You are on page 1of 12

Research Article

pubs.acs.org/journal/ascecg

Ultrastiff Biobased Epoxy Resin with High Tg and Low Permittivity:


From Synthesis to Properties
Jintao Wan,† Jianqing Zhao,‡ Bin Gan,† Cheng Li,§ Jon Molina-Aldareguia,† Ying Zhao,‡ Ye-Tang Pan,†
and De-Yi Wang*,†

IMDEA Materials Institute, C/Eric Kandel, 2, Getafe, Madrid 28906, Spain

School of Materials Science and Engineering, South China University of Technology, 510640 Guangzhou, China
§
State Key Laboratory of Chemical Engineering, College of Chemical and Biological Engineering, Zhejiang University, Zheda Road
38, Hangzhou 310027, China
*
S Supporting Information

ABSTRACT: Harvesting biobased epoxy resins with im-


proved thermomechanical properties (e.g., glass transition
temperature Tg and storage modulus), mechanical and
dielectric similar and even superior to that of bisphenol A
epoxy resin (DGEBA) is vital to many applications, yet
remains a substantial challenge. Here we develop a novel
eugenol-based epoxy monomer (TEU-EP) with a branched
topology and a very rich biobased retention (80 wt %). TEU-
EP can be well cured by 3,3′-diaminodiphenyl sulfone
(33DDS) and the resultant TEU-EP/33DDS system can be
considered as a “single” epoxy component, exhibiting adequate
reactivity at high processing temperatures. Importantly,
compared with DGEBA/33DDS, TEU-EP/33DDS achieves
a 33 °C, 39% and 55% increment in the glass transition temperature, Young’s modulus, and hardness, respectively, and shows the
improved creep resistance and dimensional stability. TEU-EP/33DDS is also characterized by the considerably reduced
permittivity, dielectric loss factor, and flammability with high yield of pyrolytic residual. Overall, TEU-EP endows the cured
epoxy with a number of the distinguished properties outperforming its DGEBA counterpart, and therefore may find practical
applications in demanding and even cutting-edge areas.
KEYWORDS: Biobased epoxy monomer, Synthesis, Curing, Mechanical and thermal properties, Dielectric properties

■ INTRODUCTION
With accelerating consumption of fossil-based unrenewable
materials has been banned in EU, USA, and Canada, and as a
result polyprolene is now widely used to fabricate the milk bottle
resources, transformation of renewable feedstocks, especially to replace bisphenol A-type polycarbonate (PC).
biomasses, to sustainable fuel and polymeric materials becomes Epoxy resins derived from the renewable biomasses3,4 become
increasingly important. Thermosetting polymeric materials are the most promising alternative for petroleum-based counter-
generally unable to rework or reshape after use, which will not parts. Among them, epoxidized soybean oils5,6 are massively
only hasten the consumption of unrenewable resources but also produced and used, but in practical applications they cannot
cause the accumulative disposition of thermosetting wastes in the serve as an epoxy matrix alone, because of their inferior
environment. Among the thermosetting polymers, epoxy resins properties (e.g., low glass temperature, modulus and thermal
find a myriad of applications as in coatings, adhesives, properties) compared with standard DGEBAs, arising from their
composites, construction, electrical engineering, and electronics long and highly flexible aliphatic skeleton. Instead, they are
owing to their attractive comprehensive properties and wide usually applied as plasticizers or processing agents for synthetic
applicability.1,2 Today approximately 90% of epoxy resins, resins, especially, polyvinyl chloride (PVC). To improve the
DGEBAs, are produced from petroleum-based bisphenol A properties of biobased epoxy resins, rosin,7−9 gallic acid,10−12
and epichlorohydrin worldwide.3 Bisphenol A is synthesized vanillic acid,11 vanillin,13,14 itaconic acid,15 dimer fatty acid,16
from unrenewable petroleum-based phenol and acetone, and sucrose fatty acid esters,17 isosorbide,18−20 cinnamic acid and
epichlorohydrin is mainly synthesized from oxidation of allyl dipentene,21 cardanol,22,23 sugar,24 furan alcohol,25 and others26
chloride though a considerable portion comes from renewable
glycerol. Also, epichlorohydrin is carcinogenic, and bisphenol A Received: March 7, 2016
can cause estrogen-like long-lasting effects on living organisms. Revised: April 6, 2016
As a result, the presence of bisphenol A in food packaging Published: April 21, 2016

© 2016 American Chemical Society 2869 DOI: 10.1021/acssuschemeng.6b00479


ACS Sustainable Chem. Eng. 2016, 4, 2869−2880
ACS Sustainable Chemistry & Engineering Research Article

Scheme 1. Molecular Structures of DGEBA, Synthesized Eugenol-Based Epoxy Resin (TEU-EP), and Curing Agent (33DDS)a

a
Theoretically, DGEBA and TEU-EP have the epoxy functionalities of two and three, respectively, and 33DDS has the amino hydrogen
functionalities of four.

have been used to synthesize biobased epoxy monomers. phase-transfer catalytic system, followed by epoxidation of the
Moreover, recent advances in biobased epoxy foams27 and inherited allyl double bonds. In the TEU-EP molecule, a low-
biobased biodegradable28,29 and flame retardant30,31 epoxy resins molecular-weight triazine concentrator links three eugenol
as well as biobased epoxy curing agents32−36 further expand the moieties together, which leads to the further increased biobased
scope of biobased epoxy chemistry and technology, which will retention as high as 80 wt %. Also thanks to this triazine
greatly extend potential applications of biobased epoxy systems. concentrator, TEU-EP has derived a branched molecular
More interestingly, eugenol, 4-allyl-2-methoxyphenol, ex- topology, which causes the increased epoxy functionalities but
tracted from plants (e.g., clove oil), is one kind of essential oils decreased melting temperature compared with our reported
with relatively low cost, consisting of the rigid aromatic ring eugenol-based epoxy resins.45,46 Moreover, this highly rigid and
substituted by an allyl group, a hydroxyl group, and a methoxyl low polarizable triazine moiety is expected to greatly contribute
group.37−39 Eugenol-based thermosets have been attracting to mechanical, thermal and electrical properties of the resulting
increasing research interests. Harvey et al.37 reported the eugenol epoxy system. Here in the presence of 3,3′-diaminodiphenylsul-
based cyanate ester that could be well cured at a high temperate, fone (33DDS) as the curing agent, the influence of TEU-EP’s
and the cured epoxy exhibited excellent thermal stability up to molecular structure on the curing behaviors, thermomechanical
350 °C with a Tg of 186 °C. Neda et al.38 reported an eugenol- and mechanical properties, dielectric characteristics, thermal
based bismaleimide with high Tg (350 °C) and modulus (2.6− decomposition, and flammability are systematically investigated,
3.1 GPa). Eugenol-based benzoxazines40−42 and eugenol-based by comparing with a standard DGEBA/33DDS system. It will be
unsaturated resins43 were reported with improved specific showed that cured TEU-EP/33DDS is highlighted by the
properties. ultrahigh Tg, Young’s modulus and hardness, outstanding
It seems more interesting to develop eugenol-based epoxy thermomechanical properties, low permittivity and dielectric
resins for their easy processing, highly tailorable and versatile loss, and reduced flammability.
properties, and thus wide potential applications. Zhang et al.44
reported an eugenol-based epoxy resin (Eu-Ep) with biobased
content of 62.7 wt % via epoxidation of eugenol’s double bond,
■ EXPERIMENTAL SECTION
Materials. Eugenol, cyanuric chloride, benzyltriethylammonium
followed by O-glycidylation of its hydroxyl group. An acid chloride, chloroperoxybenzoic acid (mCPBA, ∼ 77%), sodium
anhydride-cured Eu-Ep showed a glass temperature of 114 °C. hydroxide, chloroform, dichloromethane, acetone, menthol, sodium
On the other hand, according to our very recently developed hydroxide, and tetraethylammonium bromide were purchased from
strategy,45,46 two eugenol moieties were bridged by the aliphatic Sigma-Aldrich. 3,3′-Diaminodiphenylsulfone (33DDS, > 98%) was
ether or aromatic ester bonds, to yield to bifunctional eugenol- brought from TCI Europe. Bisphenol A epoxy resin (DGEBA) under
trademark of Epoxy Resin C (epoxy value = 0.531 mol epoxide/100 g)
based epoxy resins, DEU-EP and TPEU-EP. They had the was obtained from Faserverbundwerkstoffe Composite Technology,
increased net biobased content as high as 70.2 wt %, and the Germany. Other materials were used as received. The eugenol-based
cured epoxy showed the increased moduli and reduced epoxy resin (TEU-EP) was originally synthesized in our lab, whose
flammability. In particular, 3,3′-diaminodiphenyl sulfone epoxy value was titrated to be 0.451 mol epoxide/100 g according to
(33DDS) cured TPEU-EP was self-extinguishing in a vertical ASTM D1652 (titrant HClO4 in acetic acid; reagent tetraethylammo-
burning test,46 expressing rather low flammability. However, the nium bromide; indicator crystal violet; solvent CH2Cl2). The molecular
melting temperature of TPEU-EP was rather high (>150 °C), structures of DGEBA, TEU-EP, and 33DDS are shown in Scheme 1.
and the glass temperature of the cured epoxy product was still Molecular Structure Characterization. The FTIR spectra (KBr
lower than that of DGEBA system. pellet) of the synthesized epoxy resin and intermediate were registered
on a Nicolet iS50 spectrometer over the wavenumber range of 400−
Up to now, an open issue in the field of biobased epoxy resins 4000 cm−1. The room-temperature 1H NMR (400.1 MHz) and 13C
is to simultaneously enhance thermal resistance (especially glass NMR (101 MHz) spectra were recorded on a Varian Mercury AS400
transition temperature), mechanical properties, and flame spectrometer (5 mm Varian 4NUC probe). The solvent and internal
retardancy of biobased (including eugenol-based) epoxy resins standard were CDCl3 and tetramethylsilane (TMS), respectively. The
outperforming DGEBAs in as many properties as possible. mass spectrum of TEU-EP was recorded on a mass spectrometer
Additionally, the retained biobased contents of the resulting (Model LCD Deca xp max, Thermo Electron Corporation, USA) with
epoxy monomers still need further improving to more efficiently an ESI source. The sample was dissolved in chloroform and then diluted
use biobased precursors. Moreover, biobased epoxy resins with with methanol before measurement.
improved dielectric properties are still not yet achieved, though Curing Behavior Characterization. A differential scanning
calorimeter (Q200, TA Instruments, USA) was used to study the
epoxy-based materials with good dielectric properties are widely epoxy curing reaction. For the simplicity, we fixed the ratio of TEU-EP
used in the microelectronic industry. To meet these challenges, or DGEBA to 33DDS according to the stoichiometry (epoxy ring: N−H
here we develop an updated, facile strategy to realize a novel = 1:1 by mole). The mixture was dissolved in acetone with stirring, and
trifunctional eugenol-based epoxy monomer, TEU-EP (Scheme then, the solvent was removed at 50 °C in vacuum. The as-prepared
1), by reacting eugenol with cyanuric chloride in a high-efficient reaction mixture (10 ± 2 mg) was enclosed in an aluminum crucible and

2870 DOI: 10.1021/acssuschemeng.6b00479


ACS Sustainable Chem. Eng. 2016, 4, 2869−2880
ACS Sustainable Chemistry & Engineering Research Article

Scheme 2. Synthesis of the Triazine-Bridged Eugenol-Based Epoxy Resin (TEU-EP)a

a
Reaction conditions: (i) eugenol/NaOH/H2O + cyanuric chloride/benzyltriethylammonium chloride/CHCl3, from 10 °C to reflux, 8 h; (ii)
chloroperoxybenzoic acid/CH2Cl2, from 0 °C to room temperature, 5 days.

subjected to a programmed temperature run. Four heating scans (3, 4.5, 80:20, v/v) was used to carry out volatile products to the combustion
6.7, and 10 °C/min) were registered on each fresh sample (in nitrogen furnace (900 °C) of MCC.
flow of 50 mL/min) from 20 to 300 °C to yield the nonisothermal curing General Procedure of Synthesis. As illustrated in Scheme 2, the
curves. eugenol-based epoxy resin (TEU-EP) was synthesized by reacting
Preparation of Cure Samples. DGEBA or TEU-EP and 33DDS cyanuric chloride with eugenol to yield an intermediate (TEU) bearing
(epoxy: N−H = 1:1) without solvent were heated to 140 °C with stirring three allyl bonds followed by an epoxidation procedure to yield the
to form homogeneous mixture. Immediately, the reaction mixture was targeting epoxy monomer (TEU-EP).
casted into a preheated mold at 180 °C for 5 h. The released specimens Synthesis of TEU. In a typical procedure, eugenol (120.0 g, 710
were polished and dried in vacuum at 40 °C for further testing. mmol) was dissolved into 250 mL NaOH (29.2 g, 730 mmol) to yield a
Thermomechanical Property Characterization. The polished sodium phenolate aqueous solution, and the solution was added to the
epoxy specimens (35 mm × 10 mm × 2 mm) were subjected to the solution of cyanuric chloride (40.0 g, 216 mmol) and benzyltriethy-
thermo-mechanical analysis using a dynamic mechanical analyzer (DMA lammonium chloride (1.0 g) in chloroform (800 mL) dropwise at 10−
Q800, TA Instruments). The specimen was mounted on a single 20 °C with stirring for 1 h. The mixture was heated to 25−30 °C for 3 h
cantilever clamp, the oscillation frequency was 1 Hz, and the and then to reflux for 4 h. Subsequently, 10 g NaOH in 100 mL water
displacement was 20 μm. Testing temperature ranged from 30 to 270 was added with stirring for 30 min. The organic fraction was washed with
°C (3 °C/min). The glass temperature (Tg) was taken as the water several times and then evaporated to yield the crude yellow solid.
temperature for maximum loss factor tan δ. A static mechanical analyzer The solid was purified by digesting in the mixture of hot methanol and
(DIL 402C, Netzsch, Germany) was used to evaluate the dimensional acetone (v/v = 4:1, 400 mL) twice to remove any byproducts, reactants,
stability of the cured samples (dimension Φ 10 mm × 18 mm) from and catalyst, then washed with water, and finally dried at 100 °C in
room temperature to 250 °C (5 K/min, in nitrogen) with a statistic force vacuum for 3 h to yield white powder, viz. TEU (114.0 g, 92.6%). FTIR
of 0.2 N. (KBr): ν = 3070 (w; Ph-H and C−H), δas = 2966 (w; CH3), δas =
Nanoindentation Measurement. Nanoindentation is an effective 2937 (w; CH2), δs = 2838 (w; CH2), ν = 1638 (m, CH2), ν = 1607 and
method of probing the site-specific mechanical properties at a very fine 1508 (s, benzene ring), ω = 995 (w, CH2), γ = 909 (m, CH2). 1H
scale. Because of a very tiny plastic deformation imposed, a very small NMR (400 MHz, CDCl3-d3, δ): 3.4 (d, 6H, CH2−CHCH2), 3.8 (s,
sample is adequate for finishing testing to acquire mechanical results 9H, OCH3), 5.1 (t, 6H, CHCH2), 5.9 (m, 3H, CHCH2), 6.8 (m,
comparable to a bulk sample.47 During measurement, load−displace- 6H, CH2CH−CH2−Ph (H)), 7.1 (3H, CH2CH−CH2−Ph (H)).
ment profiles are recorded and then used to calculate Young’s (elastic) Synthesis of TEU-EP. In a typical synthesis, TEU (15.0 g) was
modulus, hardness and creep displacement.48 Nanoindentation has dissolved into CH2Cl2 (800 mL) with stirring at 0 °C, and then mCPBA
been successfully used to gauge mechanical properties of cured epoxy (60 g) added in batches. The mixture was warmed to room temperature
resins.49−51 The specimens for instrumented indentation were metal- with stirring for 5 days. After filtrating off the white solid byproduct, the
lurgically prepared with a great caution to minimize the influence of liquid phase was washed with 100 mL of 10% NaOH with stirring for 4 h
residual stresses and plastic deformation. With an extremely low load, and then with 10 g of Na2SO3 in 100 mL water with stirring overnight.
the samples were finished with a final polish using colloidal silica (0.05 The organic phase was extracted with 2% NaOH solution several times
μm). A Hysitron TI 950 nanoindenter equipped with a Berkovich and then with water until pH 7. After evaporation of the solvent, the
diamond indenter was well calibrated by fused silica before testing. The remaining solid was digested in methanol to yield a white powder (TEU-
peak indentation load was set as 1, 3, 6, 9, and 12 mN with the fixed EP) in 70% yield. FTIR (KBr): ν = 3048 (w; Ph-H), δas = 2989 (w; CH3)
loading and unloading rates of 300 and 450 μN/s, respectively. The and δas = 2939 (w; CH2), δs = 2839 (w; CH2), ν = 1605 and 1510 (s,
dwell time at the maximum indentation load was 5 s. To avoid the benzene ring), δas = 2966 (w; CH3) and δs = 2937 (w; CH3), 932 (w,
interference among the different indents, the intervals among any characteristic absorption band of epoxy group), γ = 724 (w; Ph-H). 1H
neighboring indents were greater than 100 μm. At least 36 indents were NMR (400 MHz, CDCl3-d3, δ): 2.5 (3H, epoxy ring (tran) H), 2.8 (6H,
conducted on each sample. To evaluate the creep behavior, the peak epoxy ring (cis) and CH2 H), 3.1 (3H, epoxy ring (other) H), 3.7 (s, 9H,
indentation load, loading rate, and unloading rates were fixed at 9 mN, OCH3 H), 6.8 (6H, Ph H), 7.2 (3H, Ph H). 13C NMR (101 MHz,
300 μN/s, and 450 μN/s, respectively, with the dwell time of 600 s. CDCl3, δ), 38.8 (CH2), 47.1 and 52.6 (epoxy ring), 56.0 (OCH3), 113.6,
Dielectric Property, Thermal Stability and Flammability 121.1, 121.1, 122.4, 136.5, 139.6, and 151.1 (Ph); 173.8 (triazine ring).
ESI-Mass: [M + H+] = 616.43, [M + Na+] = 638.33.


Analysis. A Concept40 broadband dielectric spectrometer (Novocon-
trol Technologies GmbH & Co. KG) was used to study dielectric
properties of the cured epoxy. The bandwidth was between 10 Hz and 2 RESULTS AND DISCUSSION
GHz, and the temperature was 25.0 °C with AC volt [Vrms] = 1.00 V. Synthesis and Structural Characterization of TEU-EP. A
The collected data were processed with the WinDETA. A thermogravi- facile strategy to prepare a triazine-bridged eugenol-based epoxy
metric analyzer (Q50, TA Instruments) was used to study the thermal
monomer was developed using cyanuric chloride to react with
decomposition of the cured epoxy. About 10 mg of cured epoxy was
heated from 40 to 850 °C (10 °C/min) under nitrogen (20 mL/min). A the hydroxyl group of eugenol in a high efficient phase-transfer
FAA Micro combustion calorimeter (MCC, Fire Testing Technology, catalytic reaction system to yield an intermediate bearing three
UK) was used to evaluate the flammability of the cured epoxy according allyl groups (TEU) in an excellent yield (92.6%). In this system,
to ASTM D7309. About 5 mg of sample was heated (1 °C/s) to 700 °C sodium eugenolate is in aqueous phase and cyanuric chloride is in
in nitrogen (80 cm3/min). A gas flow (20 cm3/min, nitrogen/oxygen = CHCl3 phase. In the presence of the phase-transfer catalyst the
2871 DOI: 10.1021/acssuschemeng.6b00479
ACS Sustainable Chem. Eng. 2016, 4, 2869−2880
ACS Sustainable Chemistry & Engineering Research Article

Figure 1. NMR spectra (CDCl3) of synthesized epoxy monomer (TEU-EP). (a) 1H NMR spectra of intermediate (TEU) and synthesized epoxy
monomer (TEU-EP). Proton signals of ally group are at 3.4, 5.1, 5.9, and 6.8 ppm, and those of epoxypropyl group are at 2.5, 2.8, and 3.1 ppm. (b) 13C
NMR spectrum of TEU-EP. The 13C chemical shifts of the epoxy rings are at 47.1 and 52.6 ppm.

Figure 2. DSC thermal and kinetic analytical curves of the curing reactions. (a) Nonisothermal thermographs from room temperature to 300 °C (10 °C/
min). (b) Multiple heating-rate nonisothermal thermographs and fractional conversion against time of TEU-EP/33DDS. (c) Effective reaction
activation energy (Eα) of TEU-EP/33DDS and averaged curing temperature against fractional conversion, α. (d) Isothermal conversion−time
dependence of TEU-EP/33DDS predicted from eq 3.

sodium eugenolate is captured and then transferred into CHCl3 reactive cyanuric chloride was used in our synthesis, there was no
phase to react with cyanuric chloride efficiently. Although highly gaseous emission during the synthesis and the whole reaction
2872 DOI: 10.1021/acssuschemeng.6b00479
ACS Sustainable Chem. Eng. 2016, 4, 2869−2880
ACS Sustainable Chemistry & Engineering Research Article

process was stable. The product could be easily to be isolated, A curing kinetic study of thermosets plays a very important
and the solvent (CHCl3) could be readily recycled. In practice, role in optimizing curing reaction cycles.33 Here the multiple
cyanuric chloride is a massively produced and used chemical with heating-rate DSC results (Figure 2b and Table 1) show that
low cost. Then, the allyl groups of TEU were epoxidized with
mCPBA in large excess to yield triazine-bridged eugenol-based Table 1. Onset and Peak Temperatures and Reaction Heat of
epoxy monomer (TEU-EP) under a mild reaction condition TEU-EP/33DDS Reaction System
(CH2Cl2/room temperature). Also, there is no need to isolate
heating rate (K/min) Tonset (°C) Tpeak (°C) reaction heat (J/g)
TEU and TEU-EP using column chromatography, which will be
attractive for their potential scaling-up synthesis for industrial 3 173 208 372
4.5 183 219 377
applications. It is worth to point out that although in the current
6.7 193 229 365
research unrenewable, toxic and high-cost mCPBA was used to
10 202 241 371
epoxidize TEU, it would be possible to find out another green
and low cost method to finish this process, which needs further
studies in the future. TEU-EP/33DDS’s exothermic curing reaction mainly occurs
Structurally, for TEU-EP, the signals related to the allyl groups over the temperature range of 150−280 °C, and the onset and
disappear, whereas the new ones due to the epoxy ring emerge peak exothermic temperatures, Tonset and Tpeak, and fractional
(Figure 1a). Additionally, IR absorption for the allyl CC bond conversional curves systematically shift to a higher temperature
at (1638 cm−1) disappears, whereas a new absorption band at as the heating rate increases. Nevertheless, the overall reaction
933 cm−1 owing to epoxy ring appears (Figure S1). The further heat changes very slightly (365−377 J/g), which implies that the
evidence from the 13C NMR spectrum (Figure 1b) of TEU-EP cure reaction arrives at essentially the same ultimate reaction
shows that carbon resonances accord well with the predicted extent. Furthermore, TEU-EP/33DDS’s normalized molar
reaction exotherm (∼100 kJ/mol epoxide) is comparable to
molecular structure. Moreover, the ESI-Mass spectrum (Figure
typical values of other epoxy-amine curing reactions.53,54
S2) further confirms that TEU-EP has the expected molecular
Therefore, TEU-EP can be sufficiently cured by 33DDS under
weight of 615.6 (experimental results [M + H+] = 616.4 and [M
suitable curing conditions.
+ Na+] = 638.3). All these results show the successful
The data collected from conversional curves (Figure 2b) are
achievement of TEU-EP. processed with the advanced isoconversional method55,56 to give
Cure Reaction, Mechanisms and Kinetic Prediction. rise to effective activation energy, Eα; see eqs 1 and 2:
Figure 2a shows the nonisothermal analytical curves of the curing
reaction of TEU-EP/33DDS and DGEBA/33DDS (10 °C/ n n
I(Eα , Tα , i)βi −1
min), from which we cannot observe any melting processes Φ(Eα) = ∑∑ = min
i=1 j≠i
I(Eα , Tα , j)βj−1 (1)
during the heating run, whereas pristine TEU-EP and 33DDS
have their respective melting temperature range (Figure S3). Tα ⎡ −E ⎤
This observation indicates that TEU-EP and 33DDS could form
a eutectoid in molecular homogeneity after removal of cosolvent
I(Eα , Tα , i) = ∫T exp⎢ α ⎥ dT
⎣ RTα ⎦
α −Δα (2)
(acetone) at low temperature (50 °C) under reduced pressure,
due probably to π−π stacking of the intermolecular aromatic where Eα is the activation energy for a given conversion α, βi and
βj are the different heating rates, I is the temperature integral
rings. Specifically, the aromatic ring of TEU-EP functions as an
approximated using a trapezoid rule, and Tα and Tα−Δα are the
electron donor, whereas that of 33DDS is electron deficient due
reaction temperature for α and Δα (typically 0.02), respectively.
to the strong electron withdrawing sulfuryl group bonded.
The effective activation energy, Eα, and averaged curing
Furthermore, the molecular branching of TEU-EP will decrease
temperature changes with α substantially (Figure 2c). As α
its crystallization ability and further destabilize TEU-EP’s and increases to 12%, the temperature increases from 176 to 198 °C,
33DDS’s respective molecular crystal lattice in the TEU-EP/ and Eα increases slightly (from ∼66 to ∼69 kJ/mol). The
33DDS mixture. These results suggest that TEU-EP/33DDS can increased temperature leads to the increased the reaction rate,
be treated as a homogeneous “single” epoxy system, which will and meanwhile engenders the rapid increase in viscosity, thus
greatly facilitate its application. Moreover, different from viscous leading to a higher energy barrier associated with the molecular
DGEBA/33DDS at room temperature, TEU-EP/33DDS is a diffusion. On the other hand, the hydroxyl groups produced from
glassy solid that can be crushed easily to fine powder, thus greatly the epoxy-amine reaction can catalyze the epoxy-amine addition
facilitating its handling, transport, and application, especially for via a hydroxyl-amine-epoxy complex mechanism,57 thus lowering
solid processing for epoxy-based molding compounds for the energetic barrier for the ring opening reaction. These two
electronic encapsulation and powder coatings. factors function in opposite ways, and the molecular diffusion
The intensive exothermic peak (Figure 2a) arises from the ring likely overweighs the hydroxyl catalytic effect, accounting for the
opening of TEU-EP under attack of the amino group of 33DDS increased Eα.
according to nucleophilic addition mechanisms. TEU-EP/ A much more dramatic increase in Eα (from ∼70 to ∼100 kJ/
33DDS shows the exothermic peak at 241 °C which is 25 °C mol) is observed for α > 70% and the rather high curing
higher than that of DGEBA/33DDS, indicating the lower temperature of >230 °C, implying a significant change of reaction
reactivity. The decreased reactivity could be attributed to the mechanisms. The increase in Eα is due probably to topological
lower polarizability of the epoxy rings of TEU-EP, because its restriction for the remaining epoxy-amine reaction. In particular,
−CH2Ph affects the neighboring epoxy ring less than the the primary amino groups (RNH2) are generally much more
−CH2OPh of DGEBA does through an electron withdrawing reactive than the corresponding substituted secondary amine
effect. Other studies also highlighted that the position of the groups (R1R2NH) to attack an epoxy ring due largely to less
epoxy ring could greatly influence the curing reactivity of steric hindrance. This difference would become more prominent
resulting biobased epoxy resins.52 for some aromatic amine-cured epoxy systems.54 Note here that
2873 DOI: 10.1021/acssuschemeng.6b00479
ACS Sustainable Chem. Eng. 2016, 4, 2869−2880
ACS Sustainable Chemistry & Engineering Research Article

Figure 3. (a) Storage modulus and dissipation factor (tan δ) against temperature with heating rate of 3 °C/min. (b) Dimension change against
temperature (5 °C/min) from room temperature to 250 °C. Onset softening temperatures of DGEBA/33DDS and TEU-EP/33DDS are 156 and 174
°C, respectively, and the averaged CTEs (25−140 °C) are 5.72 × 10−5 vs 5.94 × 10−5 K−1.

because the curing temperature at this stage is rather high, even (e.g., green solvent-free epoxy powder coatings and epoxy
above the ultimate glass transition temperature as will be molding compounds for microelectronic encapsulation) with
discussed in the following subsection, diffusion controlled easy transportation, storage, handling, and finishing.
reaction kinetics with lower activation energy seems less Dynamic Mechanical Properties of Cured Epoxy. TEU-
important.58 In this case, the reaction kinetic rate-controlled EP/33DDS shows significantly higher storage modulus (E′) than
step is probably dominated by the reaction between TEU-EP’s DGEBA/33DDS within the experimental temperature range
epoxy group and the secondary amino group derived from (Figure 3a). Even at 170 °C, TEU-EP/33DDS is still rather hard
33DDS. Such reactions have to surmount the higher energetic (E′ = 2033 MPa), but DGEBA/33DDS softens (E′ = 420 MPa).
barrier due to the much increased steric hindrance of the The glass transition temperature (Tg for tan δmax) of TEU-EP/
substituted amino groups of 33DDS to attack the epoxy ring via 33DDS is 207 °C, 33 °C higher than that of DGEBA/33DDS
the nucleophilic addition mechanism. (174 °C). To the best of our knowledge, so far no biobased epoxy
An unified model-free method, eq 3,59 was used to predict the resin could reach such a high Tg under conditions that such high
isothermal conversion: biobased content of the epoxy monomer (80 wt %) is used. In
T
this study, even in the presence of stoichiometric 33DDS, the

tα =
∫0 α exp ( ) dT
−Eα
RTα
pure TEU-EP/33DDS system can still retain the biobased
content of 61 wt %. In addition, TEU-EP/33DDS shows a
β exp( )
−Eα
broader glass relaxation than DGEBA/33DDS does, as indicated
RT0 (3) by a 9 °C increase in the half peak height width of tan δ, implying
where T0 is the given isothermal temperature and Tα is for a a broader range of cooperative segment motions involved in this
nonisothermal reaction processes to a given conversion, α, with a relaxation. The much enhanced E′ and Tg approve that TEU-EP/
constant heating rate β. Predicted conversion curves of the 33DDS possesses significantly enhanced thermomechanical
isothermal curing reaction of TEU-EP/33DDS (170, 180, 190, characteristics.
200, and 210 °C) are shown in Figure 2b. By applying the Flory The enhanced thermomechanical properties stem from the
gelation theory,60 the gel-point conversion (αgel) of stoichio- high rigidity of the triazine ring-linked eugenol structure and
metric TEU-EP/33DDS is calculated to be 40.8% where epoxy higher functionalities of TEU-EP. To illustrate, akin to 33DDS
and amino functionalities are taken as 3 and 4, respectively. could cause its cured DGEBA with reduced free volume
Raising temperature increases from 170 to 210 °C leads to a (antiplastic effect),61 the evenly distributed eugenol moieties
systematic decrease in time for αgel from 29.3 to 6.1 min. This on the triazine ring could likely reduce the free volume in cured
result implicates that in large quantity applications, it is rational TEU-EP/33DDS further. In this case, the antiplastic effect in
to process TEU-EP/33DDS at a relatively low temperature to TEU-EP/33DDS may be more pronounced, which could
prolong the pot life, and the post cure at a higher temperature, if respond for a more notably increased in the storage modulus
necessary, could be applied to reduce the curing cycle to increase at a low temperature (38% increase in E′ at 30 °C). Once at the
productivity. rubbery state, however, although the two systems hold a
In addition, to give a general impression on the storage stability relatively constant value of rubbery E′, TEU-EP/33DDS’s E′
of the epoxy system, we predicted the curing reaction carried out (∼10 MPa) is much lower than DGEBA/33DDS’s (∼21 MPa),
at room temperature (20 °C) by neglecting diffusion effect at the due probably to different contributions of the triazine-linked
deep glassy state (leading to even lower curing rate). The result eugenol and bisphenol A segments. Note here that the
indicates that TEU-EP/33DDS could reach only 4% conversion determined epoxy equivalent weight of the DGEBA (188 g/
(practically lower than this value) for 60 days, implying the mol) and TEU-EP (222 g/mol) shows no significant difference
excellent storage stability. The long storage life combined with (<20%), but the rubbery moduli of the cured epoxy resin differ
the fast curing at a relatively high processing temperature could greatly (>100%). Another reason may lie behind this finding. As
enable TEU-EP/33DDS suitable for one-pot epoxy components demonstrated in Scheme 3, the rotation of the aromatic ether
2874 DOI: 10.1021/acssuschemeng.6b00479
ACS Sustainable Chem. Eng. 2016, 4, 2869−2880
ACS Sustainable Chemistry & Engineering Research Article

Scheme 3. Illustration of the Rotation of the Aromatic Rings mechanical analysis (TMA). The TMA curves (Figure 3b) shows
along the Axis Atom in TEU-EP and DGEBA Units that the dimension change of the two epoxy systems increases
nearly linearly as the temperature increases up to ∼150 °C above
which the change becomes dramatic due to softening of the
materials. TEU-EP/33DDS displays a smaller accumulated
dimension change and a higher onset softening temperature
than DGEBA/33DDS does (174 vs 156 °C), approving that
TEU-EP/33DDS has the enhanced dimensional stability with
(−O−) will be sufficiently activated at the rubbery state and little markedly elevated upper-service temperature.
affected by other adjacent segments and cross-links, so that the In addition, we calculated the coefficient of thermal expansion
−O− linked eugenol rings and triazine could rotate more freely (CTE), α(T), from eq 4:
with the increased freedom of degree. In contrast, the benzene
ring of DGEBA’s bisphenol A unit rotating along the 1 d L (T )
α (T ) =
−C(CH3)2− linkage exhibits the decreased degree of freedom L (T ) d T (4)
due to the restriction of the two side methyl substituents on the
axial carbon. So it is reasonable to observe experimentally that the where α(T) is the temperature-dependent CTE, and L(T) is the
bisphenol A unit causes the higher rubbery modulus for the length of sample at a given temperature T. As displayed in Figure
DGEBA/33DDS network. 3b, before softening, TEU-EP/33DDS expresses a lower α(T)
Dimensional Stability of Cured Epoxy. Dimensional than DGEBA/33DDS does. For example, at 40 °C, CTE of TEU-
stability is an important consideration for epoxy materials used in EP/33DDS is 7.6% lower than that of DGEBA/33DDS. More
many demanding areas, especially for electrical and electronic importantly, over the temperature range of 25−140 °C (at glassy
(EE) applications. Dimension change of DGEBA/33DDS and state) the former shows a lower averaged CTE than the latter
TEU-EP/33DDS from room temperature to well above the does (5.72 × 10−5 vs 5.94 × 10−5 K−1). Judiciously selecting
softening temperature was evaluated based on the static thermal epoxy matrix with low CTE and high modulus plays an important

Figure 4. Typical results from the nanoindentation measurements. (a) Load−displacement curves (9000 μN for 5 s). From the releasing segments of
load−displacement curves, Young’s modulus (E) and hardness (H) are calculated from eqs 5 and 6, respectively. The different colors represents the
different intents. Comparison of E (b) and H (c) for loading of 1, 3, 6, 9, and 12 mN. (d) Creep displacement against time up to 600 s (9 mN). DGEBA/
33DDS and TEU-EP/33DDS show an accumulated creep displacement of 135.8 ± 9.5 and 128.3 ± 8.9 nm, respectively.

2875 DOI: 10.1021/acssuschemeng.6b00479


ACS Sustainable Chem. Eng. 2016, 4, 2869−2880
ACS Sustainable Chemistry & Engineering Research Article

role in ensuring reliability of semiconductor devices encapsulated Table 2. Young’s Modulus and Hardness of DGEBA/33DDS
by epoxy molding compounds (EMCs), to minimize the internal and TEU-EP/33DDS Determined from the Nanoindentation
thermal stress and cracks caused by the temperature fluctuation Measurements under Different Loadings
when devices run, since today >80% semiconductor devices are
DGEBA/33DDS TEU-EP/33DDS
embedded using EMCs.62 For example, fundamental EMC
formulations usually contain not only a small fraction of an load Young’s modulus hardness Young’s modulus hardness
(mN) (GPa) (MPa) (GPa) (MPa)
organic epoxy binder (higher CTE) but a predominant fraction
of inorganic fillers (>80 wt %) such as silica powder having lower 1 4.45 ± 0.04 402 ± 5 6.28 ± 0.04 634 ± 7
CTE as well. The greater the CTE and modulus difference 3 4.22 ± 0.03 380 ± 5 5.90 ± 0.01 597 ± 4
between an epoxy matrix and fillers, the higher the thermal stress 6 4.16 ± 0.03 376 ± 3 5.80 ± 0.02 582 ± 4
developed and therefore the higher the probability of micro- 9 4.12 ± 0.02 373 ± 2 5.74 ± 0.02 577 ± 3
cracking which would eventually lead to the failure of EMC- 12 4.10 ± 0.01 367 ± 2 5.71 ± 0.03 574 ± 4
encapsulated electronic devices.
Nanoindentation Analysis. Nanoindentation was used to
examine mechanical properties of the cured epoxy by analyzing increases slower than that of DGEBA/33DDS, with a lower
Young’s (elastic) modulus, hardness, and creep displacement. cumulative creep displacement (128.3 ± 8.9 vs 135.8 ± 9.5 nm)
The typical load−displacement curves of DGEBA/33DDS and at 600 s. Therefore, TEU-EP endows the cured epoxy with a
TEU-EP/33DDS (Figure 4a) show that parallel indents better creep resistance at room temperature, particularly within a
(different colors) are superimposed on each other, highlighting relatively short dwell time (<200 s). The improved creep
the excellent reproducibility and homogeneous (isotropic) resistance can be attributed to more rigid molecular structure of
properties of the sample. DGEBA/33DDS shows a greater TEU-EP leads to the less plastic distortion of the cured epoxy
indentation depth and therefore lower resistance to the network.
indentation force than TEU-EP/33DDS does. To illustrate, To summarize, in contrast to the standard DGEBA, TEU-EP
with a peak load of 9000 μN (for 5 s), DGEBA/33DDS, and bestows the notably improved mechanical indentation resistance,
TEU-EP/33DDS show the maximum indentation depth (hmax) increased Young’s modulus, surface hardness, and creep
of 1283 and 1051 nm, respectively. The lower hmax value could be resistance on the cured TEU-EP/33DDS system.
related to the highly rigid triazine-concentrated eugenol Broadband Dielectric Property. The dielectric properties
segments of TEU-EP/33DDS that lead to the stronger chain of DGEBA/33DDS and TEU-EP/33DDS are compared in
interaction such as π−π stacking among the aromatic rings. On Figure 5 including the permittivity, ε′, (approximate to dielectric
the other hand, although DGEBA contains the rigid bisphenol A constant k) and dielectric loss factor, ε″/ε′, over a broad
units, they are separated by a relatively flexible glycidol unit, so bandwidth (10−106 Hz). TEU-EP/33DDS (5.0−5.2) shows a
that the overall rigidity of DGEBA/33DDS at room temperature lower ε′ than DGEBA/33DDS (5.5−5.8) does (Figure 5a),
reduces. which means that less electric energy is absorbed by the material
Young’s modulus, the most important parameters character- subjected to the alternative electric field. This finding also
izing mechanical rigidity of materials, can be calculated from eq suggests that TEU-EP/33DDS may express the improved
5:63 electromagnetic wave transparency, if the material is used as
insulation materials for microwave communication, for example,
1 1−ν 1 − νi mobiles and radars. Decreased ε′ of TEU-EP/33DDS could be
= + explained based on the Debye equation, eq 7:65,66
Er E Ei (5)

where Ei and νi are Young’s modulus (1141 GPa) and Poisson’s k−1 4π ⎛ μ2 ⎞
= N ⎜αe + αd + ⎟
rate (0.07) of a diamond indenter, Er is the reduced elastic k+2 3 ⎝ 3k bT ⎠ (7)
modulus, E and ν are Young’s modulus and Poisson’s ratio of a
where k, N, αe, αd, μ, kb, and T are the dielectric constant, number
sample, respectively, and ν is taken as 0.35 for polymer
density of dipoles, electric polarization, distortion polarization,
materials.50 Sample’s hardness (H) can be determined by
orientation polarization related to the dipole moment,
dividing the peak load (Pmax) by contact area (A), eq 6.63
Boltzmann constant, and temperature, respectively.
Pmax For TEU-EP, the highly symmetric low polarizable triazine
H= ring can form a highly delocalized p−π conjugation with its
A (6)
linked three eugenol moieties via the −O− bonds, resulting in
As the indentation load increases, E and H (Table 2) of the cured more evenly distributed electrons on dipoles and thus reduced
epoxy decreases slightly, but tends to level off for above 9 mN. orientated electric polarizability, in response to the reduction in
This observation is associated with the “indentation size effect”, αe and N. Moreover, the branched molecular structure of TEU-
resulting in a higher E and H values for a rather lower load.64 EP could disturb the regular arrangement of dipoles in the TEU-
Importantly, under the same loading, TEU-EP/33DDS shows EP/33DDS network, so that the number density of the dipoles,
the much higher E and H than DGEBA/33DDS dose (Figures. N, further decreases. Furthermore, considering the molecular
4b and c). Especially, TEU-EP/33DDS shows a 39% and 55% branching of TEU-EP, the chain distribution of TEU-EP/33DDS
enhancement in E and H (9 mN), and therefore, TEU-EP/ may be more isotropic, rendering a smaller μ. As a result, the
33DDS is much more rigid and harder than DGEBA/33DDS is. decreased αe, N, and μ values causes lowered k and in turn ε′
The creep behavior of the cured epoxy was evaluated over a values of TEU-EP/33DDS.
relatively long dwell time of 600 s (9 mN) at room temperature. In addition to the lowered permittivity, TEU-EP/33DDS also
The resulting typical profiles (Figure 4d) display that the creep shows a lower dielectric dissipation factor (ε″/ε′) than DGEBA/
displacement increases quickly initially yet the increase slows 33DDS does (Figure 5b), especially in a relatively high-frequency
down after 200 s. The creep displacement of TEU-EP/33DDS range (>105 Hz). Once the polarization can no longer follow the
2876 DOI: 10.1021/acssuschemeng.6b00479
ACS Sustainable Chem. Eng. 2016, 4, 2869−2880
ACS Sustainable Chemistry & Engineering Research Article

Figure 5. (a) Permittivity, ε′, and (b) dielectric loss factor, ε″/ε′, of DGEBA/33DDS and TEU-EP/33DDS over frequency of 10 Hz-2 MHz. The
maximum ε″/ε′ values (0.027 and 0.014) appear at 6.5 × 104 Hz.

Figure 6. (a) Thermogravimetric analytical curves of cured DGEBA/33DDS and TEU-EP/33DDS from 50 to 800 °C in N2. (b) Heat release rate
(HRR)vs temperature curves from micro combination calorimeter (MCC) tests.

change of the externally applied alternative electric field, the dielectric properties, compromising as-discussed excellent
relaxation occurs,67 as indicated by the peak for ε″/ε′. thermomechanical properties and mechanical performances
Importantly, TEU-EP/33DDS shows the far lower maximum (high Tg, ultrahigh elastic modulus and harness, improved
ε″/ε′ values than DGEBA/33DDS does (0.014 vs 0.027), which creep resistance, and lowered thermal expansion), we could
indicates the less electric energy dissipated as heat. The reason foresee that TEU-EP holds a high potential to replace DGEBA as
for the lowered ε″/ε′ is likely that the highly conjugated, triazine- insulation materials for electrical and electronic (EE) applica-
linked eugenol segments will greatly restrict the relaxation of the tions.
dipoles in the TEU-EP/33DDS network, thus reducing the Thermal Decomposition and Flammability of Cured
dielectric polarization relaxation compared to DGEBA/33DDS. Epoxy. The thermal decomposition of TEU-EP/33DDS and
Noticeably, the maximum ε″/ε′ of the two systems coincides at DGEBA/33DDS was comparatively studied with TGA in N2
the same frequency (6.5 × 104 Hz), implicating the same origin (Figure 6a). No apparent weight loss up to 290 °C occurs, and
of the dielectric relaxation owing to the dipole motion in the DGEBA/33DDS begins to decompose at a higher temperature
cross-linked network, probably the sulfonyl-related dipoles from than TEU-EP/33DDS does. The onset decomposition of TEU-
33DDS. EP/33DDS could be assigned to the dissociation of the methoxyl
As we know, epoxy based materials are widely used in groups from eugenol moieties.68 Despite that, TEU-EP/33DDS
microelectronics, especially semiconductor encapsulation and starts to degrade at a high enough temperature above its Tg
printed circuit boards. To meet the ever increasing tendency (Figure 3a) and softening temperature (Figure 3b), so that it can
toward miniaturization of electronics, epoxy materials with low be used as a hard plastic safely. Also, TEU-EP/33DDS shows a
permittivity and low dielectric loss become very crucial to reduce much lower peak decomposition rate than DGEBA/33DDS does
heat generation, temperature rise, energy consumption, and and affords the much higher pyrolytic (800 °C) residual (25.1%
electric leakage. Therefore, TEU-EP would be attractive in vs 13.8%), an increment by 82%. This result is related to the
electric and microelectronic areas. In addition to the improved highly compact triazine-linked aromatic structure of TEU-EP
2877 DOI: 10.1021/acssuschemeng.6b00479
ACS Sustainable Chem. Eng. 2016, 4, 2869−2880
ACS Sustainable Chemistry & Engineering Research Article

that promotes charring and reduces liberation of gaseous FTIR spectra of TEU and TEU-EP (Figure S1), ESI-Mass
products during the anaerobic pyrolysis. Noticeably, the nitrogen spectrum of TEU-EP (Figure S2), and DSC analytical
on the triazine ring could enhance the char formation of the curves of TEU-EP and 33DDS (Figure S3) (PDF)


TEU-EP/33DDS during the pyrolysis, whereas DGEBA cannot
afford any additional nitrogen atoms to promote charring of the
cured epoxy matrix.69
AUTHOR INFORMATION
On the other hand, flammability is a very important Corresponding Author
consideration for epoxy resins used as electrical and micro- *Tel: +34 91 549 34 22. Fax: +34 91 550 30 47. E-mail: deyi.
electronic materials to reduce fire risks.62 Here flammability of wang@imdea.org.
the cured epoxy was examined according to the micro Notes
combustion calorimeter (MCC) results as shown in Figure 6b The authors declare no competing financial interest.


where the variation of the heat release rate (HHR) with
temperature is displayed. DGEBA/33DDS shows an intensive ACKNOWLEDGMENTS
exothermic peak around 370−520 °C with the peak temperature,
peak HHR and total HR values of 434 °C, 400 W/g, and 24.9 kJ/ The research leading to these results has received funding from
g, respectively. TEU-EP/33DDS shows a slightly lowered the European Union Seventh Framework Programme (FP7
maximum exothermic temperature (409 °C) yet markedly 2007-2013) under grant agreement PIIF-GA-2013-626682. In
decreased peak HHR (191 W/g) and total HR (17.6 kJ/g), , a addition, this work is partly funded by Spanish Ministry of
reduction by 52% and 29%, respectively. These results suggest Economy and Competitiveness (MINECO) under Ramón y
Cajal grant (RYC-2012-10737).


that the gaseous combustible production is reduced notably,
consisting with the aforementioned high residual yield (Figure
6a). The MCC and TG results support the low flammability of REFERENCES
TEU-EP/33DDS, which is associated with the highly aromatic (1) May, C. A. Epoxy Resins Chemistry and Technology, 2nd ed.; Marcel
nitrogen-rich triazine ring linked eugenol segments of TEU-EP. Dekker, Inc: New York, 1988.
Such highly compact aromatic rings will reduce the production of (2) Petrie, E. M. Epoxy Adhesive Formulations; McGraw-Hill
gaseous combustible by promoting charring. In short, TEU-EP Publishing: New York, 2006.
(3) Auvergne, R.; Caillol, S.; David, G.; Boutevin, B.; Pascault, J.-P.
endows the cured epoxy with the reduced flammability, hence
Biobased Thermosetting Epoxy: Present and Future. Chem. Rev. 2014,
improving the fire safety. 114, 1082−1115.

■ CONCLUSIONS
A novel triazine-linked, eugenol-based epoxy resin (TEU-EP)
(4) Raquez, J. M.; Deléglise, M.; Lacrampe, M. F.; Krawczak, P.
Thermosetting (Bio)Materials Derived from Renewable Resources: A
Critical Review. Prog. Polym. Sci. 2010, 35, 487−509.
(5) Tan, S. G.; Chow, W. S. Biobased Epoxidized Vegetable Oils Andits
has been synthesized and its molecular structure has been Greener Epoxy Blends: A Review. Polym.-Plast. Technol. Eng. 2010, 49,
identified with FTIR, 1H NMR, 13C NMR, and ESI-Mass. TEU- 1581−1590.
EP could be sufficiently cured by 3,3′-diaminodiphenyl sulfone (6) Tan, S. G.; Ahmad, Z.; Chow, W. S. Relationships of Cure Kinetics
(33DDS). The curing kinetics analysis showed that TEU-EP/ and Processing for Epoxidized Soybean Oil Bio-Thermoset. Ind. Crops
33DDS has excellent storage stability and high-temperature Prod. 2013, 43, 378−385.
processability. In comparison with DGEBA/33DDS, TEU-EP/ (7) Deng, L. L.; Shen, M. M.; Yu, J.; Wu, K.; Ha, C. Y. Preparation,
Characterization, and Flame Retardancy of Novel Rosin-Based Siloxane
33DDS expressed significantly improved thermomechanical
Epoxy Resins. Ind. Eng. Chem. Res. 2012, 51, 8178−8184.
properties (33 °C, 39%, and 55% enhancement in the glass (8) Huang, K.; Zhang, J.; Li, M.; Xia, J.; Zhou, Y. Exploration of the
transition temperature, Young’s modulus, and hardness, Complementary Properties of Biobased Epoxies Derived from Rosin
respectively). TEU-EP/33DDS was also characterized by the Diacid and Dimer Fatty Acid for Balanced Performance. Ind. Crops Prod.
better creep resistance and dimensional stability, higher 2013, 49, 497−506.
softening temperature, and lower permittivity and dielectric (9) Liu, X.; Zhang, J. High-Performance Biobased Epoxy Derived from
loss factor, as well as the much higher pyrolytic residual yield and Rosin. Polym. Int. 2010, 59, 607−609.
lowered flammability. (10) Aouf, C.; Nouailhas, H.; Fache, M.; Caillol, S.; Boutevin, B.;
Due to these attractive properties, TEU-EP holds a great Fulcrand, H. Multi-Functionalization of Gallic Acid. Synthesis of a
Novel Bio-Based Epoxy Resin. Eur. Polym. J. 2013, 49, 1185−1195.
potential as a sustainable alternative to satisfy more demanding
(11) Aouf, C.; Lecomte, J.; Villeneuve, P.; Dubreucq, E.; Fulcrand, H.
and even cutting-edge applications, in particular, in EE industries. Chemo-Enzymatic Functionalization of Gallic and Vanillic Acids:
Extending our current work, a low-cost, scalable approach to Synthesis of Bio-Based Epoxy Resins Prepolymers. Green Chem. 2012,
synthesizing TEU-EP as well as fabrication of related high- 14, 2328−2336.
performance nanocomposites will be of great interest in future (12) Cao, L.; Liu, X.; Na, H.; Wu, Y.; Zheng, W.; Zhu, J. How a Bio-
study. In particular, it would be very interesting and important to Based Epoxy Monomer Enhanced the Properties of Diglycidyl Ether of
work out the up-scaling of synthesis reaction of TEU-EP based Bisphenol a (Dgeba)/Graphene Composites. J. Mater. Chem. A 2013, 1,
more on the principles of green chemistry and using as many 5081−5088.
renewable/recyclable reagents as possible. (13) Fache, M.; Darroman, E.; Besse, V.; Auvergne, R.; Caillol, S.;


Boutevin, B. Vanillin, a Promising Biobased Building-Block for
Monomer Synthesis. Green Chem. 2014, 16, 1987−1998.
ASSOCIATED CONTENT (14) Fache, M.; Auvergne, R.; Boutevin, B.; Caillol, S. New Vanillin-
*
S Supporting Information
Derived Diepoxy Monomers for the Synthesis of Biobased Thermosets.
Eur. Polym. J. 2015, 67, 527−538.
The Supporting Information is available free of charge on the (15) Ma, S.; Liu, X.; Jiang, Y.; Tang, Z.; Zhang, C.; Zhu, J. Bio-Based
ACS Publications website at DOI: 10.1021/acssusche- Epoxy Resin from Itaconic Acid and Its Thermosets Cured with
meng.6b00479. Anhydride and Comonomers. Green Chem. 2013, 15, 245−254.

2878 DOI: 10.1021/acssuschemeng.6b00479


ACS Sustainable Chem. Eng. 2016, 4, 2869−2880
ACS Sustainable Chemistry & Engineering Research Article

(16) Huang, K.; Zhang, P.; Zhang, J.; Li, S.; Li, M.; Xia, J.; Zhou, Y. (37) Harvey, B. G.; Sahagun, C. M.; Guenthner, A. J.; Groshens, T. J.;
Preparation of Biobased Epoxies Using Tung Oil Fatty Acid-Derived Cambrea, L. R.; Reams, J. T.; Mabry, J. M. A High-Performance
C21 Diacid and C22 Triacid and Study of Epoxy Properties. Green Renewable Thermosetting Resin Derived from Eugenol. ChemSusChem
Chem. 2013, 15, 2466−2475. 2014, 7, 1964−1969.
(17) Pan, X.; Sengupta, P.; Webster, D. C. High Biobased Content (38) Neda, M.; Okinaga, K.; Shibata, M. High-Performance Bio-Based
Epoxy-Anhydride Thermosets from Epoxidized Sucrose Esters of Fatty Thermosetting Resins Based on Bismaleimide and Allyl-Etherified
Acids. Biomacromolecules 2011, 12, 2416−2428. Eugenol Derivatives. Mater. Chem. Phys. 2014, 148, 319−327.
(18) Chrysanthos, M.; Galy, J.; Pascault, J.-P. Preparation and (39) Rojo, L.; Vazquez, B.; Parra, J.; Bravo, A. L.; Deb, S.; Roman, J. S.
Properties of Bio-Based Epoxy Networks Derived from Isosorbide From Natural Products to Polymeric Derivatives of ″Eugenol″: A New
Diglycidyl Ether. Polymer 2011, 52, 3611−3620. Approach for Preparation of Dental Composites and Orthopedic Bone
(19) Feng, X.; East, A.; Hammond, W.; Ophir, Z.; Zhang, Y.; Jaffe, M. Cements. Biomacromolecules 2006, 7, 2751−2761.
Thermal Analysis Characterization of Isosorbide-Containing Thermo- (40) Thirukumaran, P.; Shakila Parveen, A.; Sarojadevi, M. Synthesis
sets. J. Therm. Anal. Calorim. 2012, 109, 1267−1275. and Copolymerization of Fully Biobased Benzoxazines from Renewable
(20) Chang, R.; Qin, J.; Gao, J. Fully Biobased Epoxy from Isosorbide Resources. ACS Sustainable Chem. Eng. 2014, 2, 2790−2801.
Diglycidyl Ether Cured by Biobased Curing Agents with Enhanced (41) Thirukumaran, P.; Shakila, A.; Muthusamy, S. Synthesis and
Properties. J. Polym. Res. 2014, 21, 501. Characterization of Novel Bio-Based Benzoxazines from Eugenol. RSC
(21) Xin, J.; Zhang, P.; Huang, K.; Zhang, J. Study of Green Epoxy Adv. 2014, 4, 7959−7966.
Resins Derived from Renewable Cinnamic Acid and Dipentene: (42) Dumas, L.; Bonnaud, L.; Olivier, M.; Poorteman, M.; Dubois, P.
Synthesis, Curing and Properties. RSC Adv. 2014, 4, 8525−8532. Eugenol-Based Benzoxazine: From Straight Synthesis to Taming of the
(22) Jaillet, F.; Darroman, E.; Ratsimihety, A.; Auvergne, R.; Boutevin, Network Properties. J. Mater. Chem. A 2015, 3, 6012−6018.
B.; Caillol, S. New Biobased Epoxy Materials from Cardanol. Eur. J. Lipid (43) Dai, J.; Jiang, Y.; Liu, X.; Wang, J.; Zhu, J. Synthesis of Eugenol-
Sci. Technol. 2014, 116, 63−73. Based Multifunctional Monomers Via a Thiol-Ene Reaction and
(23) Chen, J.; Nie, X.; Liu, Z.; Mi, Z.; Zhou, Y. Synthesis and Preparation of Uv Curable Resins Together with Soybean Oil
Application of Polyepoxide Cardanol Glycidyl Ether as Biobased Derivatives. RSC Adv. 2016, 6, 17857−17866.
Polyepoxide Reactive Diluent for Epoxy Resin. ACS Sustainable Chem. (44) Qin, J.; Liu, H.; Zhang, P.; Wolcott, M.; Zhang, J. Use of Eugenol
Eng. 2015, 3, 1164−1171. and Rosin as Feedstocks for Biobased Epoxy Resins and Study of Curing
(24) Rapi, Z.; Szolnoki, B.; Bakó, P.; Niedermann, P.; Toldy, A.; and Performance Properties. Polym. Int. 2014, 63, 760−765.
Bodzay, B.; Keglevich, G.; Marosi, G. Synthesis and Characterization of (45) Wan, J.; Gan, B.; Li, C.; Molina-Aldareguia, J. M.; Li, Z.; Wang, X.;
Biobased Epoxy Monomers Derived from D-Glucose. Eur. Polym. J. Wang, D.-Y. A Novel Biobased Epoxy Resin with High Mechanical
Stiffness and Low Flammability: Synthesis, Characterization and
2015, 67, 375−382.
(25) Hu, F.; La Scala, J. J.; Sadler, J. M.; Palmese, G. R. Synthesis and Properties. J. Mater. Chem. A 2015, 3, 21907−21921.
(46) Wan, J.; Gan, B.; Li, C.; Molina-Aldareguia, J.; Kalali, E. N.; Wang,
Characterization of Thermosetting Furan-Based Epoxy Systems.
X.; Wang, D.-Y. A Sustainable, Eugenol-Derived Epoxy Resin with High
Macromolecules 2014, 47, 3332−3342.
Biobased Content, Modulus, Hardness and Low Flammability:
(26) Maiorana, A.; Spinella, S.; Gross, R. A. Bio-Based Alternative to
Synthesis, Curing Kinetics and Structure−Property Relationship.
the Diglycidyl Ether of Bisphenol a with Controlled Materials
Chem. Eng. J. 2016, 284, 1080−1093.
Properties. Biomacromolecules 2015, 16, 1021−1031.
(47) Patel, N. G.; Sreeram, A.; Venkatanarayanan, R. I.; Krishnan, S.;
(27) Altuna, F. I.; Ruseckaite, R. A.; Stefani, P. M. Biobased
Yuya, P. A. Elevated Temperature Nanoindentation Characterization of
Thermosetting Epoxy Foams: Mechanical and Thermal Character-
Poly(Para-Phenylene Vinylene) Conjugated Polymer Films. Polym. Test.
ization. ACS Sustainable Chem. Eng. 2015, 3, 1406. 2015, 41, 17−25.
(28) De, B.; Gupta, K.; Mandal, M.; Karak, N. Biodegradable (48) Pharr, G. M.; Bolshakov, A. Understanding Nanoindentation
Hyperbranched Epoxy from Castor Oil-Based Hyperbranched Polyester Unloading Curves. J. Mater. Res. 2002, 17, 2660−2671.
Polyol. ACS Sustainable Chem. Eng. 2014, 2, 445−453. (49) Li, Y.; Rios, O.; Kessler, M. R. Thermomagnetic Processing of
(29) Ma, S.; Webster, D. C. Naturally Occurring Acids as Cross- Liquid-Crystalline Epoxy Resins and Their Mechanical Characterization
Linkers to Yield Voc-Free, High-Performance, Fully Bio-Based, Using Nanoindentation. ACS Appl. Mater. Interfaces 2014, 6, 19456−
Degradable Thermosets. Macromolecules 2015, 48, 7127−7137. 19464.
(30) Menard, R.; Negrell, C.; Fache, M.; Ferry, L.; Sonnier, R.; David, (50) Shen, L.; Wang, L.; Liu, T.; He, C. Nanoindentation and
G. From a Bio-Based Phosphorus-Containing Epoxy Monomer to Fully Morphological Studies of Epoxy Nanocomposites. Macromol. Mater.
Bio-Based Flame-Retardant Thermosets. RSC Adv. 2015, 5, 70856− Eng. 2006, 291, 1358−1366.
70867. (51) Paiva, A.; Sheller, N.; Foster, M. D.; Crosby, A. J.; Shull, K. R.
(31) Ma, S.; Liu, X.; Jiang, Y.; Fan, L.; Feng, J.; Zhu, J. Synthesis and Microindentation and Nanoindentation Studies of Aging in Pressure-
Properties of Phosphorus-Containing Bio-Based Epoxy Resin from Sensitive Adhesives. Macromolecules 2001, 34, 2269−2276.
Itaconic Acid. Sci. China: Chem. 2014, 57, 379−388. (52) Nicolau, A.; Samios, D.; Piatnick, C. M. S.; Reiznautt, Q. B.;
(32) Qin, J.; Woloctt, M.; Zhang, J. Use of Polycarboxylic Acid Derived Martini, D. D.; Chagas, A. L. On the Polymerisation of the Epoxidised
from Partially Depolymerized Lignin as a Curing Agent for Epoxy Biodiesel: The Importance of the Epoxy Rings Position, the Process and
Application. ACS Sustainable Chem. Eng. 2014, 2, 188−193. the Products. Eur. Polym. J. 2012, 48, 1266−1278.
(33) Roudsari, G. M.; Mohanty, A. K.; Misra, M. Study of the Curing (53) Wan, J.; Li, B.-G.; Fan, H.; Bu, Z.-Y.; Xu, C.-J. Nonisothermal
Kinetics of Epoxy Resins with Biobased Hardener and Epoxidized Reaction Kinetics of Dgeba with Four-Armed Starlike Polyamine with
Soybean Oil. ACS Sustainable Chem. Eng. 2014, 2, 2111−2116. Benzene Core (Mxbdp) as Novel Curing Agent. Thermochim. Acta
(34) El-Thaher, N.; Mussone, P.; Bressler, D.; Choi, P. Kinetics Study 2010, 510, 46−52.
of Curing Epoxy Resins with Hydrolyzed Proteins and the Effect of (54) Rozenberg, B. A. Kinetics, Thermodynamics and Mechanism of
Denaturants Urea and Sodium Dodecyl Sulfate. ACS Sustainable Chem. Reactions of Epoxy Oligomers with Amines. Adv. Polym. Sci. 1986, 75,
Eng. 2014, 2, 282−287. 113−165.
(35) Ding, C.; Matharu, A. S. Recent Developments on Biobased (55) Vyazovkin, S. Evaluation of Activation Energy of Thermally
Curing Agents: A Review of Their Preparation and Use. ACS Sustainable Stimulated Solid-State Reactions under Arbitrary Variation of Temper-
Chem. Eng. 2014, 2, 2217−2236. ature. J. Comput. Chem. 1997, 18, 393−402.
(36) Tachibana, Y.; Torii, J.; Kasuya, K.-i.; Funabashi, M.; Kunioka, M. (56) Vyazovkin, S. Modification of the Integral Isoconversional
Hardening Process and Properties of an Epoxy Resin with Bio-Based Method to Account for Variation in the Activation Energy. J. Comput.
Hardener Derived from Furfural. RSC Adv. 2014, 4, 55723−55731. Chem. 2001, 22, 178−183.

2879 DOI: 10.1021/acssuschemeng.6b00479


ACS Sustainable Chem. Eng. 2016, 4, 2869−2880
ACS Sustainable Chemistry & Engineering Research Article

(57) Smith, I. T. The Mechanism of the Crosslinking of Epoxide Resins


by Amines. Polymer 1961, 2, 95−108.
(58) Wan, J.; Li, C.; Bu, Z.-Y.; Xu, C.-J.; Li, B.-G.; Fan, H. A
Comparative Study of Epoxy Resin Cured with a Linear Diamine and a
Branched Polyamine. Chem. Eng. J. 2012, 188, 160−172.
(59) Vyazovkin, S. A Unified Approach to Kinetic Processing of
Nonisothermal Data. Int. J. Chem. Kinet. 1996, 28, 95−101.
(60) Odian, G. Principles of Polymerization, Fourth Ed.; John Wiley &
Sons, Inc.: Hoboken, NJ, 2004.
(61) Tu, J.; Tucker, S. J.; Christensen, S.; Sayed, A. R.; Jarrett, W. L.;
Wiggins, J. S. Phenylene Ring Motions in Isomeric Glassy Epoxy
Networks and Their Contributions to Thermal and Mechanical
Properties. Macromolecules 2015, 48, 1748−1758.
(62) Kinjo, N.; Ogata, M.; Nishi, K.; Kaneda, A.; Dusek, K. Epoxy
Molding Compounds as Encapsulation Materials for Microelectronic
Devices. Adv. Polym. Sci. 1989, 88, 1−48.
(63) Oliver, W. C.; Pharr, G. M. Measurement of Hardness and Elastic
Modulus by Instrumented Indentation: Advances in Understanding and
Refinements to Methodology. J. Mater. Res. 2004, 19, 3−20.
(64) Swadener, J. G.; George, E. P.; Pharr, G. M. The Correlation of the
Indentation Size Effect Measured with Indenters of Various Shapes. J.
Mech. Phys. Solids 2002, 50, 681−694.
(65) Yuan, C.; Jin, K.; Li, K.; Diao, S.; Tong, J.; Fang, Q. Non-Porous
Low-K Dielectric Films Based on a New Structural Amorphous
Fluoropolymer. Adv. Mater. 2013, 25, 4875−4878.
(66) Volksen, W.; Miller, R. D.; Dubois, G. Low Dielectric Constant
Materials. Chem. Rev. 2010, 110, 56−110.
(67) Kim, J.-Y.; Lee, W. H.; Suk, J. W.; Potts, J. R.; Chou, H.;
Kholmanov, I. N.; Piner, R. D.; Lee, J.; Akinwande, D.; Ruoff, R. S.
Chlorination of Reduced Graphene Oxide Enhances the Dielectric
Constant of Reduced Graphene Oxide/Polymer Composites. Adv.
Mater. 2013, 25, 2308−2313.
(68) Haw, J. F.; Schultz, T. P. Carbon-13 Cp/Mas Nmr and Ft-Ir Study
of Low-Temperature Lignin Pyrolysis. Holzforschung 1985, 39, 289.
(69) Liu, H.; Wang, X.; Wu, D. Novel Cyclotriphosphazene-Based
Epoxy Compound and Its Application in Halogen-Free Epoxy
Thermosetting Systems: Synthesis, Curing Behaviors, and Flame
Retardancy. Polym. Degrad. Stab. 2014, 103, 96−112.

2880 DOI: 10.1021/acssuschemeng.6b00479


ACS Sustainable Chem. Eng. 2016, 4, 2869−2880

You might also like