Anda di halaman 1dari 9

Article

pubs.acs.org/Biomac

Biobased Epoxy Resins from Deconstructed Native Softwood Lignin


Daniel J. van de Pas* and Kirk M. Torr
Scion, Private Bag 3020, Rotorua 3046, New Zealand

ABSTRACT: The synthesis of novel epoxy resins from lignin


hydrogenolysis products is reported. Native lignin in pine
wood was depolymerized by mild hydrogenolysis to give an oil
product that was reacted with epichlorohydrin to give epoxy
prepolymers. These were blended with bisphenol A diglycidyl
ether or glycerol diglycidyl ether and cured with diethylenetri-
amine or isophorone diamine. The key novelty of this work
lies in using the inherent properties of the native lignin in
preparing new biobased epoxy resins. The lignin-derived epoxy
prepolymers could be used to replace 25−75% of the
bisphenol A diglycidyl ether equivalent, leading to increases of up to 52% in the flexural modulus and up to 38% in the
flexural strength. Improvements in the flexural strength were attributed to the oligomeric products present in the lignin
hydrogenolysis oil. These results indicate lignin hydrogenolysis products have potential as sustainable biobased polyols in the
synthesis of high performance epoxy resins.

■ INTRODUCTION
With the growth of the biobased economy, the chemical sector
isosorbide,15 gallic acid16 and lignin.4,6,17 Lignin is the only
renewable aromatic polymer that is available in larger quantities
is increasingly looking toward using biomass-derived molecules at low cost that offers a potential source of ready-made
and materials with new functionality, enhanced performance aromatic polyols. Lignin is a complex aromatic polymer derived
attributes, and low carbon footprints. This has fueled interest in from p-coumaryl alcohol, coniferyl alcohol (G units), or sinapyl
replacing nonrenewable polymers, or components thereof, with alcohol (S units) building blocks, the proportions dictated by
sustainable alternatives. Significant advances have been made in the type of lignocellulosic species.18 The C9 phenylpropanoid
developing both biobased thermoplastics1,2 and thermosets.3 lignin units are linked together in the polymer by a variety of
For thermosets, aromatic intermediates are favored for their different linkages, the most abundant being the β-O-4 ether
structural rigidity as they impart desirable mechanical and linkage.18 A number of strategies have been used for
thermal properties to the material.4−6 The availability of incorporating lignin as a polyol precursor in epoxy resins.
petroleum-derived aromatic compounds could decline in the Lignin can be blended and reacted with epoxy prepolymers
future as petrochemical feedstocks shift away from traditional such as BADGE.6,19,20 Lignin can also be reacted with
crude oil to shale gas and lighter tight oils that contain less epichlorohydrin to produce an epoxy prepolymer that can
aromatics.7 then be cross-linked with curing agents.4,6 However, using
Epoxy resins represent a major class of thermosetting resins lignin directly can be challenging due to its poor solubility and
that demand a higher price than other polymers. Epoxy resins reactivity.21 Chemical modification of lignin (e.g., phenolation,
are versatile thermosets because they can be combined with a hydroxypropylation) has been effective in addressing these
wide range of curing agents. This enables them to be tailored to limitations.4,6,17
suit a diverse range of applications such as coatings, adhesives, In recent years there has been renewed interest in developing
composites, and electrical encapsulation.8 The most commonly epoxy resins from lignin.22−26 A new approach to producing
used epoxy prepolymer is bisphenol A diglycidyl ether lignin-derived polyols is to depolymerize the lignin to lower
(BADGE), which is commercially produced from bisphenol A molecular weight compounds. There have been several reports
and epichlorohydrin. While a commercial route exists for the on producing epoxy resins from lignin model compounds.
production of biobased epichlorohydrin,9 there is considerable Examples include vanillin and derivatives thereof,27−29 vanillyl
interest in replacing bisphenol A with biobased polyol alcohol derivatives,30 phenol,31 isoeugenol,32 syringaresinol,33
alternatives that are equally capable of delivering a high and compounds based on propyl guaiacol and its demethylated
performance resin. Alternatives for bisphenol A are also being product.34−36 Model compounds are useful in assessing the
sought on health and environmental grounds as it is a known potential of lignin in materials applications, however it is
endocrine disrupting compound that can potentially leach from important that more structurally relevant and readily available
BADGE-based epoxy resins used in food applications.10
A number of biobased alternatives to bisphenol A have been Received: June 1, 2017
investigated.3,9,11−13 Examples of biobased epoxy prepolymers Revised: July 4, 2017
include epoxidized cardanol,14 and glycidylation products of Published: July 6, 2017

© 2017 American Chemical Society 2640 DOI: 10.1021/acs.biomac.7b00767


Biomacromolecules 2017, 18, 2640−2648
Biomacromolecules Article

lignin streams be utilized in developing new polymers and cooled, neutralized with phosphate buffer (pH 7.0, 0.3 M, 10 mL) and
materials. Zhao and Abu-Omar (2017), for example, combined excess epichlorohydrin was removed. The product was extracted into
demethylated lignin with a glycidylated derivative of dichloromethane, and the solvent was removed before dissolving in
dihydroeugenol to produce lignin-containing epoxy resins.37 dioxane and treating with 20% aqueous NaOH. The final product was
an amber-colored viscous oil (LHEP, ca. 88% yield w/w). The diethyl
Processed lignins such as Kraft and organosolv lignins have ether-insoluble fraction of the lignin hydrogenolysis product (LHO)
been depolymerized and used to produce epoxy resins.23,24,38,39 was reacted in the same way to give the oligomeric lignin
The chemical structures of processed lignins are often heavily hydrogenolysis epoxy prepolymer (LHOEP, ca. 84% yield w/w).
modified compared to native lignins due to depolymerization Chemical Analysis. Analysis of the hydrogenolysis products by
(cleavage of C−O bonds) and repolymerization (formation of GCMS was performed as previously described.42 Gel permeation
C−C bonds) reactions that occur during extraction of the chromatography was performed on a PSS SDV Lux 1000 Å column set
lignin from biomass. Utilizing native lignin within biomass is an eluting with tetrahydrofuran using a Knauser Smartline GPC system
important emerging strategy in the biorefinery concept. In this with UV detection at 254 nm. A calibration curve was constructed
approach, lignin is fractionated first from the lignocellulosic using 4-n-propyl guaiacol, dihydro-diisoeugenol, and polystyrene
standards. 1H, 13C, 31P, and 2D HSQC NMR spectroscopy was
matrix by catalytic upstream processing to produce a solvent- performed on a Bruker AVIII 400 MHz spectrometer equipped with a
soluble depolymerized lignin oil and a holocellulose-enriched BBO 5 mm probe. Hydroxyl content was determined by quantitative
residue.18,40,41 31
P NMR spectroscopy as described previously.43 The EEW of the
We have previously reported that mild hydrogenolysis can epoxy resins was determined by potentiometric titration following
depolymerize lignin in situ in pine wood and an isolated native ASTM Standard D1652−04 scaled down to analyze 20 mg samples.
lignin into a mixture of phenolic monomers, dimers and Total chlorine content was determined by the Campbell Micro-
oligomers.42,43 Mild hydrogenolysis selectively cleaves the β-O- analytical Lab (Otago University, New Zealand).48 Hydrolyzable
4 and α-O-4 ether linkages and stabilizes the lignin fragments chlorine content was determined by titration following ASTM
against repolymerization through catalytic reduction. In recent Standard D1726−03 scaled down to analyze 25−50 mg samples.
The viscosity of the products, as 40% solutions in diethylene glycol
years, there has been increased interest in lignin hydrogenolysis
butyl ether, were measured at 25 °C using a Brookfield cone and plate
as a pathway to valuable low molecular weight chemicals from viscometer.
lignin.40,41,44−47 Most approaches use noble metal or Ni-based Preparation of Cured Epoxy Specimens. LHEP or LHOEP
catalysts and have focused on maximizing yields of monomeric were mixed with BADGE or GDGE in the desired proportions to give
products. In this regard, lignins rich in S units and with high β- a total mass of 0.6 g. Phenyl glycidyl ether (60 mg) was added as a
O-4 contents, such as those found in hardwoods, are preferred. reactive diluent to all mixtures, unless indicated otherwise, and the
Mild hydrogenolysis of softwood lignins, which are mainly components were mixed thoroughly with warming to about 40 °C.
composed of G units, give lower yields of monomers and The EEW of the individual epoxy resin components was used to
higher yields of oligomeric products.40,42 Research on utilizing calculate the equimolar amount of curing agent required to cure the
hydrogenolysis products of native lignins to produce new resin. The curing agent, either DETA or IPDA, was added and mixed
thoroughly before centrifuging at 4000 rpm for 1 min to remove air
biobased materials is yet to be reported. In this study we bubbles. The mixture was warmed to about 40 °C and transferred into
describe the development of novel biobased epoxy prepolymers an aluminum mold (40 × 5.7 × 0.9 mm), which was fixed between
from hydrogenolysis products derived from softwood lignin and glass plates. Resins containing DETA or IPDA were cured at 40 and 80
their use as replacements for BADGE in new epoxy °C, respectively for 24 h followed by a postcure at 120 and 150 °C,
thermosetting polymers. The novelty of this approach is that respectively for 24 h.
it takes advantage of the inherent properties of the native lignin Thermal Analysis and Flexural Testing. Glass transition
in the wood before it is chemically altered as in conventional temperature was determined by differential scanning calorimetry on
wood pulping processes. a Q1000 instrument (TA Instruments, U.S.A.) by subjecting cured
epoxy resins to a heat−cool−heat cycle from 0 to 200 °C at 10 °C/

■ EXPERIMENTAL SECTION
Materials. Ground, dry, extractive-free Pinus radiata wood was
min under a nitrogen atmosphere. Cure behavior was determined
using the same instrument by heating freshly prepared resins from 30
to 200 °C at 5 °C/min under a nitrogen atmosphere. Cured epoxy
prepared as described previously.42 The Pd/C (5% Pd) catalyst, resins were analyzed on a Q500 thermogravimetric analyzer (TA
epichlorohydrin, BADGE with an epoxy equivalent weight (EEW) of Instruments, U.S.A.) to determine mass loss as a function of
174 g/mol, glycerol diglycidyl ether (GDGE) with an EEW of 143 g/ temperature by heating to 500 °C at 10 °C/min under a nitrogen
mol, phenyl glycidyl ether, diethylenetriamine (DETA), and atmosphere. The statistic heat-resistant index temperature (Ts) was
isophorone diamine (IPDA) were purchased from Sigma-Aldrich. calculated using T5% and T30% (temperature at 5% and 30% weight
Hydrogenolysis. Hydrogenolysis reactions were carried out in a loss) according to the following equation:
450 mL limbo autoclave (Buchiglasuster, Switzerland) charged with Ts = 0.49[T5% + 0.6(T30% − T5%)] (1)
wood (12.9 g) and Pd/C (1.03 g, 8% w/w of wood) in dioxane/water
(1:1, 300 mL). The reactor was pressurized to 3.4 MPa with hydrogen Cured epoxy resins were conditioned at 23 °C and 50% relative
and heated to 195 °C for 24 h with stirring at 750 rpm.42 After cooling, humidity for at least 24 h. Three-point bending was performed using
the contents of the reactor were filtered, extracted into dichloro- an RSA-G2 dynamic mechanical thermal analyzer (TA Instruments,
methane and the extract was dried to constant weight to give the lignin U.S.A.). The support span length was 15 mm (span to thickness ratio:
hydrogenolysis oil product (LH, 2.9 g). This product was fractionated 16−17). The specimens were tested to failure, where possible. Average
into a diethyl ether-insoluble fraction (LHO, ca. 50% (w/w)) and a values and 95% confidence intervals were determined from at least
diethyl ether-soluble fraction. four replicates, except for LHOEP/BADGE 1:9 and the IPDA-cured
Preparation of Lignin Hydrogenolysis Epoxy Prepolymer specimens that were tested in duplicate due to limited amounts of
(LHEP). The lignin hydrogenolysis product (8.22 g) was dissolved in sample.


refluxing epichlorohydrin (13.7 mL). A 20% aqueous solution of
NaOH (1.40 g, 34.9 mmol) was added dropwise to the reaction RESULTS AND DISCUSSION
mixture over a 90 min period. During the reaction, water was removed
by azeotropic distillation with epichlorohydrin using a modified dean- Hydrogenolysis. Native lignin in Pinus radiata wood was
stark apparatus. The reaction was continued for a further 30 min, depolymerized by mild hydrogenolysis to a complex mixture of
2641 DOI: 10.1021/acs.biomac.7b00767
Biomacromolecules 2017, 18, 2640−2648
Biomacromolecules Article

Figure 1. Approach used to prepare cured epoxy resins from lignin hydrogenolysis products.

aromatic polyols composed of monomers, dimers, and


oligomers, as described previously.42 In this study, scaling-up
the reaction 6-fold in a 450 mL reactor gave the lignin
hydrogenolysis oil product (LH) in a yield of 78% w/w on
starting lignin. Consistent with previous findings, the main
monomers in the oil product were dihydroconiferyl alcohol
(DCA, 4-(3-hydroxypropyl)-2-methoxyphenol) and 4-propyl
guaiacol (PG, 4-propyl-2-methoxyphenol) constituting approx-
imately 22% and 3% w/w, respectively (Figure 1).42 Various
dimers and oligomers of DCA and PG represented
approximately 75% of the oil product. The oligomers ranged
in degree of polymerization from 3 to 7, as determined by gel
permeation chromatography (Figure 2). LH was fractionated

Figure 3. Quantitative 31P NMR spectrum of lignin hydrogenolysis oil


after reaction with the phosphitylating agent.

Figure 2. Molecular weight profiles of lignin hydrogenolysis products hydroxyl groups, respectively (Table 1). This compares with 1.1
and their epoxy derivatives. mmol/g for enzymatic mild acidolysis lignin isolated from P.
radiata wood42 and 2.3 mmol/g for a depolymerized lignin that
by solubility in diethyl ether. The diethyl ether-insoluble was used to produce epoxy resins.50 The hydrogenolysis
fraction (LHO), which was enriched in oligomers (Figure 2B), reaction cleaves β-O-4 ether linkages in the native wood lignin
made up 50% w/w of LH and was used along with LH to to liberate new phenolic hydroxyl groups. 31P NMR spectros-
prepare epoxy resins. The weight-average molecular weight copy also gives information on the chemical environment of the
(Mw) for LH and LHO was 667 and 898 g/mol, respectively. phenolic hydroxyl groups, namely, whether these groups
Both LH and LHO had a polydispersity index (PDI) of 1.5. present in C9 phenylpropanoid units are (i) bonded to H at
Quantitative 31P NMR spectroscopy is an important the 5-position (uncondensed) or (ii) bonded to another C9
analytical technique for quantifying the different types of unit at the 5-position (condensed), such as β-5, 5-5, or 4-O-5
hydroxyl groups in lignin (Figure 3).43,49 A high number of linked units (Figures 3 and 4, Table 1).43 The phenolic
phenolic hydroxyl groups is generally advantageous as they can hydroxyl groups in the monomers, DCA and PG (Figure 1),
be reacted with epichlorohydrin to generate epoxide function- analyze as uncondensed hydroxyl groups, whereas the dimers
ality, that is, glycidyl ethers (Figure 1). The LH and LHO and oligomers of DCA and PG could contain both
products contained 4.1 and 4.4 mmol/g of total phenolic uncondensed and condensed phenolic hydroxyl groups.
2642 DOI: 10.1021/acs.biomac.7b00767
Biomacromolecules 2017, 18, 2640−2648
Biomacromolecules Article

Table 1. Hydroxyl Group Content of the Lignin Hydrogenolysis Products and Their Epoxy Derivatives by Quantitative 31P
NMR Spectroscopy
phenolic OH (mmol/g)
sample β-5 4-O-5 5-5 uncondensed guaiacyl (G) p-hydroxy phenyl total acidic OH (mmol/g)
LH 0.53 (0.008)a 0.20 (0.003) 0.76 (0.012) 2.56 (0.019) 0.06 (0.007) 4.11 (0.06) 0.14 (0.005)
LHEP 0.07 0.01 0.02 0.03 0.001 0.13 0.003
% reacted 87 95 98 99 99 97 98
LHO 0.81 0.34 1.17 1.95 0.10 4.36 0.16
LHOEP 0.16 0.02 0.05 0.13 NDb 0.37 0.003
% reacted 80 93 96 93 100 92 98
a
Standard deviations in parentheses for LH analyzed in triplicate. bNot detected.

Figure 4. Examples of condensed and uncondensed structures in


lignin hydrogenolysis products.

Consequently, the oligomeric fraction of the hydrogenolysis oil


(LHO) contained a greater proportion of condensed phenolics
Figure 5. Partial HSQC NMR spectrum of LHEP showing glycidyl
(Table 1) compared to the whole oil (LH). In this study, ether functionality.
molecules with at least two phenolic hydroxyl groups (i.e.,
dimers and oligomers) were required to make cross-linking
epoxy resins. Both LHEP and LHOEP had higher molecular weights than
Glycidylation Reaction. The main product from the the lignin hydrogenolysis starting materials, LH and LHO
reaction of LH with epichlorohydrin was confirmed as the (Figure 2). Potential oligomerization reactions between newly
expected glycidyl ether by 1H, 13C and 2D HSQC NMR formed glycidyl groups and unreacted phenolic hydroxyl
spectroscopy. Signals in the HSQC spectrum of the lignin groups, which would lead to higher molecular weight resins,
hydrogenolysis epoxy prepolymer (LHEP), labeled as 1, 2, and were minimized by using an excess of epichlorohydrin in the
3, are characteristic of a glycidyl ether group connected to a reaction. Indeed, the Mw for LHEP and LHOEP increased by
lignin guaiacyl unit (Figure 5). The 1H and 13C chemical shifts only 60−70% to 1084 and 1514 g/mol (Table 2), respectively,
of the glycidyl signals were consistent with those reported in compared to the starting materials. When compared against
the literature.28 commercial BADGE resins, the average molecular weight of
Quantitative 31P NMR spectroscopy of LHEP and the LHEP was intermediate between typical liquid and solid
oligomeric epoxy prepolymer (LHOEP) showed that 97% and BADGE resins. The molecular weight of LHOEP was
92% of the phenolic hydroxyl groups were derivatized, equivalent to a solid BADGE resin with a degree of
respectively (Table 1). The condensed β-5 phenolic hydroxyl polymerization of 2.8 The PDI for the LHEP and LHOEP
group was less reactive compared to the uncondensed phenolic products was 1.8 (Table 2), which was similar to that of low
hydroxyl group (Table 1). This could be due to steric molecular weight BADGE resins.51 Higher PDI values have
hindrance of the reaction. Similarly, Fache and co-workers been reported for epoxy resins produced from depolymerized
(2016) found that the phenolic hydroxyl groups in syringyl- Kraft (PDI 3.5), organosolv (PDI 2.8), and hydrolysis (PDI
based (S-type) model compounds, with a methoxy group at the 4.6) lignins.39,50 Polydispersity is a measure of the molecular
5-position, gave lower conversions on glycidylation compared weight distribution. In this study, a low PDI was desirable as it
to guaiacyl-based (G-type) lignin models.27 Interestingly, the gave a more molecularly homogeneous prepolymer which was
other condensed phenolic hydroxyl groups (e.g., 5-5 and 4-O-5 reflected in the ability to process the resin.
linked) were not affected to the same degree as the phenolic An important criterion for epoxy resins is the epoxide
hydroxyl groups in β-5 structures (Table 1), suggesting factors content, which is commonly referred to as the epoxy equivalent
other than steric hindrance influence the reactivity. weight (EEW), expressed as the weight of resin to obtain one
2643 DOI: 10.1021/acs.biomac.7b00767
Biomacromolecules 2017, 18, 2640−2648
Biomacromolecules Article

Table 2. Chemical Properties of the Epoxy Prepolymers


resin Mn (g/mol) Mw (g/mol) PDI EEW (g/eq) hydrolyzable chlorine (%) total chlorine (%) viscositya (cP)
b
LHEP 610 1084 1.8 359 (17) 0.1 (0.04) 1.0 (0.03) 55
LHOEP 837 1514 1.8 452 (1) 0.3 (0.04) 1.7 (0.02) 132
a
As 40% solutions in diethylene glycol monobutyl ether. bStandard deviations are given in parentheses.

equivalent epoxide group.8 The lower the EEW the greater the hydrolyzable and total chlorine contents range from 0.02−0.1%
potential cross-linking density in the cured epoxy resin. The and 0.1−0.2%, respectively.8
EEW values for LHEP and LHOEP were 359 and 452 g/eq, Cured Epoxy Resins Containing BADGE. Blending and
respectively (Table 2). This translates to approximately one Curing. The viscosities of LHEP and LHOEP as 40% solutions
epoxide group per 1.5 lignin C9 units in the LHEP resin and in diethylene glycol butyl ether were 55 and 132 cP,
one epoxide group per 1.9 lignin C9 units in the LHOEP resin, respectively (Table 2). These viscosities were higher than for
for an estimated molecular weight of a glycidylated C9 lignin a typical liquid BADGE resin, but lower than for a solid
unit of 238 g/mol. The theoretical EEW values expected for BADGE resin with a degree of polymerization of 2.8 Adding 9%
LHEP and LHOEP were 300 and 307 g/eq, respectively, based of a reactive diluent was helpful in reducing the viscosity of the
on the quantitative 31P NMR results (Table 1). The difference epoxy resins so that they could be mixed with the curing agent
between the actual and theoretical values was possibly due to and remain workable at 40 °C. Reactive diluents are typically
side reactions that can occur during the glycidylation reaction.8 used to reduce the viscosity of epoxy resins without severely
Commercial BADGE resins can have a wide range of EEW impacting on their overall properties53 and are used at levels up
values, from 174 up to 6000 g/eq.8 BADGE used in this study to 20% w/w.8 Using this system gave workable resins, whereas
had an EEW of 174 g/eq. The EEW values for LHEP and other researchers have incorporated solvents or used elevated
LHOEP were lower than reported for epoxy resins made from temperatures (70−130 °C) in order to effectively process
depolymerized softwood Kraft lignin (768 g/eq) and hardwood lignin-based epoxy resins.24,27,28,33,39
organosolv lignin (537 g/eq),39 indicating more cross-linking is We first investigated DETA for curing the lignin-based epoxy
possible with lignin hydrogenolysis products from native prepolymers as it is used commercially as an aliphatic curing
softwood lignin. agent for curing epoxy resins at ambient temperature.8 Cured
Chlorine content is an important quality indicator for epoxy epoxy resins produced from LHEP and DETA resulted in
resins.8 High chlorine values can result in corrosivity issues in specimens that were brittle and easily broken when removed
cured resins, therefore low values are desirable. The total from the mold. The brittleness was attributed to the oligomer
chlorine content of LHEP and LHOEP was 1.0% and 1.7%, molecules in the resin that contained multiple glycidyl ether
respectively (Table 2). The hydrolyzable chlorine content of groups. These are analogous to multifunctional epoxy novolac
LHEP and LHOEP was 0.1% and 0.3%, respectively (Table 2). resins based on phenol formaldehyde novolacs that have high
Hydrolyzable chlorine content provides a measure of the cross-linking densities, which can result in increased brittle-
presence of chlorohydrin intermediates (Figure 6) from ness.8 It is common for commercial epoxy resins to be
formulated with multiple components to give optimum
processability and performance.8 Therefore, we investigated
the potential to formulate LHEP and LHOEP with BADGE.
DETA-cured epoxy resin specimens suitable for mechanical
testing could be prepared by blending LHEP with BADGE in
mass ratios of up to 3:1 and blending LHOEP with BADGE in
ratios of up to 1:3 (Figure 7). Preparation of specimens for

Figure 6. Main glycidylation reaction pathway of lignin hydrogenolysis


products showing chlorohydrin intermediate.

Figure 7. Cured epoxy resin specimens.


incomplete dehydrohalogenation that can reduce the potential
for the resin to cross-link with curing agents. Chlorine content
is rarely reported in publications on biobased epoxy resins. property testing using higher proportions of LHEP and
Zhao and Abu-Omar (2015) reported a hydrolyzable chlorine LHOEP was difficult due to processing limitations, except for
content of 0.5% and a total chlorine content of 3.4% for an a thermal analysis specimen of the 100% LHEP epoxy resin.
epoxy functionalized lignin model compound.34 A low level of Hence, LHEP was tested as a partial replacement for BADGE,
hydrolyzable chlorine was achieved in this study by applying a while LHOEP was tested as a potential performance enhancing
postreaction alkali treatment.52 For glycidylation reactions, additive.
optimized conditions are needed to achieve low levels of total The cured epoxy resins were all transparent, although those
chlorine. This is difficult to achieve at a small scale. Undesirable containing LHEP and LHOEP were amber in color (Figure 7).
side reactions can lead to bound chlorides which are not readily The color is unlikely to be a major limitation for most epoxy
saponified with NaOH solutions.8 In commercial resins, typical resin applications.
2644 DOI: 10.1021/acs.biomac.7b00767
Biomacromolecules 2017, 18, 2640−2648
Biomacromolecules Article

Thermal Properties. The glass transition temperature (Tg) Mechanical Properties. The measured mechanical proper-
values of the DETA-cured epoxy resins produced from LHEP/ ties of the LHEP/BADGE and LHOEP/BADGE blends cured
BADGE blends were lower than that of the BADGE control with DETA were superior to the BADGE control resin (Figure
resin, with a range of 68−80 °C (Table 3). Higher proportions 8). The cured LHEP/BADGE 2:1 blend resulted in flexural

Table 3. Thermal Analysis Data for Epoxy Resin Blends


Containing BADGE and Cured with DETA
resin Tg (°C) T5% (°C) Ts (°C)
a
BADGE 117 (0.7) 328 (0.8) 169 (0.2)
LHEP/BADGE 1:1 80 289 161
LHEP/BADGE 2:1 70 270 156
LHEP/BADGE 3:1 68 258 151
LHEP 53 236 144
LHOEP/BADGE 1:9 111 325 169
LHOEP/BADGE 1:3 104 311 165
a
95% confidence intervals are given in parentheses for the BADGE
resin based on four replicates.

of LHEP resulted in a lower Tg, which was consistent with the


100% LHEP resin having the lowest Tg of 53 °C. This trend
was the same for the cured LHOEP/BADGE resins, although
the differences relative to the BADGE resin were less
pronounced due to the lower proportions of LHOEP used in Figure 8. Flexural properties of epoxy resin blends containing BADGE
the blends. The Tg of the cured BADGE resin was similar to and cured with DETA (error bars are 95% confidence intervals of at
that reported previously (117 vs 119 °C).54 The lower Tg of the least four replicates).
cured blends could be due to several factors. BADGE is a rigid
molecule with the aromatic rings joined by a isopropylidene modulus and strength values that were 52% and 28% greater,
bridge (Figure 1), whereas LHEP and LHOEP contain dimers respectively, than BADGE alone. Increasing the proportion of
and oligomers with both rigid (e.g., 5-5) and flexible (e.g., β-5) LHEP in the blend from 67% to 75% resulted in a decline in
linkages (Figure 4). The flexible structures are likely to reduce both the flexural modulus and, to a greater degree, the flexural
the Tg. Methoxy groups that are associated with the lignin strength. The flexural strength values of replicate specimens of
structures are known to reduce the Tg of cured epoxy resins.30 the cured LHEP/BADGE 3:1 blend were highly varied
By combining a depolymerized lignin with an annulation (ranging from 80 to 159 MPa), which was attributed to
reaction prior to glycidylation, Kaiho and co-workers (2016) increased brittle behavior. The results indicated there was no
were able to produce epoxy resins with Tg values as high as 134 advantage in testing cured LHEP/BADGE blends at ratios
°C.55 lower than 1:1 (Figure 8). The cured LHOEP/BADGE 1:3
The cured LHEP/BADGE and LHOEP/BADGE resins were blend resulted in a flexural modulus and strength that was 26%
less thermally stable than the BADGE resin (Table 3). The and 38% greater, respectively, than BADGE alone. The increase
initial decomposition temperature (T5%) was lowest for in the flexural modulus was greater than for the LHEP/BADGE
specimens containing the highest proportion of LHEP and 1:1 blend and the flexural strength was the highest of all the
LHOEP, and was consistent with the 100% LHEP resin having specimens measured at 163 MPa. An increase in the flexural
the lowest T5% of 236 °C. The statistic heat-resistant index strength was observed even when the proportion of LHOEP to
temperature (Ts) is characteristic of the thermal stability of the BADGE was only 10% compared to the BADGE resin (140 vs
cured resins.35 The Ts values had the same trend as for the T5% 118 MPa).
values (Table 3). The reduction in thermal stability is likely due These results imply that the lignin-derived oligomers play a
to the presence of methoxy groups on the aromatic ring, as they key role in improving the flexural modulus and strength of the
are known to decrease thermal stability by means of electron cured epoxy resin blends. For the LHEP/BADGE blends,
donation to the aromatic ring.56 The thermal results suggest monomers in the lignin hydrogenolysis oil with single glycidyl
that cured epoxy resins containing LHEP and LHOEP may be groups potentially act more as reactive diluents, contributing
less suitable than BADGE-based resins in high temperature little to the mechanical properties of the cured resins. In this
applications. regard, choice of the lignin feedstock for hydrogenolysis is
The onset curing temperature (To) and the peak curing important, as softwood native lignins give higher yields of
temperature (Tp) for the BADGE resin was 67 and 91 °C, oligomeric products compared to hardwood lignins.40,42 Lignin
respectively. The LHEP/BADGE 1:1 and 2:1 blends had a monomers can be used to make a cross-linked epoxy polymer
reduced To of 46 and 43 °C and a reduced Tp of 80 and 78 °C, if, for example, they contain at least two phenolic hydroxyl
respectively. This curing behavior is attributed to the presence groups that can be glycidylated.34 The presence of methoxy
of primary aliphatic hydroxyl groups associated with the DCA groups in the lignin hydrogenolysis products is likely to
moieties (Figure 1) in LHEP. Hydroxyl groups are known to contribute to the superior mechanical properties of the cured
accelerate the rate of cure of epoxy resins with amine curing LHEP/BADGE and LHOEP/BADGE blends. Hernandez and
agents,8 and have been implicated in promoting the curing co-workers (2016) found that methoxy groups improved the
reaction of epoxy resins produced from lignin model modulus of cured epoxy resins derived from vanillyl alcohol,30
compounds and other depolymerized lignins.34,39 and hydrogen bonding between methoxy groups and hydroxyl
2645 DOI: 10.1021/acs.biomac.7b00767
Biomacromolecules 2017, 18, 2640−2648
Biomacromolecules Article

Table 4. Effect of Curing Agent (DETA, IPDA) on Thermal and Mechanical Properties of BADGE and Blended Cured Resins
DETA IPDA
resin Tg (°C) flexural modulus (GPa) flexural strength (MPa) Tg (°C) flexural modulus (GPa) flexural strength (MPa)
a
BADGE 117 (0.7) 3.1 (0.03) 118 (1) 162 (0.2) 2.9 (0.2) 146 (6)
LHEP/BADGE 1:1 80 3.7 129 100 3.8 141
LHOEP/BADGE 1:3 104 3.9 163 135 3.1 155
a
95% confidence intervals are given in parentheses for the BADGE resin based on at least three replicates.

groups formed during curing has been implicated in improving Table 5. Thermal and Mechanical Properties of Epoxy Resin
mechanical strength.4 Nonetheless, other factors, such as Blends Containing GDGE and Cured with DETA
monomer to oligomer ratio and cross-linking densities, can
Tg T5% Ts flexural modulus flexural strength
also influence mechanical properties. resin (°C) (°C) (°C) (GPa) (MPa)
Other researchers have prepared cured epoxy resins based on GDGE 58 260 139 2.8 (0.6)a 95 (8)
monomers that can be derived from lignin, such as vanillin,27,28 LHEP/ 59 242 139 3.7 (0.1) 121 (2)
vanillyl alcohol,30 isoeugenol,32 and propyl guaiacol,34,36 GDGE 1:1
however, no comparable information is provided on mechanical LHEP/ 60 220 136 4.0 (0.2) 140 (7)
properties. In related work, Ferdosian and co-workers (2016) GDGE 2:1
measured mechanical properties of glass fiber-reinforced a
95% confidence intervals are given in parentheses.
plastics prepared using BADGE blended with epoxy resins
made from depolymerized softwood Kraft and hardwood
organosolv lignins.39 They also observed improvements in LHEP/GDGE epoxy resin blends would be restricted to lower
flexural modulus and strength, but mainly at a low blend ratio temperature applications.
of 1:3 lignin-based epoxy to BADGE. In our work, we were able The flexural modulus and strength of the cured epoxy resins
to increase both the flexural modulus and strength in higher were improved with increasing proportions of LHEP in the
blend ratios up to 2:1 lignin-based epoxy to BADGE. LHEP/GDGE blends (Table 5). The cured LHEP/GDGE 2:1
Alternative Curing Agent. Selected epoxy resin blends were blend resulted in flexural modulus and strength values that were
cured with IPDA, a cycloaliphatic curing agent commonly used 43% and 47% greater, respectively, than the GDGE control
in industry, to investigate the effect of a different curing agent. resin, and were intermediate between the modulus and strength
These cured epoxy resins had higher Tg values compared with values obtained for the cured LHEP/BADGE 1:1 and 2:1
blends. The LHEP/GDGE 2:1 blend produced a cured resin
those cured with DETA (Table 4). Cycloaliphatic curing agents
with a 29% greater flexural modulus and 19% greater flexural
are known to enhance the thermal properties of cured resins
strength than the BADGE control resin. This indicates that it is
compared to aliphatic amine curing agents.8 As with DETA,
possible to make 100% biobased epoxy resins that, when cured,
blending LHEP and LHOEP with BADGE resulted in cured
have superior mechanical properties compared to BADGE
epoxy resins with lower Tg values relative to the BADGE resins.
control resin. The flexural modulus of the LHEP/BADGE 1:1 Summary of Mechanical Performance. In summary, the
blend resin was similar when cured with either curing agent flexural properties were very promising, with the cured LHEP/
(Table 4). This was also true for the BADGE control resin. The BADGE and LHOEP/BADGE blends performing better than
LHEP/BADGE 1:1 blend that was cured with IPDA no longer the industry standard BADGE resin. LHEP has potential as a
had a flexural strength performance advantage over the BADGE biobased partial replacement for BADGE. Alternatively, LHEP
resin, compared to curing with DETA. Curing BADGE with used in combination with GDGE warrants further investigation
IPDA has previously been reported to increase flexural strength as a complete replacement for BADGE. LHOEP appears to
compared to curing with DETA.54 The flexural properties of have potential as a biobased performance enhancing additive
the IPDA-cured LHOEP blend were similar to the IPDA-cured for BADGE-based resins. More extensive formulation and
BADGE resin and reduced compared to the DETA-cured property testing is needed to fully assess the potential of these
LHOEP blend (Table 4). Overall, these results indicated that lignin hydrogenolysis products in developing novel biobased
blending LHEP and LHOEP with BADGE, and curing with epoxy resins.
IPDA, gave mechanical properties that were similar to the
BADGE control resin.
Cured Epoxy Resins Containing Glycerol Diglycidyl
■ CONCLUSIONS
Mild hydrogenolysis of native softwood lignin produces a
Ether. LHEP was blended with GDGE to produce epoxy resins polyol product that can be used to prepare novel biobased
containing entirely biobased components. GDGE has been epoxy resins. Reacting this lignin-based polyol with epichlor-
used as an epoxy prepolymer in lignin-containing epoxy ohydrin gave an epoxy prepolymer that, when blended with
resins.22,57 The addition of the reactive diluent was not BADGE or GDGE, resulted in cured epoxy resins with a greater
required for the LHEP/GDGE blends as the viscosity was flexural modulus and strength to those made from BADGE
sufficiently low to cast specimens at 40 °C (e.g., Figure 7). The alone. This is the first reported example of a deconstructed
Tg values of the blends that were cured with DETA were similar native softwood lignin being used to produce concept epoxy
to the GDGE control resin (Table 5) and approximately 60 °C resins with superior mechanical properties to the industry
lower than the BADGE control resin. The cured GDGE resin standard. Further work is needed to improve and scale-up the
had reduced thermal stability compared to the BADGE resin lignin hydrogenolysis process, and to optimize the glycidylation
and addition of LHEP decreased this further (Table 5), as was and resin formulation chemistry. The promising performance of
seen with the LHEP/BADGE blends. These results suggest that these concept epoxy resins suggests that with this further work,
2646 DOI: 10.1021/acs.biomac.7b00767
Biomacromolecules 2017, 18, 2640−2648
Biomacromolecules Article

a viable and fully biobased epoxy resin with properties (16) Aouf, C.; Nouailhas, H.; Fache, M.; Caillol, S.; Boutevin, B.;
comparable to the bisphenol A-derived industry standard may Fulcrand, H. Multi-functionalization of gallic acid. Synthesis of a novel
be realized. bio-based epoxy resin. Eur. Polym. J. 2013, 49, 1185−1195.


(17) Engelmann, G.; Ganster, J. Lignin Reinforcement in Thermosets
Composites. In Lignin in Polymer Composites; Faruk, O., Sain, M., Eds.;
AUTHOR INFORMATION William Andrew, 2016; pp 119−151.
Corresponding Author (18) Rinaldi, R.; Jastrzebski, R.; Clough, M. T.; Ralph, J.; Kennema,
M.; Bruijnincx, P. C.; Weckhuysen, B. M. Paving the Way for Lignin
*E-mail: daniel.vandepas@scionresearch.com. Phone: +64-7- Valorisation: Recent Advances in Bioengineering, Biorefining and
343-5899. Fax: +64-7-348-0952. Catalysis. Angew. Chem., Int. Ed. 2016, 55, 8164−8215.
ORCID (19) Doherty, W. O. S.; Mousavioun, P.; Fellows, C. M. Value-adding
to cellulosic ethanol: Lignin polymers. Ind. Crops Prod. 2011, 33, 259−
Daniel J. van de Pas: 0000-0002-6415-9569 276.
Notes (20) Stewart, D. Lignin as a base material for materials applications:
The authors declare no competing financial interest. Chemistry, application and economics. Ind. Crops Prod. 2008, 27,


202−207.
(21) Ma, S.; Li, T.; Liu, X.; Zhu, J. Research progress on bio-based
ACKNOWLEDGMENTS thermosetting resins. Polym. Int. 2016, 65, 164−173.
This research was supported by the New Zealand Ministry of (22) Engelmann, G.; Ganster, J. Bio-based epoxy resins with low
Business Innovation and Employment via Scion Core funding. molecular weight kraft lignin and pyrogallol. Holzforschung 2014, 68,
The authors wish to acknowledge Ibrar Hussain for technical 1−12.
(23) Xin, J.; Li, M.; Li, R.; Wolcott, M. P.; Zhang, J. Green Epoxy
assistance with lignin hydrogenolysis.


Resin System Based on Lignin and Tung Oil and Its Application in
Epoxy Asphalt. ACS Sustainable Chem. Eng. 2016, 4, 2754−2761.
REFERENCES (24) Ferdosian, F.; Yuan, Z.; Anderson, M.; Xu, C. Sustainable lignin-
(1) Chen, G. Q.; Patel, M. K. Plastics derived from biological sources: based epoxy resins cured with aromatic and aliphatic amine curing
present and future: a technical and environmental review. Chem. Rev. agents: Curing kinetics and thermal properties. Thermochim. Acta
2012, 112, 2082−2099. 2015, 618, 48−55.
(2) Wang, C.; Kelley, S. S.; Venditti, R. A. Lignin-Based (25) Asada, C.; Basnet, S.; Otsuka, M.; Sasaki, C.; Nakamura, Y.
Thermoplastic Materials. ChemSusChem 2016, 9, 770−783. Epoxy resin synthesis using low molecular weight lignin separated
(3) Raquez, J. M.; Deléglise, M.; Lacrampe, M. F.; Krawczak, P. from various lignocellulosic materials. Int. J. Biol. Macromol. 2015, 74,
Thermosetting (bio)materials derived from renewable resources: A 413−419.
critical review. Prog. Polym. Sci. 2010, 35, 487−509. (26) Over, L. C.; Grau, E.; Grelier, S.; Meier, M. A. R.; Cramail, H.
(4) Koike, T. Progress in development of epoxy resin systems based Synthesis and Characterization of Epoxy Thermosetting Polymers
on wood biomass in Japan. Polym. Eng. Sci. 2012, 52, 701−717. from Glycidylated Organosolv Lignin and Bisphenol A. Macromol.
(5) Llevot, A.; Grau, E.; Carlotti, S.; Grelier, S.; Cramail, H. From Chem. Phys. 2017, 218, 1600411.
Lignin-derived Aromatic Compounds to Novel Biobased Polymers. (27) Fache, M.; Boutevin, B.; Caillol, S. Epoxy thermosets from
Macromol. Rapid Commun. 2016, 37, 9−28. model mixtures of the lignin-to-vanillin process. Green Chem. 2016, 18,
(6) Upton, B. M.; Kasko, A. M. Strategies for the Conversion of 712−725.
Lignin to High-Value Polymeric Materials: Review and Perspective. (28) Fache, M.; Viola, A.; Auvergne, R.; Boutevin, B.; Caillol, S.
Chem. Rev. 2016, 116, 2275−2306. Biobased epoxy thermosets from vanillin-derived oligomers. Eur.
(7) Bruijnincx, P. C.; Weckhuysen, B. M. Shale gas revolution: an Polym. J. 2015, 68, 526−535.
opportunity for the production of biobased chemicals? Angew. Chem., (29) Wang, S.; Ma, S.; Xu, C.; Liu, Y.; Dai, J.; Wang, Z.; Liu, X.;
Int. Ed. 2013, 52, 11980−11987. Chen, J.; Shen, X.; Wei, J.; Zhu, J. Vanillin-Derived High-Performance
(8) Pham, H. Q.; Marks, M. J. Epoxy Resins. Ullmann’s Encyclopedia Flame Retardant Epoxy Resins: Facile Synthesis and Properties.
of Industrial Chemistry; Wiley-VCH Verlag GmbH & Co. KGaA: Macromolecules 2017, 50, 1892−1901.
Weinheim, 2000. (30) Hernandez, E. D.; Bassett, A. W.; Sadler, J. M.; La Scala, J. J.;
(9) Auvergne, R.; Caillol, S.; David, G.; Boutevin, B.; Pascault, J. P. Stanzione, J. F. Synthesis and Characterization of Bio-based Epoxy
Biobased thermosetting epoxy: present and future. Chem. Rev. 2014, Resins Derived from Vanillyl Alcohol. ACS Sustainable Chem. Eng.
114, 1082−1115. 2016, 4, 4328−4339.
(10) Vandenberg, L. N.; Maffini, M. V.; Sonnenschein, C.; Rubin, B. (31) Maiorana, A.; Spinella, S.; Gross, R. A. Bio-based alternative to
S.; Soto, A. M. Bisphenol-A and the great divide: a review of the diglycidyl ether of bisphenol A with controlled materials
controversies in the field of endocrine disruption. Endocr. Rev. 2009, properties. Biomacromolecules 2015, 16, 1021−1031.
30, 75−95. (32) François, C.; Pourchet, S.; Boni, G.; Fontaine, S.; Gaillard, Y.;
(11) Gandini, A. Epoxy Polymers Based on Renewable Resources. In Placet, V.; Galkin, M. V.; Orebom, A.; Samec, J.; Plasseraud, L.
Epoxy Polymers; Pascault, J.-P., Williams, R. J. J., Eds.; Wiley-VCH Diglycidylether of iso-eugenol: a suitable lignin-derived synthon for
Verlag GmbH & Co. KGaA: Weinheim, 2010; pp 55−78. epoxy thermoset applications. RSC Adv. 2016, 6, 68732−68738.
(12) Baroncini, E. A.; Kumar Yadav, S.; Palmese, G. R.; Stanzione, J. (33) Janvier, M.; Hollande, L.; Jaufurally, A. S.; Pernes, M.; Menard,
F. Recent advances in bio-based epoxy resins and bio-based epoxy R.; Grimaldi, M.; Beaugrand, J.; Balaguer, P.; Ducrot, P. H.; Allais, F.
curing agents. J. Appl. Polym. Sci. 2016, 133, 44103. Syringaresinol: A Renewable and Safer Alternative to Bisphenol A for
(13) Ng, F.; Couture, G.; Philippe, C.; Boutevin, B.; Caillol, S. Bio- Epoxy-Amine Resins. ChemSusChem 2017, 10, 738−746.
Based Aromatic Epoxy Monomers for Thermoset Materials. Molecules (34) Zhao, S.; Abu-Omar, M. M. Biobased Epoxy Nanocomposites
2017, 22, 149. Derived from Lignin-Based Monomers. Biomacromolecules 2015, 16,
(14) Voirin, C.; Caillol, S.; Sadavarte, N. V.; Tawade, B. V.; Boutevin, 2025−2031.
B.; Wadgaonkar, P. P. Functionalization of cardanol: towards biobased (35) Zhao, S.; Abu-Omar, M. M. Renewable Thermoplastics Based
polymers and additives. Polym. Chem. 2014, 5, 3142−3162. on Lignin-Derived Polyphenols. Macromolecules 2017, 50, 3573−3581.
(15) Chrysanthos, M.; Galy, J.; Pascault, J.-P. Preparation and (36) Zhao, S.; Abu-Omar, M. M. Renewable Epoxy Networks
properties of bio-based epoxy networks derived from isosorbide Derived from Lignin-Based Monomers: Effect of Cross-Linking
diglycidyl ether. Polymer 2011, 52, 3611−3620. Density. ACS Sustainable Chem. Eng. 2016, 4, 6082−6089.

2647 DOI: 10.1021/acs.biomac.7b00767


Biomacromolecules 2017, 18, 2640−2648
Biomacromolecules Article

(37) Zhao, S.; Abu-Omar, M. M. Synthesis of Renewable Thermoset linkage and the phenylnaphthalene structure coupled with selective
Polymers through Successive Lignin Modification Using Lignin- β-O-4 bond cleavage for synthesizing lignin-based epoxy resins with a
Derived Phenols. ACS Sustainable Chem. Eng. 2017, 5, 5059−5066. controlled glass transition temperature. Green Chem. 2016, 18, 6526−
(38) Ferdosian, F.; Yuan, Z.; Anderson, M.; Xu, C. Chemically 6535.
modified lignin through epoxidation and its thermal properties. J. Sci. (56) Harvey, B. G.; Guenthner, A. J.; Lai, W. W.; Meylemans, H. A.;
Technol. For. Prod. Processes 2012, 2, 11−15. Davis, M. C.; Cambrea, L. R.; Reams, J. T.; Lamison, K. R. Effects of o-
(39) Ferdosian, F.; Zhang, Y.; Yuan, Z.; Anderson, M.; Xu, C. Curing Methoxy Groups on the Properties and Thermal Stability of
kinetics and mechanical properties of bio-based epoxy composites Renewable High-Temperature Cyanate Ester Resins. Macromolecules
comprising lignin-based epoxy resins. Eur. Polym. J. 2016, 82, 153− 2015, 48, 3173−3179.
165. (57) Ismail, T. N. M. T.; Hassan, H. A.; Hirose, S.; Taguchi, Y.;
(40) Van den Bosch, S.; Schutyser, W.; Vanholme, R.; Driessen, T.; Hatakeyama, T.; Hatakeyama, H. Synthesis and thermal properties of
Koelewijn, S. F.; Renders, T.; De Meester, B.; Huijgen, W. J. J.; ester-type crosslinked epoxy resins derived from lignosulfonate and
Dehaen, W.; Courtin, C. M.; Lagrain, B.; Boerjan, W.; Sels, B. F. glycerol. Polym. Int. 2009, 59, 181−186.
Reductive lignocellulose fractionation into soluble lignin-derived
phenolic monomers and dimers and processable carbohydrate pulps.
Energy Environ. Sci. 2015, 8, 1748−1763.
(41) Parsell, T.; Yohe, S.; Degenstein, J.; Jarrell, T.; Klein, I.; Gencer,
E.; Hewetson, B.; Hurt, M.; Kim, J. I.; Choudhari, H.; Saha, B.; Meilan,
R.; Mosier, N.; Ribeiro, F.; Delgass, W. N.; Chapple, C.; Kenttamaa, H.
I.; Agrawal, R.; Abu-Omar, M. M. A synergistic biorefinery based on
catalytic conversion of lignin prior to cellulose starting from
lignocellulosic biomass. Green Chem. 2015, 17, 1492−1499.
(42) Torr, K. M.; van de Pas, D. J.; Cazeils, E.; Suckling, I. D. Mild
hydrogenolysis of in-situ and isolated Pinus radiata lignins. Bioresour.
Technol. 2011, 102, 7608−7611.
(43) van de Pas, D. J.; Nanayakkara, B.; Suckling, I. D.; Torr, K. M.
Comparison of hydrogenolysis with thioacidolysis for lignin structural
analysis. Holzforschung 2014, 68, 151−155.
(44) Song, Q.; Wang, F.; Cai, J.; Wang, Y.; Zhang, J.; Yu, W.; Xu, J.
Lignin depolymerization (LDP) in alcohol over nickel-based catalysts
via a fragmentation−hydrogenolysis process. Energy Environ. Sci. 2013,
6, 994−1007.
(45) Galkin, M. V.; Samec, J. S. Selective route to 2-propenyl aryls
directly from wood by a tandem organosolv and palladium-catalysed
transfer hydrogenolysis. ChemSusChem 2014, 7, 2154−2158.
(46) Schutyser, W.; Van den Bosch, S.; Renders, T.; De Boe, T.;
Koelewijn, S. F.; Dewaele, A.; Ennaert, T.; Verkinderen, O.; Goderis,
B.; Courtin, C. M.; Sels, B. F. Influence of bio-based solvents on the
catalytic reductive fractionation of birch wood. Green Chem. 2015, 17,
5035−5045.
(47) Shuai, L.; Amiri, M. T.; Questell-Santiago, Y. M.; Héroguel, F.;
Li, Y.; Kim, H.; Meilan, R.; Chapple, C.; Ralph, J.; Luterbacher, J. S.
Formaldehyde stabilization facilitates lignin monomer production
during biomass depolymerization. Science 2016, 354, 329−333.
(48) Celon, E.; Bresadola, S. On the use of sodium borohydride in
the microdetermination of chlorine or bromine in highly halogenated
organic compounds. Microchim. Acta 1969, 57, 441−448.
(49) Granata, A.; Argyropoulos, D. S. 2-chloro-4,4,5,5-tetramethyl-
1,3,2-dioxaphospholane, a reagent for the accurate determination of
the uncondensed and condensed phenolic moieties in lignins. J. Agric.
Food Chem. 1995, 43, 1538−1544.
(50) Ferdosian, F.; Yuan, Z.; Anderson, M.; Xu, C. Synthesis and
characterization of hydrolysis lignin-based epoxy resins. Ind. Crops
Prod. 2016, 91, 295−301.
(51) Mc Aninch, I. M.; Palmese, G. R.; Lenhart, J. L.; La Scala, J. J.
Epoxy-amine networks with varying epoxy polydispersity. J. Appl.
Polym. Sci. 2015, 132, 41503.
(52) Bauer, R. S.; De La Mere, H. E.; Klarquist, J. M.; Newman, S. F.;
McKetta, J. J. Epoxy resins and epoxides. Encyclopedia of Chemical
Processing and Design; Marcel Dekker: New York, 1983; Vol. 19, pp
260−297.
(53) May, C. A. Epoxy Resins: Chemistry and Technology; Marcel
Dekker, Inc.: New York, 1988.
(54) Garcia, F. G.; Soares, B. G.; Pita, V. J. R. R.; Sánchez, R.;
Rieumont, J. Mechanical properties of epoxy networks based on
DGEBA and aliphatic amines. J. Appl. Polym. Sci. 2007, 106, 2047−
2055.
(55) Kaiho, A.; Mazzarella, D.; Satake, M.; Kogo, M.; Sakai, R.;
Watanabe, T. Construction of the di(trimethylolpropane) cross

2648 DOI: 10.1021/acs.biomac.7b00767


Biomacromolecules 2017, 18, 2640−2648

Anda mungkin juga menyukai