal 370 (2019) 420–431

Fig. 5. (a) FT-IR spectra of phenolic resins and (b) FT-IR spectrum of PEI.

Compared with the PF resin, the T10% increased from 193.5 °C to was close to 100%. The mass loss after hydrolysis of the cured phenolic
239.5 °C for the TPF resin, to 227.1 °C for the DTPF resin, and to resins is shown in Fig. 7b. The mass loss of the PF resin was 21.87%,
228.0 °C for the DTPF-d resin; the weight residual at 600 °C increased while the mass losses of the TPF and DTPF resins were 27.20% and
from 67.2% to 68.4% for the TPF resin, to 68.7% for the DTPF resin, 25.67%, respectively. The increased mass loss of the TPF and DTPF
and to 69.0% for the DTPF-d resin. DTPF and DTPF-d resins had similar resins may be caused by the presence of soluble carbohydrates in
thermal stability based on the TGA analysis. The T10% of the TPF-PEI condensed tannins [28]. Compared with the TPF resin, the decreased
and DTPF-PEI resins was 214.6 °C and 258.6 °C, respectively. The PEI- mass loss of the DTPF resin may be explained by the reduced molecular
modified phenolic resins exhibited inferior thermal stability in the weight and enhanced reactivity of the DAMT. The mass loss of the
second thermal event, which may be caused by the degradation of the DTPF-d resin was 25.08%, which was similar to that of the DTPF resin.
PEI framework [45]. The residual weight of the DTPF-PEI resin at After the incorporation of PEI, the mass loss of TPF-PEI and DTPF-PEI
600 °C was higher than that of the TPF-PEI resin. These results led us to resins decreased to 24.66% and 20.05%, respectively. This is because
the conclusions that the DTPF and DTPF-d resins had good thermal PEI cross-linked with the molecular chains of the phenolic resin, in-
stability, as evidenced by the high residual weight at 600 °C. The DTPF- creased the cross-linking density and reduced the content of small so-
PEI resin exhibited the best thermal stability in the first thermal event, luble molecules (e.g., free formaldehyde) of the phenolic resins. In
as evidenced by the high T10%. The high thermal stability of the DTPF- addition, the enhanced cross-linking density can improve water re-
PEI resin can be explained by the highly cross-linked resin network, sistance of the phenolic resins and limit the dissolution of small mole-
which can be confirmed by FT-IR measurement. cules from the resins.
Fig. 8a shows the gel time and the viscosity of the tannin-based
phenolic resins. The viscosity of the TPF resin was higher than that of
3.4. Physical properties of the tannin-based phenolic resins the PF resin, possibly due to the higher molecular weight of the AMT
than phenol [28]. The DTPF resin exhibited lower viscosity than the
The yield of the phenolic resins is shown in Fig. 7a. As seen, the TPF resin, caused by the decreased molecular weight of the DAMT. The
yield of the PF, TPF, DTPF, and DTPF-d resins was 99.43%, 99.31%, viscosity of the DTPF-d resin was approximately 166 mPa·s, which was
99.35%, and 99.33%, respectively. During the resin preparation pro- similar to that of the DTPF resin (ca. 173 mPa·s). With the addition of
cess, phenols, formaldehyde, and catalysts were added to a flask PEI, the viscosity of the tannin-based phenolic resins significantly in-
equipped with a condenser, and there were almost no raw materials creased. This might result from the increased hydrogen bonding
volatilized from the flask. Therefore, the yield of the phenolic resins

Fig. 6. (a) TGA and (b) DTG curves of the phenolic resins.

427
J. Li, et al. Chemical Engineering Journal 370 (2019) 420–431

Fig. 7. (a) The yield of phenolic resins; (b) The mass loss after hydrolysis of the cured phenolic resins.

interactions between the amino groups of PEI and the hydroxyl groups reactivity of the DAMT and the fast curing rate of the DTPF resin. The
of the phenolic resins, because the depolymerization of tannins resulted bonding strength and formaldehyde emission of the DTPF-d resin were
in an increased content of hydroxyl groups as confirmed by FTIR and 1H 1.06 MPa and 0.35 mg/L, respectively, which were similar to those of
NMR measurements (Fig. 4). Therefore, higher viscosity was de- the DTPF resin. The bonding strength of the tannin-based phenolic re-
termined for the DTPF-PEI resin, compared with the TPF-PEI resin sins increased with the addition of PEI, achieving ca. 1.17 MPa and ca.
(Fig. 8a). As a consequence, the gel time of the TPF resin decreased due 1.44 MPa for the TPF-PEI and DTPF-PEI resins, respectively. The for-
to the high viscosity of the TPF resin and the high molecular weight of maldehyde emission of the TPF-PEI and DTPF-PEI resins decreased after
the AMT. The gel time of the DTPF and DTPF-d resins were 552 s and the addition of PEI, which can be explained by the reactions between
560 s, respectively, which were shorter than that of the TPF resin, al- amino groups and the phenolic resins [41]. It is worth noting, the DTPF-
though the DTPF and DTPF-d resins had low viscosities. These ob- PEI resin had higher bonding strength and lower formaldehyde emis-
servations may be explained by the high reactivity of the depolymer- sion than the neat PF resin.
ized tannins, which benefits from the depolymerization reaction. The Due to the low curing temperature of the MAP adhesives, the in-
PEI-modified phenolic resins exhibited shorter gel time than the other troduction of the simulated DOPA structures and polyamine was sup-
phenolic resins, due to their high viscosity and the reactions between posed to decrease the curing temperature of the phenolic resins. To
the PEI and the tannin-based phenolic resins [20]. Notably, compared evaluate whether the synthesized tannin-based phenolic resins can be
with the PF resin, the gel time of the DTPF-PEI resin decreased from ca. cured at low temperature, the bonding strength of plywood prepared by
702 s to ca. 204 s. These results led us to the conclusion that the ad- hot pressing at 115 °C was investigated. As shown in Fig. 8b, the
hesives synthesized by this approach can promote the preparation of bonding strength of TPF, DTPF, and DTPF-d resins were 0.54 MPa,
fast-curing tannin-based phenolic resins. 0.62 MPa, and 0.61 MPa, respectively, all below 0.70 MPa. After the
The bonding strength and formaldehyde emission of the phenolic addition of PEI, the bonding strengths of TPF-PEI and DTPF-PEI resins
resins are shown in Fig. 8b. After being cured at 135 °C, the bonding increased to 0.82 MPa and 1.14 MPa, respectively. It is worth noting
strength of phenolic resins was higher than the requirement for ex- that, the DTPF-PEI resin exhibited higher bonding strength than the
terior-grade plywood (0.70 MPa) according to the Chinese National neat PF resin (0.98 MPa), which was similar to the trend observed when
Standard (GB/T 14732-2017), and the formaldehyde emissions were all cured at 135 °C. This means that the DTPF-PEI resin has high adhesion
below 0.50 mg/L, meeting the E0 grade [25]. The bonding strength and strength and can be cured at low temperature, which has great potential
formaldehyde emission of TPF resin were 0.88 MPa and 0.45 mg/L, and in industrial applications.
those of the DTPF resin were 1.07 MPa and 0.32 mg/L, respectively. The properties of the DTPF resin were similar to that of the DTPF-d
The enhanced properties of the DTPF resin were caused by the high resin according to the above-mentioned results. Therefore, the dialysis

Fig. 8. (a) Gel time and viscosity of the phenolic resins; (b) Bonding strength and formaldehyde emission of the phenolic resins.

428
J. Li, et al. Chemical Engineering Journal 370 (2019) 420–431

of the depolymerized tannin had little effect on the properties of the of aging time, which was consistent with the FT-IR measurement. It is
prepared tannin-based phenolic resins. noteworthy that the content of nitrogen atom increased as the increase
The adhesive mechanisms of the DTPF-PEI resins are summarized in of aging time. This observation may be because the PEI molecular
Scheme 4. Compared with the TPF resin, the increased adhesion chains did not release from the TPF-PEI resin systems during the ac-
properties of the DTPF-PEI resin can be explained by three main rea- celerated aging test. In this approach, the dosage of PEI in the tannin-
sons. First, the depolymerization of AMT decreased the molecular based phenolic resins was low, PEI can react with the phenolic resin
weight and increased the content of reactive functional groups. The systems through Michael addition, Schiff base formation, and con-
increased reactive sites promoted the methylolation and polymerization densation reactions, the amino groups in PEI can form strong H-
reactions between tannins and formaldehyde, resulting in the formation bonding with the hydroxyl groups in the phenolic resin systems, and the
of a network structure with high cross-linking density. Second, the branched chain of PEI can entangle with the molecular chains of phe-
amino groups in PEI can react with the ortho-quinone in tannins nolic resins. In addition, after curing of the phenolic resins, the liquid
through Michael addition and/or Schiff base formation reactions, in adhesives were converted to a stable, three-dimensional network.
addition to the condensation reactions with the methylol groups in the Therefore, it is not easy for the PEI molecular chains to release from the
phenolic resins [21]. All these reactions can increase the cross-linking phenolic resin systems during the accelerated aging test.
density and enhance the physical and/or chemical crosslinking between
the tannin-based phenolic resins and the adhered surfaces. Third, the
catechol structures in the cross-linked networks can interact with the 4. Conclusions
adhered surfaces via hydrogen bonding and pi-cation interactions [1],
which can further enhance the adhesion properties of the phenolic re- Tannin-based phenolic resins have been successfully fabricated by
sins. using depolymerized natural AMT. PEI was subsequently added to the
synthesized phenolic resins to mimic the MAP adhesive systems. The
MALDI-TOF-MS measurement revealed that the flavonoid repeating
3.5. Accelerated aging test of the tannin-based phenolic resin units forming the AMT were (A) fisetinidin, (B) catechin and robineti-
nidin, and (C) delphinidin. The AMT was successfully depolymerized
The FT-IR spectra of the accelerated aged TPF-PEI resins are shown and the 2-methylfuran was added to the flavonoid units in DAMT. The
in Fig. 9a. After the accelerated aging test, the accelerated aged TPF-PEI enhanced reactivity of DAMT, the Michael addition and/or Schiff base
resins exhibited decreased content of hydroxyl groups (approximately formation reactions between PEI and tannins, and the condensation
at 3191 cm−1) and methylol groups (approximately at 1010 cm−1) reactions between the amino groups and the methylol groups facilitated
[37]. In addition, absorption bands related to the CeO stretching vi- the formation of cross-linked, three-dimensional phenolic resin network
brations gradually disappeared as the increase of accelerated aging structures with high cross-linking density. Compared with the PF resin,
time. The intensity of peaks associated with the aromatic rings gradu- the gel times of the TPF-PEI and DTPF-PEI resins decreased from 702 s
ally decreased and new peaks related to the C]O groups generated in to 322 s (decreased by 54.1%) and 204 s (decreased by 70.9%), re-
the spectra of the accelerated aged TPF-PEI resins [38]. These ob- spectively. Compared with the TPF resin, the mass loss after hydrolysis
servations indicated that TPF-PEI resins were gradually degraded of the DTPF-PEI resin decreased from 27.20% to 20.05% (decreased by
during the accelerated aging test. 26.3%), the bonding strength increased from 0.88 MPa to 1.44 MPa
Fig. 9b shows the XPS spectra of the accelerated aged TPF-PEI re- (increased by 63.6%), and the formaldehyde emission decreased from
sins. As seen, the relative content of carbon atom increased first and 0.45 mg/L to 0.16 mg/L (decreased by 64.4%), which were close to
then decreased and the relative content of oxygen atom decreased first those of the neat PF resin. When cured at 115 °C, the bonding strength
and then increased as the aging time increased. These observations of the DTPF-PEI resin was 1.14 MPa, which was higher than that of the
indicated the TPF-PEI resins were gradually decomposed as the increase neat PF resin (0.98 MPa) and the TPF resin (0.54 MPa). The strong

Scheme 4. Adhesion mechanisms of the DTPF-PEI resins.

429
J. Li, et al. Chemical Engineering Journal 370 (2019) 420–431

Fig. 9. (a) FT-IR spectra and (b) XPS spectra of the accelerated aged TPF-PEI resins.

bonding strength resulted from the highly cross-linked network and the Synthesis and characterization of phenolic novolacs modified by chestnut and mi-
hydrogen bonding and pi-cation interactions between the catechol mosa tannin extracts, J. Appl. Polym. Sci. 100 (2006) 4412–4419.
[12] G.B. Laurent Rouméas, Chahinez Aouf, Eric Dubreucq, and Helene Fulcrand,
structures and the adhered surfaces. The tannin-based adhesives syn- Furylated flavonoids: fully biobased building blocks produced by condensed tannins
thesized in this work exhibited a fast curing rate, good adhesion depolymerization, ACS Sustainable, Chem. Eng. 6 (2018) 1112–1120.
properties, low formaldehyde emission, and high thermal stability, [13] L.Y. Foo, L.J. Porter, Prodelphinidin polymers: definition of structural units, J.
Chem. Soc., Perkin Trans. 1 (1978) 1186–1190.
which have great potential for industrial applications. [14] R.S. Thompson, D. Jacques, E. Haslam, R.J.N. Tanner, Plant proanthocyanidins.
Part 5. Sorghum polyphenols, J. Chem. Soc., Perkin Trans. 1 (1978) 892–896.
Conflicts of interest [15] M.J. Betts, B.R. Brown, P.E. Brown, W.T. Pike, Degradation of condensed tannins:
structure of the tannin from common heather, Chem. Commun. (London) (1967)
1110–1112.
There are no conflicts to declare. [16] X.S. Ariadna Selga, Raül Bobet, Josep Lluís Torres, Efficient one pot extraction and
depolymerization of grape (vitis vinifera) pomace procyanidins for the preparation
of antioxidant thio-conjugates, J. Agric. Food Chem. 52 (2004) 467–473.
Acknowledgements
[17] Richard W. Hemingway, Joseph J. Karchesy, Gerald W. McGraw, Richard
A. Wielesek, Heterogeneity of interflavanoid bond location in loblolly pine bark
This work was financially supported by the National Natural Science procyanidins, Phytochemistry 22 (1983) 275–281.
Foundation of China [Project 31700491], and the Fundamental [18] James A. Kennedy, Graham P. Jones, Analysis of proanthocyanidin cleavage pro-
ducts following acid-catalysis in the presence of excess phloroglucinol, J. Agric.
Research Funds for the Central Universities [No. 2016ZCQ01]. Food Chem. 49 (2001) 1740–1746.
[19] M. Bordiga, J.D. Coïsson, M. Locatelli, M. Arlorio, F. Travaglia, Pyrogallol: an
Appendix A. Supplementary data Alternative Trapping Agent in Proanthocyanidins Analysis, Food Analytical
Methods 6 (2012) 148–156.
[20] Y. Liu, K. Li, Preparation and characterization of demethylated lignin-poly-
Supplementary data to this article can be found online at https:// ethylenimine adhesives, J Adhesion. 82 (2006) 593–605.
doi.org/10.1016/j.cej.2019.03.211. [21] K. Li, X. Geng, J. Simonsen, J. Karchesy, Novel wood adhesives from condensed
tannins and polyethylenimine, Int. J. Adhes. Adhes. 24 (2004) 327–333.
[22] J. Li, J. Zhang, S. Zhang, Q. Gao, J. Li, W. Zhang, Alkali lignin depolymerization
References under eco-friendly and cost-effective NaOH/urea aqueous solution for fast curing
bio-based phenolic resin, Ind. Crop. Prod. 120 (2018) 25–33.
[23] W. Zhang, Y. Ma, Y. Xu, C. Wang, F. Chu, Lignocellulosic ethanol residue-based
[1] J.S. Poseu, J.M. Aracil, F. Nador, F. Busqué, D.R. Molina, The Chemistry behind
lignin–phenol–formaldehyde resin adhesive, Int. J. Adhes. Adhes. 40 (2013) 11–18.
Catechol-Based Adhesion, Angew, Chem, 2018.
[24] J. Luo, J. Luo, X. Li, K. Li, Q. Gao, J. Li, Toughening improvement to a soybean
[2] J. Yu, W. Wei, E. Danner, R.K. Ashley, J.N. Israelachvili, J.H. Waite, Mussel protein
meal-based bioadhesive using an interpenetrating acrylic emulsion network, J.
adhesion depends on interprotein thiol-mediated redox modulation, Nat. Chem.
Mater. Sci. 51 (2016) 9330–9341.
Biol. 7 (2011) 588–590.
[25] W. Zhang, Y. Ma, C. Wang, S. Li, M. Zhang, F. Chu, Preparation and properties of
[3] Q. Ye, F. Zhou, W. Liu, Bioinspired catecholic chemistry for surface modification,
lignin–phenol–formaldehyde resins based on different biorefinery residues of agri-
Chem. Soc. Rev. 40 (2011) 4244–4258.
cultural biomass, Ind. Crop. Prod. 43 (2013) 326–333.
[4] E.P. Holowka, T.J. Deming, Synthesis and crosslinking of L-DOPA containing
[26] S. Abdalla, A. Pizzi, N. Ayed, F. Charrier, F. Bahabri, A. Ganash, MALDI-TOF and
polypeptide vesicles, Macromol. Biosci. 10 (2010) 496–502.
13C NMR analysis of Tunisian Zizyphus jujuba root bark tannins, Ind. Crop. Prod.
[5] F. Yu, S. Chen, Y. Chen, H. Li, L. Yang, Y. Chen, Y. Yin, Experimental and theoretical
59 (2014) 277–281.
analysis of polymerization reaction process on the polydopamine membranes and
[27] Y.B. Hoong, A. Pizzi, P.Md. Tahir, H. Pasch, Characterization of Acacia mangium
its corrosion protection properties for 304 Stainless Steel, J Mol. Struct. 982 (2010)
polyflavonoid tannins by MALDI-TOF mass spectrometry and CP-MAS 13C NMR,
152–161.
Eur. Polym. J. 46 (2010) 1268–1277.
[6] B.P. Lee, P.B. Messersmith, J.N. Israelachvili, J.H. Waite, Mussel-inspired adhesives
[28] A. Zhang, J. Li, S. Zhang, Y. Mu, W. Zhang, J. Li, Characterization and acid-cata-
and coatings, Annu. Rev. Mater. Res. 41 (2011) 99–132.
lysed depolymerization of condensed tannins derived from larch bark, RSC Adv. 7
[7] A. Arbenza, L. Avérous, Chemical modification of tannins to elaborate aromatic
(2017) 35135–35146.
biobased macromolecular architectures, Green Chem. 17 (2015) 2626–2646.
[29] W.J. Lee, W.C. Lan, Properties of resorcinol-tannin-formaldehyde copolymer resins
[8] S. Feng, S. Cheng, Z. Yuan, M. Leitch, C. Xu, Valorization of bark for chemicals and
prepared from the bark extracts of Taiwan acacia and China fir, Bioresour. Technol.
materials: A review, Renew. Sust. Energ. Rev. 26 (2013) 560–578.
97 (2006) 257–264.
[9] M.C. Basso, C. Lacoste, A. Pizzi, E. Fredon, L. Delmotte, MALDI-TOF and 13C NMR
[30] C.W. Oo, M.J. Kassim, A. Pizzi, Characterization and performance of Rhizophora
analysis of flexible films and lacquers derived from tannin, Ind. Crop. Prod. 61
apiculata mangrove polyflavonoid tannins in the adsorption of copper (II) and lead
(2014) 352–360.
(II), Ind. Crop. Prod. 30 (2009) 152–161.
[10] G. Foyer, B.H. Chanfi, B. Boutevin, S. Caillol, G. David, New method for the
[31] L.Y. Foo, Proanthocyanidins: Gross chemical structures by infrared spectra,
synthesis of formaldehyde-free phenolic resins from lignin-based aldehyde pre-
Phytochemistry 20 (1981) 1397–1402.
cursors, Eur. Polym. J. 74 (2016) 296–309.
[32] R. Zhou, S. Si, Q. Zhang, Water-dispersible hydroxyapatite nanoparticles synthe-
[11] C. Peña, M. Larrañaga, N. Gabilondo, A. Tejado, J.M. Echeverria, I. Mondragon,
sized in aqueous solution containing grape seed extract, Appl. Surf. Sci. 258 (2012)

430
J. Li, et al. Chemical Engineering Journal 370 (2019) 420–431

3578–3583. for recognition of microRNA, Biomaterials 32 (2011) 3875–3882.


[33] E.C. Ramires, E. Frollini, Tannin–phenolic resins: Synthesis, characterization, and [40] K.L.C. Zhang, J. Simonsen, A novel wood-binding domain of a wood–plastic cou-
application as matrix in biobased composites reinforced with sisal fibers, Compos. pling agent: Development and characterization, J. Appl. Polym. Sci. 89 (2003)
Part B-Eng. 43 (2012) 2851–2860. 1078–1084.
[34] V.C. Jean-Marc Souquet, Franck Brossaud, Michel Moutounet, Polymeric proan- [41] Y.B. Hoong, A. Pizzi, L.A. Chuah, J. Harun, Phenol–urea–formaldehyde resin co-
thocyanidins from grape skins, Phytochemistry 43 (1996) 509–512. polymer synthesis and its influence on Elaeis palm trunk plywood mechanical
[35] A. Duval, L. Averous, Characterization and Physicochemical Properties of performance evaluated by 13 C NMR and MALDI-TOF mass spectrometry, Int. J.
Condensed Tannins from Acacia catechu, J. Agric. Food Chem. 64 (2016) Adhes. Adhes. 63 (2015) 117–123.
1751–1760. [42] Y. Lv, H.C. Yang, H.Q. Liang, L.S. Wan, Z.K. Xu, Nanofiltration membranes via co-
[36] H.C. Yang, K.J. Liao, H. Huang, Q.Y. Wu, L.S. Wan, Z.K. Xu, Mussel-inspired deposition of polydopamine/polyethylenimine followed by cross-linking, J. Membr.
modification of a polymer membrane for ultra-high water permeability and oil-in- Sci. 476 (2015) 50–58.
water emulsion separation, J. Mater. Chem. A 2 (2014) 10225–10230. [43] Y. Hu, K. Cai, Z. Luo, R. Hu, Construction of Polyethyleneimine-β-cyclodextrin/
[37] S. Cheng, I. D'Cruz, Z. Yuan, M. Wang, M. Anderson, M. Leitch, C.C. Xu, Use of pDNA Multilayer Structure for Improved In Situ Gene Transfection, Adv. Eng.
biocrude derived from woody biomass to substitute phenol at a high-substitution Mater. 12 (2010) B18–B25.
level for the production of biobased phenolic resol resins, J. Appl. Polym. Sci. 121 [44] J. Li, J. Zhang, S. Zhang, Q. Gao, J. Li, W. Zhang, Fast curing bio-based phenolic
(2011) 2743–2751. resins via lignin demethylated under mild reaction condition, Polymers 9 (2017)
[38] B. Li, S.H. Feng, H.S. Niasar, Y.S. Zhang, Z.S. Yuan, J. Schmidt, C. Xu, Preparation 428.
and characterization of bark-derived phenol formaldehyde foams, RSC Adv. 6 [45] M. Shabanian, M. Khoobi, F. Hemati, H.A. Khonakdar, S.e.S. ebrahimi,
(2016) 40975–40981. U. Wagenknecht, A. Shafiee, NewPLA/PEI-functionalized Fe3O4 nanocomposite:
[39] H. Dong, L. Ding, F. Yan, H. Ji, H. Ju, The use of polyethylenimine-grafted graphene Preparation and characterization, J. Ind. Eng. Chem. 24 (2015) 211–218, https://
nanoribbon for cellular delivery of locked nucleic acid modified molecular beacon doi.org/10.1016/j.jiec.2014.09.032.

431

You might also like