Sie sind auf Seite 1von 38

Subscriber access provided by University | of Minnesota Libraries

Article
A branched 1,6-diaminohexane-derived aliphatic polyamine as curing agent
for epoxy: Isothermal cure, network structure and mechanical properties
Jintao Wan, Cheng Li, Hong Fan, and Bo-Geng Li
Ind. Eng. Chem. Res., Just Accepted Manuscript DOI: 10.1021/acs.iecr.7b00610 Publication Date (Web): 10 Apr 2017
Downloaded from http://pubs.acs.org on April 24, 2017

Just Accepted

Just Accepted manuscripts have been peer-reviewed and accepted for publication. They are posted
online prior to technical editing, formatting for publication and author proofing. The American Chemical
Society provides Just Accepted as a free service to the research community to expedite the
dissemination of scientific material as soon as possible after acceptance. Just Accepted manuscripts
appear in full in PDF format accompanied by an HTML abstract. Just Accepted manuscripts have been
fully peer reviewed, but should not be considered the official version of record. They are accessible to all
readers and citable by the Digital Object Identifier (DOI). Just Accepted is an optional service offered
to authors. Therefore, the Just Accepted Web site may not include all articles that will be published
in the journal. After a manuscript is technically edited and formatted, it will be removed from the Just
Accepted Web site and published as an ASAP article. Note that technical editing may introduce minor
changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers
and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors
or consequences arising from the use of information contained in these Just Accepted manuscripts.

Industrial & Engineering Chemistry Research is published by the American Chemical


Society. 1155 Sixteenth Street N.W., Washington, DC 20036
Published by American Chemical Society. Copyright American Chemical Society.
However, no copyright claim is made to original U.S. Government works, or works
produced by employees of any Commonwealth realm Crown government in the course
of their duties.
Page 1 of 37 Industrial & Engineering Chemistry Research

1 A branched 1,6-diaminohexane-derived aliphatic polyamine as curing agent for epoxy: Isothermal cure, network
2 structure and mechanical properties
3
4 Jintao Wan1,2, Cheng Li1, Hong Fan1* and Bo-Geng Li1
5
6
7 1. State Key Laboratory of Chemical Engineering, College of Chemical and Biological Engineering, Zhejiang University,
8
9 Hangzhou 310027, China
10
11 2. IMDEA Materials Institute, C/Eric Kandel, 2, 28906 Getafe, Madrid, Spain.
12
13
14 Corresponding author:
15
16 *Hong Fan (E-mail: hfan@zju.edu.cn, Tel: 86-571-87957371)
17
18
Abstract: Aliphatic diamines and polyamines are long used as important curing agents for epoxy resins, especially in room
19
20
temperature-cure epoxy coatings and adhesives due to its high reactivity and low cost. Herein we systematically evaluate our
21
22
23 newly developed liquid branched aliphatic polyamine, N,N,N',N'-tetra(3-aminopropyl)-1,6-diaminohexane (TADH), as the
24
25 curing agent for bisphenol-A epoxy (DGEBA), emphasizing on the isothermal cure reaction, network structure and
26
27 mechanical properties. The isothermal curing reaction of DGEBA/TADH at 40, 50 60 and 70 oC is autocatalytic, and we
28
29 adequately simulate the curing kinetic rate with the extended autocatalytic Kamal model. Further isoconversional kinetic
30
31 analysis reveals that effective activation energy changes dramatically as the reaction goes, especially for the diffusion-
32
33 controlled stage due to the chemical vitrification (T g(DSC) =119 oC). Compared to DGEBA/1,6-diaminohexane (a starting
34
35 material of TADH), the dynamic mechanical analysis illustrates the cured DGEBA/TADH network feature three relaxations,
36
37 and shows the increased storage modulus (up to 54 oC), Alfa and Bata transition temperatures, and activation energy of glass
38
39 relaxation. Also, DGEBA/TADH shows enhanced mechanical properties: flexural strength (~95 MPa), flexural modulus
40
41 (~2440 MPa) and shear strength (~8 MPa), with a good processing ability (gel time of 130 - 150 min at 25 oC). Due to these
42
43 merits, TADH may be suitable to be used in epoxy systems, highlighting its good promise in adhesive applications.
44
45
46 Keywords: Epoxy resin; aliphatic branched polyamine; isothermal cure; structure-property relationship.
47
48
49 1. Introduction
50
51
52 Epoxy resins are mainly used in coatings, adhesives, structural composites, electronic encapsulation, printed circuit
53
54 broads, sealants, and so on, owing to their excellent mechanical, thermal and electric properties, as well as good processing
55
56 ability. Bisphenol -A epoxy resins (DGEBAs), produced from condensation of bisphenol A and epichlorohydrin, are most
57
58
59 1
60
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 2 of 37

1 commonly used to meet huge global demand (estimated roughly US$18.6 billion in 2013 1). Curing is an essential step for
2
3 processing and applications of thermosetting epoxy resins to achieve desired ultimate properties. Particularly, curing agents
4
5 determine fundamental crosslinking reaction mechanisms, and greatly affect the processing and ultimate properties of the
6
7 shaped epoxy products. Curing agents can be divided into several main catalogues according to their chemical attributes:
8
9 amines, organic anhydrides, phenolic resins, polymercaptans, imidazole, complexes, etc. Very recently, the renewed
10
11 attentions are paid to some especial epoxy curing agents with unique molecular structures 2-7
and, in particular, bio-based
12
13 sustainable curing agents 8-12. This tendency reflects two important directions in development of epoxy curing agents:
14
15 engineering ultimate properties through more sophisticated molecular design and by fostering sustainability by making use
16
17 of abundant renewable starting materials.
18
19
20 Shifting our attention to highly versatile epoxy curing agents commercially available, amine-based curing agents, bearing
21
22 more than two reactive amino hydrogen atoms, are of primary importance, because of their excellent properties such as high
23
24
and tailorable reactivity, excellent corrosion resistance, easy access, and low cost. Among them, aliphatic amine curing
25
26
agents are characterized by high reactivity towards opening epoxy rings and low cost, and they can well cure epoxy resins
27
28
29 even at room temperature, accounting for their major applications in room-temperature coatings, adhesives, potting materials
30 13.
31 and sealants However, conventional aliphatic amine curing agents always suffer from strong irritation with high vapor
32
33 pressure at room temperature, fast absorption of carbon dioxide, high toxicity, and strict dosage respective to epoxy
34
35 monomer or oligomers. Moreover, some aliphatic amine curing agents with long aliphatic chains are a crystal solid at room
36
37 temperature, thus restricting their application in room-temperature epoxy coatings and adhesives where heating is
38
39 inapplicable or infeasible.
40
41
42 To resolve these issues, simultaneously increasing molecular weights and reducing crystallization ability of aliphatic
43
44 amine curing agents are very promising 14. Increasing molecular weights will notably reduce vapor pressure and toxicity, and
45
46 increase equivalent weight of curing agent; lowering crystallization ability will result in aliphatic amines at a liquid state. To
47
48 achieve these goals, our previous work has demonstrated that introducing the branched structure into aliphatic amino
49
50 molecules is very effective15-19.
51
52
53 To illustrate, in the previous publications18-19, we reported the non-isothermal, isothermal curing kinetics and dynamic
54
55 mechanical properties of a four-directional benzene-centered aliphatic polyamine curing agent for epoxy resins.
56
57 Metaxylenediamine and acrylonitrile (CAN) were used as the starting materials for the curing agent synthesis.
58
59 2
60
ACS Paragon Plus Environment
Page 3 of 37 Industrial & Engineering Chemistry Research

1 Metaxylenediamine is a liquid at room temperature with the melting point of 14 oC, whereas hexamethylendiamine (HAD) is
2
3 much cheaper and is wildly used in huge volume (>3 billion pounds per year20) especially for PA-66 synthesis. However,
4
5 HAD is a crystal solid with the melting point of 41-42 oC, leading to their limited applications in curing agent for epoxy
6
7 resin due mainly to poor processing ability without heating. In this case, modifying HAD via a chemical means to result in
8
9 liquid curing agent to expand their application for epoxy curing agent for room-temperature purposes is of great interest. In
10
11 another paper21, we reported a new aliphatic polyamine curing agent (TADH) from HAD and acrylonitrile (CAN), with
12
13 emphasis on the non-isothermal curing kinetics of the resulting epoxy systems. TADH is a liquid at the room temperature,
14
15 thus facilitating its well mixing with epoxy at a low temperature range. In addition, the raw materials of TADH, i.e., 1,6-
16
17 diaminohexane, acrylonitrile and H 2 , are easy to access and inexpensive. Moreover, compared to 1,6-diaminohexane (Mw
18
19
=116 Da), the much increased molecular weight of TADH (Mw= 345 Da) will reduce its volatility and toxicity, and increase
20
21
equivalent weight with respect to epoxy monomers or oligomers simultaneously. On the other hand, with the updated
22
23
24 conception of sustainability in mind, HAD can be derived from biomass as glucose via a renewable process. In particular, in
25
26 2013 Rennovia Inc.20, 22 has announced that 100% biobased nylon 6,6 can be made by using renewable adipic acid and 1,6-
27
28 diaminohexane with relatively low cost. CAN, widely used in production of carbon fiber, can be derived from a biomass as
29
30 glutamic acid via a two-step method involving an its oxidative decarboxylation in water to 3-cyanopropanoic acid followed
31
32 by decarbonylation elimination23. Therefore, our newly developed aliphatic polyamine (TADH) seems more attractive based
33
34 on some important industrial criteria, because the two key starting materials at least ideally come out of renewable resources,
35
36 which will promote its sustainability.
37
38
39 Due to as-said merits of TADH, it will be interesting to further examine the applicability of TADH in epoxy system in a
40
41 more comprehensive way. To better understand the TADH-cured DGEBA system, herein report an original study of the
42
43 isothermal curing reaction kinetics of DGEBA/TADH using model-fitting and isoconversional methods, and cured network
44
45 characteristics from multiple frequency dynamic mechanical analysis and mechanical properties of the cured epoxy. We also
46
47 try to use TADH in an room-temperature epoxy adhesive formation. Our data will demonstrate that by comparing with 1,6-
48
49 diaminohexane, TADH endows the resultant DGEBA system with high reactivity, enhanced thermomechanical properties,
50
51 and flexural strength and modulus and shear strength, as well as good processing ability.
52
53
54 2. Experimental
55
56
57 2.1 Materials
58
59 3
60
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 4 of 37

1 1,6-Diaminohexane (DAH) and acrylonitrile were purchased from Shanghai Reagent Co., Ltd., China, and purified by
2
3 distillation under reduced pressure prior to use. Diglycidyl ether of bisphenol A (DGEBA) was obtained from Heli Resin Co.,
4
5 Ltd., China with the epoxide equivalent weight (EEW) of 196 g/equiv epoxy. The nonlinear aliphatic amine curing agent,
6
7 N,N,N',N'-tetra(3-aminopropyl)-1,6-diaminohexane (TADH), shown in Scheme 1, was synthesized by ourselves, and the
8
9 related spectrum data can be found in our previous publication21. Due to the importance of TADH as discussed in the
10
11 introduction section, herein we give the reproducible embodiment of TADH synthesis which is still not yet reported before.
12
13
14 To a three-necked flask equipped with a dropping funnel, a condensation column, a thermometer and a magnetic stirrer,
15
16 60 g of 1,6-diaminohexane (at 50-60 oC) and distilled water (120 ml) was mixed together to from transparent solution. At 25
17
18 o
C, to this solution acrylonitrile was added dropwise with stirring in two hours, and the mixture was heated to 40 oC for one
19
20 hour. The mixture was slowly heated to reflux temperature (about 80 oC) for 20 hours. After that, the water and unreacted
21
22 acrylonitrile was removed with a rotatory evaporator under reduced pressure. The crude produce was washed with warmed
23
24
distilled water several times, and subsequently dried an in vacuum oven at 80 oC for 5 h. Finally, the yellowish viscous liquid
25
26
product, N,N,N',N'-tetra(3-nitrilepropyl)-1,6-diaminohexane (Compound 1), was obtained in a very good yield (~93%).
27
28
29
To a high-pressure autoclave (2000 ml), Compound 1 (100 g), ethanol (95%, 1200 mL), Raney nickel (100 g) and NaOH
30
31
(48 g) were loaded and mixed together. After replacing air in the autoclave, H 2 was charged to 20 bar with vigorous agitation
32
33
34 (800 rpm) at room temperature. The hydrogenation process would spend about 10 h, as indicated by no decreased in pressure.
35
36 After that, filter off the catalyst and then concentrate the filtrate, leading to two layers. The upper layer was collected and
37
38 extracted with toluene combined with a small fraction of water several times. Then extract was concentrated and further
39
40 purified through a flash silica-gel column with menthol as the eluent. Completely drying collected fraction yields the
41
42 targeting polyamine, N,N,N',N'-tetra(3-aminopropyl)-1,6-diaminohexane (TADH) in a good yield (75%).
43
44
CH3 CH3
45
CH2 CH CH2 O O CH2 CH CH2 O O CH2 CH CH2
46
O CH3 OH CH3 n=0-1 O
47
DGEBA
48
NH2
49 H 2N
50 N H2 N
N NH2
51
TADH NH2 1,6-diaminohexane (DAH)
52 H 2N
53
54
55 Scheme 1 Molecular structures of N,N,N',N'-tetra(3-aminopropyl)-1,6-diaminohexane (TADH), 1,6-diaminohexane (DAH)
56
57
and epoxy resin (DGEBA).
58
59 4
60
ACS Paragon Plus Environment
Page 5 of 37 Industrial & Engineering Chemistry Research

1 2.2 Preparation of cured epoxy sample


2
3
4 Stoichiometric DGEBA and DAH (epoxy: N-H=1:1 by mole) were mixed well at 45 to form a transparent solution,
5
6 and immediately the mixture was poured into a preheated steeliness module. Then the module was transferred into an oven
7
8 under reduced pressure for 5 min to drive off entrapped bubbles. The following schedules were applied to cure the epoxy
9
10 resin in an air-blast oven: 70 /1.5h + 150 /2.5 h. The obtained rectangular epoxy specimens were used for the
11
12 mechanical tests. Similarly, the cured DGEBA/TADH was prepared. Thanks to the liquid feature of TADH, DGEBA and
13
14 TADH could be mixed well with stirring at room temperature.
15
16
17 2.3 Instrumentation and characterization
18
19
20 The curing reaction of DGEBA/TADH was studied using a differential scanning calorimeter (Perkin Elemer DSC-7) in
21
22 an isothermal mode. The DSC instrument was calibrated with Indium before any measurement. The reaction exotherm as a
23
24 function of time was recorded for the different temperatures of 40, 50, 60 and 70 , followed by a second run from 25 to
25
26 250 at 10 /min to measure heat residual. The glass transition temperature of completely cured epoxy resin is
27
28 determined with heating rate of 10 oC/min. Stoichiometric DGEBA and TADH were mixed well as quickly as possible at
29
30 <20 oC, and immediately ~10 mg of the reaction mixture was enclosed in aluminum curable hermetically and heated with
31
32
fast heating rate (100 oC/min) to the desired temperature of isothermal run. From these experiments, such kinetic data as
33
34
35 reaction heat, curing temperature, fractional conversion and reaction time could be extracted, and use for the quantitative
36
37 analysis of the curing reaction. A further more experiment regarding data processing can be found elsewhere 24.
38
39
40 The dynamic mechanical properties of the completely cured epoxy casting (35 mm 10 mm 2 mm) were examined
41
42 with a dynamic mechanical analyzer (DMA Q800) with the heating rate of 3 /min from -100 to well above the glass-
43
44 transition temperature of the sample. A single-cantilever mode and the multiple frequencies of 1, 3, 6, 12 and 24 Hz were
45
46 applied.
47
48
49 A Charpy impact tester (CEST, Italy) was used to determine the impact strength of the V notched cured epoxy bars
50
51 (80 mm 10 mm 4 mm) at 25 according to GB/T25671995.
52
53
54 A universal testing machine (Zwick/Roell Z020) was used to measure the shear and flexural strength and modulus of
55
56 the epoxy-amine systems. For the shear strength test, to the two bonded contact surface of the two steel pieces (after cure at
57
58 room temperature for 72 h), a vertical load was applied at the heading speed of 2 mm/min until failure. The shear strength
59 5
60
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 6 of 37

1 was estimated with = P/BL where is the shear strength, P is the maximum load applied to the bonded surface, and BL is
2
3 the bonded surface area. The flexural strength and modulus were measured according to GB T2570-1995 at temperature of
4
5 25 oC. Because DAH is a crystallized solid at room temperature, so it is immiscible with DGEBA at room temperature.
6
7 Therefore, here DAH is not applicable to DGEBA for shear strength test and gel time determination.
8
9
10 The gel time of the epoxy-amine mixture was estimated. The stoichiometric amount of DGEBA and TADH (~5 g) was
11
12 mixed well with stirring in a test tube immersed in a water bath (25 ). The reaction time for the mixture to lose its flow
13
14 ability was recorded and taken as the gel time.
15
16
17 2.3 Theoretical aspect of curing kinetics
18
19
20 Curing reactions of the epoxy resins were studied using DSC. In this case, for kinetic analysis the heat flux rate was assumed
21
22 to be directly proportional to the reaction rate; thus, the kinetic rate (conversion rate) can be represented as the following
23
24 expression:
25
26
27 1 dH
t

28
29
=
H total 0 dt
dt (1)

30
31
32 where is the conversion or the extent of cure, H the heat flow, t the reaction time, and H total is the overall reaction heat to
33
34 complete cure.
35
36
37 Two main methodologies are frequently applied to investigating the curing reaction of epoxy resins, i.e., model-fitting
38
39 kinetic and the model-free isoconversional ones. The former need a kinetic model which is used to fit experimental thermal
40
41 data to estimate the model parameters. For epoxy systems, classic nth-order reaction and the autocatalytic Kamal models 25-
42
43 26, Eqs. (1-2), are most frequently applied by different workers.
44
45
46 d
= k (T ) (1 )
n
47 (2)
48 dt
49
50
51 d
k1 (T ) + k2 (T ) m (1 ) (3)
=
n
52
53
dt
54
55
In Eq. 2, k(T) is the temperature-dependent rate constant by following the Arrhenius relation, Eq. (4). In Eq. (3), k 1 (T) is
56
57
58
the non-autocatalytic rate constant corresponding to the initial rate constant, k 2 (T) the autocatalytic rate constant associated
59 6
60
ACS Paragon Plus Environment
Page 7 of 37 Industrial & Engineering Chemistry Research

1 with the catalytic species formed in reaction systems, and m and n are the reaction orders for the non-autocatalytic and
2
3 autocatalytic reactions, respectively.
4
5
6 =k A exp ( Ea RT ) (4)
7
8
9
On the other hand, without applying any kinetic models, the model-free kinetic analytical methodology27 is based on the
10
11
12 isoconversional principle stating that the reaction rate is a sole function of temperature T; see Eq. (5):
13
14
15 d ln ( d dt ) E
16 = (5)
17 dt R
18
19
20 where E is the effective energy for a specified conversion . The several isoconversional methods with varied analytical
21
22 expressions have been advanced, with which effective activation energy E can be determined for any specific , yielding an
23
24 E - correlation or a thermal-induced polymer process which usually serve as a sensitive indicator for kinetic mechanisms.
25
26 Especially, in the past decade, isoconversional kinetic analysis of thermally induced processes occurring in polymers
27
28 received great attention in academic community, which is really a very powerful tool 28. The most frequently used include
29
30 the differential Friedman 29, integral Ozawa-Flynn-Wall 30-31, KissingerAkahiraSunose 32 and Vyazovkin methods 33-34.
31
32 Among these analytical methods, the Vyazovkin method combines the high accuracy and wide applicability to any arbitrary
33
34 temperature programs 35 including the most commonly encountered isothermal and nonisothermal thermal-induced polymer
35
36 processes because of the more advanced computing method incorporated without the interference of preexponential factors
37
38 to determined activation energy 33-34.
Here the Vyazovkin method will be used to perform the isothermal kinetic analysis.
39
40
First, a series of thermal measurement should be conducted according to varied temperature programs, and then the effective
41
42
43
activation energy E can be determined for any specific by minimizing Eq. (6):
44
45
n n J E , Ti ( t )

( E ) =
46
(6)
47
=i 1 j i J E , T j ( t )
48
49
50
51 t
E
52 J E , Ti ( t ) RTi ( t ) dt (7)
exp
53 t
54
55
56
57
58
59 7
60
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 8 of 37

1 where subscripts, i and j, denote the different experiments under varied heating programs, indicates the small-step
2
3 increment in conversion (usually 0.02, which is enough to eliminate accumulation of systematic errors in E values), and
4
5 integral J can be evaluated numerically with a trapezoid rule. Repeat this minimization procedure for each , and an E -
6
7 relationship results. E - dependence from the isoconversional analysis is instructive to untangle mechanisms or kinetic
8
9 schemes of many thermally stimulated polymer processes 36. The uncertain of obtained E was 0.1 kJ/mol based on our
10
11 designed computing program. Nevertheless, this value can be further decreased by simply adjusting some parameters in the
12
13 program, but it will exponentially increase the computing time.
14
15
16 3. Results and discussion
17
18
19 3.1 Isothermal curing reaction
20
21
22 Fig. 1 displays the isothermal heat flow-time thermographs for DGEBA/TADH at 40, 50, 60 and 70 . From these
23
24 curves, only a single exothermic peak can be observed for each run, and the peak becomes more and more intensive as the
25
26 reaction temperature increases. The reaction enthalpy is determined by integrating these exothermic peaks whose values 341,
27
28 389, 425 and 443 J/g for 40, 50, 60 and 70 , whereas the averaged overall reaction enthalpy is 456 J/g, with the uncertainty
29
30 of isothermal reaction heat within 5 J/g. From Eq. (1), the ultimate conversion for 40, 50, 60, and 70 has been
31
32 determined to be 74.8, 85.3, 93.2, and 97.1%, respectively. Also, these data suggest that DGEBA/TADH system can be well
33
34
cured up to conversion of 97.1% at 70 oC within a relative short time (<1 h), indicating the high relativity of TADH towards
35
36
DGEBA.
37
38
39
Fig. 1
40
41
42 The fractional conversion-time curves for each isothermal temperature (T c ) are shown in Fig. 2. increases rapidly with
43
44
the reaction time, then slows down, and finally approaches a limiting conversion T (Fig. 2A). Moreover, the higher T c
45
46
leads to the shorter reaction time and higher T . Nevertheless, the complete cure is not reached under these isothermal
47
48
49
conditions, due to the chemical vitrification, which causes a minority of epoxy and amine groups losing their ability to
50
51 access each other. The vitrification will occur as the glass transition temperature of a reaction system increases somewhat
52
53 higher than the isothermal cure temperatures 37. For this reason, when the isothermal cure of DGEBA/TADH is carried out at
54
55 a rather lower temperature (e.g. 40 oC) than the glass temperature of the completely cured epoxy resin (T g(DSC) =119 oC,
56
57 Fig.2B), then the chemical vitrification becomes much more influential for the curing reaction.
58
59 8
60
ACS Paragon Plus Environment
Page 9 of 37 Industrial & Engineering Chemistry Research

1 Fig. 2
2
3
4 3.2 Model-fitting kinetic analysis
5
6
7 Fig. 3 shows that reaction rate, d/dt, increases with reaction time, t, arriving at the maximum at essentially same of
8
9 0.3 for each isothermal run (Fig. 4), which indicates the isothermal cure reaction is autocatalytic and the curing temperatures
10
11 selected here little affect this mechanism. Increasing the cure temperature leads to systematically reduced time for the
12
13 maximum d/dt. Therefore, we applied the autocatalytic Kamal model, Eq. (3), to the subsequent model-fitting kinetic study
14
15 of DGEBA/TADH.
16
17
18 Fig. 3
19
20
21 Fig. 4
22
23
24 In Eq. (3), k 1 can be determined by extrapolating , and then the remaining three parameters can be evaluated using a
25
26 least-squared procedure (Original 7.5). The calculated values of the model parameters are summarized in Table 1. These
27
28 data show that k 1 and k 2 increase rapidly with T c , whereas the reaction orders, m and n, changes slightly, with m from
29
30 0.934 to 1.06 and n from 0.976 to 1.19.
31
32
33 Table 1
34
35
36 Arrhenius plots of ln k 1 and ln k 2 versus 1/T c are constructed in Fig. 5, from which an excellent linear relationship can
37
38 be established. The slopes of the fitted lines give rise to the values E a1 and E a2 . The obtained E a1 (57.4 1.8 kJ/mol) is
39
40 reasonable higher than E a2 (46.3 1.6 kJ/mol), because the autocatalytic reaction has a lower energetic barrier than the non-
41
42 autocatalytic reaction does. The autocatalytic mechanism is associated with the -OH groups from the epoxy-amine addition
43
44 that can catalyze the epoxy-amine ring opening via a more energetically favorable trimolecular transition state, i.e.,
45
46 hydroxyl-epoxy-amine intermediate 38-40.
47
48
49 Fig. 5
50
51
52 Fig. 4 compares the experimental curing rate (full line) and predicted rate from the Kamal model. At a low conversion
53
54 range, the simulated rates agree well with the experimental rate, but the markedly deviation occurs at the higher-conversion
55
56
57
58
59 9
60
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 10 of 37

1 range, especially at the lower curing temperature. This deviation is associated with the mobility of remaining epoxy and
2
3 amino groups affected greatly by the vitrification, leading to the diffusion-controlled kinetics.
4
5
6 To adequately describe the diffusion-controlled kinetics, the extended Kamal model (Eq. (8)) was used to the diffusion
7
8 control which contains an additional term, diffusion factor f() 41:
9
10
d
k1 (T ) + k2 (T ) m (1 ) f ( )
11
=
n
12 (8)
13
dt
14
15
In Eq. (8), f() can be expressed by
16
17
18
ke 1
19 f (=
) = (9)
20 kc 1 + exp C ( c )
21
22
23
where k c is the chemical reaction constant, k e the effective reaction rate constant, C the fitting constant, and c the critical
24
25
conversion, an indicative of the extent of diffusion control. The value of f() equals the ratio of the experimental curing rate
26
27
28 to the reaction rate predicated from the original Kamal model.
29
30
Fig. 6 illustrates f() as a function of for each isothermal temperature, and the experimental rate agrees well with the
31
32
33
predication from the unmodified Kamal model (f() = 1) in a relatively low-conversion range. After that, f() drops rapidly,
34
35 approaching zero eventually, as a result of the diffusion-controlled kinetics. Furthermore, elevating the curing temperature
36
37 leads to the decrease of f() at the higher conversion, because the molecular chains mobility increases with temperature.
38
39
40 Fig. 6
41
42
43 Table 2
44
45
Inserting the values of f() and into Eq. (9), then C and c were calculated using a least-squared procedure (Origin 7.5).
46
47
48
Shown in Table 2, curing temperature (T c ) increases from 40 to 70 , c increases from 0.709 to 0.948. Nevertheless, no
49
50 definite tendency of C can be found, similar to the previous reports on other epoxy-amine polymerizations 42-43.
51
52
53
Substituting the obtained k 1 , k 2 , m, n, C and c values into the extended Kamal model gives rise to the explicit rate
54
55 equation, and the simulated reaction rate is shown as the discrete dots (Fig. 7). By comparison, the experimental rates (full
56
57 line) fit well with the predicated throughout the conversion range of interest. Therefore, the extended Kamal model is
58
59 10
60
ACS Paragon Plus Environment
Page 11 of 37 Industrial & Engineering Chemistry Research

1 adequate to numerically simulate the isothermal curing reaction rate of DGEBA/TADH throughout the reaction-controlled
2
3 and the diffusion-controlled stages.
4
5
6 Fig. 7
7
8
9 3.3 Isoconversional analysis
10
11
12 To achieve a better understanding of curing mechanisms, we shall present the result of the model-free kinetic analysis
13
14 of the curing reaction by analyzing the effective activation energy (E ) very sensitive to the change of reaction mechanisms
15
16 or kinetic schemes 44-45. Fig. 8 shows E as a function of determined using the isoconversional Vyazovkin method,
17
18 expressing a strong E - correlation. E takes relatively constant value of ~50 kJ/mol up to = ~0.3, which is mainly
19
20 ascribed to the epoxy-primary amine reaction. As further increases from 0.3 to 0.8, E decreases from ~50 to ~35 kJ/mol,
21
22 owing probably to the more and more restricted diffusion of the small molecules or sol into the viscous liquid/solid reaction
23
24 interface 36. Once exceeds ~0.8, E decreases quickly from ~35 to ~20 kJ/mol, indicating the rate-determined step shifting
25
26 to the diffusion-controlled kinetics. Note here that the previous study pointed out that while proceeding in a glass transition
27
28 regime, the thermosetting reactions showed very low activation energies (only a few kilojoule or less) 46-47. Such a low value
29
30 was due to the contribution of the small-scale chain segment motions at the glassy state instead of commonly encountered
31
32 chemical reaction kinetics expressing much higher activation energy 48. Here once the curing of DGEBA/TADH proceeds
33
34
into the glass-transition region, the diffusion of separated epoxy and amino groups will become more and more influential on
35
36
overall reaction kinetics, expressing a rather low effective activation energy.
37
38
39
To conclude, on the basis of the isoconversional kinetic analysis, the isothermal curing reaction of DGEBA/TADH
40
41
transverses the reaction-controlled zone at the lower conversion stage, whereas the diffusion effect begins to take effect as
42
43
44
the high-conversion stage, and eventually dominates the overall kinetics, leading to the fast decreased effective activation
45
46 energy to a low value.
47
48
49 Fig. 8
50
51
3.4 Network structure of cured epoxy
52
53
54 Dynamic mechanical analysis (DMA) was used to study the network characteristics of the cured epoxy. Fig. 9
55
56
compared DGBEA/TADH and DGEBA/DAH networks regarding to the typical dynamic mechanical spectra for the storage
57
58
59 11
60
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 12 of 37

1 modulus (E) and loss factors (Tan ) against temperature over a low-to-high temperature range. The two networks exhibit
2
3 the high E values of 5~6 GPa at a low temperature (Fig. 9 and Table 3), showing the rather high stiffness. In comparison,
4
5 DGEBA/TADH exhibits somewhat higher E than DGEBA/DAH does at a lower temperature (up to 54 oC), indicating the
6
7 more restricted motion of the molecular chains, showing increased stiffness. This finding results from the molecular
8
9 branching of TADH providing the two additional crosslink points, which further restricts the motion of molecular chains,
10
11 causing high stiffness of the network especially at a lower temperature.
12
13
14 As the temperature rises, however, E of DGEBA/TADH decreases faster than that of DGEBA/DAH (Fig. 9), and at a
15
16 temperature (>54 oC) DGEBA/DAH shows higher E than DGEBA/TADH does (Table 3). This observation somewhat
17
18 resembles the previous finding on the DGEBA/DAH network with the tailored crosslink density: as the crosslink density
19
20 increased, the storage modulus at room-temperature range decreased 49. Also, this observation likely agrees with our recent
21
22 report on the eugenol-based epoxy monomer-aromatic diamine systems with a lower crosslinking density comprising other
23
24
kinds of aromatic epoxy resins instead of DGEBA 50-51. The reason is that the compared to DAH, TADH negatively disturbs
25
26
the stacking of DGEBA segments in the network, and moreover introduces the more flexible aliphatic chains with low
27
28
29 internal chain interaction, which will be decisive at a higher temperature. Nevertheless, at the lower temperature, the motions
30
31 of most of the segments are frozen, so that DGEBA/TADH still shows the higher stiffness than DGEBA/DAH does. On
32
33 the other hand, to maintain the high stiffness at a higher temperature, introducing more condensed aromatic structure in the
34
35 resulting curing network seems more effective52, especially for high-performance cutting-edge applications.
36
37
38 Fig. 9
39
40
41 Table 3
42
43
44
Fig. 9 also displays a minor relaxation process ( transition) from ~-30 to -60 , attributed to the crank movement of
45
46 the hydroxyl ether segments (CH 2 CH(OH)CH 2 O)53 (Scheme 2). DGEBA/DAH shows the lower -transition temperature
47
48 (T ) than DGEBA/TADH does (Table 3), which is attributed to the more restricted motion of the CH 2 CH(OH)CH 2 O
49
50 sequences of the DGEBA/TADH network, because the TADH units act as a chain concentrator to link the neighboring
51
52 chains together, thus creating the stronger adhesion in the network at a lower temperature.
53
54
55
56
57
58
59 12
60
ACS Paragon Plus Environment
Page 13 of 37 Industrial & Engineering Chemistry Research

1
2
3 relaxation O O relaxation
4
5 OH relaxation HO
6 O N N O
7 OH OH
N N
8 OH OH
9 O N N O
OH HO
10
11 O O
12
13
14
15
16
17 Scheme 2. Schematic illustration of origination of , and relaxations in DGEBA/TADH network.
18
19
20 E of the two networks decreases dramatically at a higher temperature, eventually reaching a plateau (<100 MPa), while
21
22 Tan goes through its maximum, due to the glass transition ( transition). Table 3 shows that DGEBA/TADH (136.7 oC)
23
24 has a higher transition n temperature T than DGEBA/DAH (120.5 oC) does, demonstrating that TADH endows the cured
25
26 epoxy with the improved upper-limit service temperature.
27
28
29 Table 4
30
31
32 As temperature further increases, DGEBA/TADH shows a higher rubbery modulus than DGEBA/DAH (56 MPa vs. 35
33
34 MPa). From the rubbery modulus, the crosslink density of the networks is calculated by using Eq. (10) 54:
35
36
37
Er = 3RTr ve (10)
38
39
40
41 where E r is the rubber modulus, R is the universal gas constant (8.314 J mol-1 K-1), T r the temperature for E r , and e the
42
43 crosslink density. The results (Table 3) show that the DGEBA/TADH network takes the higher e value (5195 mol m-3) than
44
45 DGEBA/DAH (3128 mol m-3). The increased e is due to the tertiary amino functionalities of TADH acting as the additional
46
47 crosslinks in the cured network, thus increasing the concentration of the effective crosslinks (Scheme 2).
48
49
50 In addition to the aforementioned and relaxations, a weak relaxation, noted as the relaxation, can be observed 30-
51
52 90 for the DGEBA/TADH network. The relaxation likely results from the motion of the tertiary amine functionalities
53
54 (Nt) from the TADH molecules (Scheme 1). To illustrate, the tertiary amine functionalities take a role of extra crosslinks in
55
56 the cured epoxy network. They are indirectly connected with the rigid bisphenol A moieties, so that the motions of these
57
58
59 13
60
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 14 of 37

1 crosslinks will cause the less cooperation of the adjacent chains. Moreover, the number of tertiary amino functionalities is
2
3 only the half of the bisphenol A moieties in the cured epoxy. These two aspects could interpret why the observed
4
5 relaxation show the low intensity compared with and relaxations.
6
7
8 The cured epoxy is further examined based on the multiple frequency DMA testing, and Tan as a function of
9
10 temperature at the different frequency is shown in Fig. 10. Increasing the frequency leads to the systematic shift of the - and
11
12 -relaxations to a higher temperature, but the peak height (peak height) and width changes very slightly. The obtained - and
13
14 -transition temperatures are summarized in Table 4. Interestingly, for both and relaxations observed at the different
15
16 frequencies, DGEBA/TADH always shows a lower peak height (Tan ) than DGEBA/DAH does, which can be attributed to
17
18 the higher crosslink density of the former network. The increased crosslinked density will decrease plastic deformation
19
20 network, leading to less energy loss under a dynamic loading cycle.
21
22
23 The activation energies of the relaxation can be calculated from the Arrhenius relation (Eq. (11)):
24
25
d ( ln f )
26
27 Ear = R (11)
28
d (1/ T )
29
30
31
where E ar is the activation energy for a relaxation, f is the frequency, and T is the transition temperature.
32
33
34 Fig. 10
35
36
37 Table 4
38
39
40 Arrhenius plots of ln f against 1/T for the and relaxations are displayed in Fig. 11 showing a good linear correlation
41
42 can be established. The slopes of the lines give rise to relaxation activation energies, E (652 85.0 kJ/mol for
43
44 DGEBA/TADH and 601 43 kJ/mol for DGEBA/DAH). These obtained E values are comparable to the previously
45
46
reported epoxy-amine networks 55-58.
The glass () relaxation of the epoxy networks can be attributed to the cooperative
47
48
motion of the whole network with a very complex molecular relaxation mechanisms 59. Here DGEBA/TADH expresses the
49
50
51 higher E value than DGEBA/DAH does, due reasonably to the much higher crosslink density of the former leading to the
52
53 more restriction of the chain movement. However, the difference is still minor, when considering the uncertainty, likely due
54
55 to the complexity of the glass transition especially for DGEBA/TADH with the higher uncertainty.
56
57
58
59 14
60
ACS Paragon Plus Environment
Page 15 of 37 Industrial & Engineering Chemistry Research

1 In contrast to the relaxation discussed, here the calculated activation energies of the relaxation (E s) are much small
2
3 (<100 kJ/mol), but comparable with and close to the reported literature data extracted from other epoxy-amine networks [14,
4
5 43, 53-55]. This similarity implicates that the relaxation of the different epoxy-amine networks follows essentially the
6
7 same molecular mechanisms. A closer examination shows that DGEBA/TADH (61.4 5.7 kJ/mol) has lower E than
8
9 DGEBA/DAH (77.1 0.65 kJ/mol) does, which is the reflection of the relaxation temperatures of the DGEBA/TADH
10
11 network more sensitive to the loading frequency (Table 4). This observation is likely associated with TADH molecules
12
13 introducing more aliphatic segments into the cured epoxy network with prolonged relaxation time.
14
15
16
17 Fig. 11
18
19 3.5 Mechanical properties and processing ability
20
21
22 The flexural strength, modulus, impact strength, shear strength and gel time of DGEBA/DAH and DGEBA/TADH
23
24 systems are compared in Table 5. The data suggest that DGEBA/TADH exhibits a higher the flexural strength and modulus
25
26 than DGEBA/DAH dose. The increased flexural strength and rigidness of the DGEBA/TADH network is attributed to the
27
28 notably higher crosslink density of DGEBA/TADH (Table 3). To illustrate, the increased crosslink density will make the
29
30 deformation of the epoxy network more difficult and need higher stress to break the increased chemical links, thus showing
31
32 higher stiffness of the network. However, DGEBA/TADH reasonably shows the lower impact strength than DGEBA/DAH
33
34 does due to the higher crosslink density leading to the reduced plastic deformation under a high speed impact loading.
35
36
37 At room temperature, liquid TADH and DGEBA can be mixed well by sampling fast stirring, and the shear strength of
38
39 the resulting epoxy adhesive can reach as high as 8.0 0.4 MPa. In contrast, DAH is a crystal solid (Mp = 41-42 ) and
40
41 immiscible with DGEBA at room temperature, suggesting its inferior processing ability. For this reason, DAH is not
42
43 applicable as a curing agent for room temperature epoxy adhesive systems, whereas TADH is compatible with DGEBA,
44
45 showing improved processing ability and applicability. Moreover, time to reach the gel point is a critically important
46
47
parameter to characterize the processing ability and applications of epoxy formations. Here, DGEBA/TADH system has
48
49
relatively short gel time (130-150 min), which indicates that TADH is suitable for the fast-cure epoxy formulations,
50
51
52 especially room temperature epoxy adhesives.
53
54
Table 5
55
56
57
4. Conclusions
58
59 15
60
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 16 of 37

1 We have systematically studied the isothermal curing reaction kinetics, network characteristics and mechanical
2
3 properties of DGEBA/TADH. The isothermal curing reaction followed the autocatalytic mechanisms, and TADH showed
4
5 the high enough reactivity to be able to well cure epoxy resin even at a relative low temperature range. The model-fitting
6
7 analysis demonstrated that the Kamal model coupled with diffusion term could well simulate isothermal curing kinetic rate
8
9 over both reaction- and diffusion-controlled stages. Furthermore, the model-free isoconversional analysis revealed the
10
11 effective activation energy (E ) changed substantially with conversion. Specifically, the isothermal cure was controlled by
12
13 chemical reaction kinetics at the early stage with the relatively constant E , whereas the diffusion-controlled kinetics
14
15 dominated in the high-conversion stage because of the chemical vitrification, leading to the observed dramatic decrease in E .
16
17
18 The cured DGEBA/TADH network expressed the higher - and -transition temperatures and crosslink density than
19
20 DGEBA/1,6-diaminohexane (DAH) did, and showed an additional relaxation in 30-90 due to the tertiary amino related
21
22 groups from TADH. Cured DGEBA/TADH showed higher activation energy for relaxation (652 85.0 kJ/mol vs. 601
23
24
43 kJ/mol) but lower activation energy for relaxation (64.4 5.7 kJ/mol kJ/mol vs. 77.1 0.65 kJ/mol) than DGEBA/DAH
25
26
does. The further mechanical property analysis demonstrated that cured DGEBA/TADH showed the higher flexural strength,
27
28
29 flexural modulus, but the lower impact strength than DGEBA/DAH does. Noticeably, TADH was miscible with DGEBA
30
31 even at room temperature in contrast to parent crystalline 1,6-Diaminohexane immiscible with DGEBA without heating,
32
33 which will greatly facilitate processing of the epoxy system. The resulting epoxy adhesive with TADH as the curing agent
34
35 showed the relatively short gel time (120-150 min) and high adhesion strength (~8 MPa).
36
37
38 In conclusion, TADH endows the resulting epoxy system with many interesting properties, thus showing high promise
39
40 as curing agent for room-temperature epoxy coatings and adhesives. In the future, low-cost, large-scale synthesis of TADH
41
42 will be of practical interest to broaden its possibilities for manty epoxy formulations of commercial importance.
43
44
45 Acknowledgements
46
47
48 We acknowledge the finical support of the Program for Changjiang Scholars and Innovative Research Team in
49
50 University, China (PCSIRT). We would like appreciate the referees to give our helpful comments.
51
52
53 References
54
55
56 1. Brochure Epoxy Resin Market Report. http://www.acmite.com/market-reports/materials/global-
57 epoxy-resin-market.html 2014.
58
59 16
60
ACS Paragon Plus Environment
Page 17 of 37 Industrial & Engineering Chemistry Research

1 2. Hu, J.; Shan, J.; Zhao, J.; Tong, Z., Water resistance and curing kinetics of epoxy resins with a novel
2 curing agent of biphenyl-containing amine synthesized by one-pot method. Thermochim. Acta 2015, 606,
3 58-65.
4 3. Raimondo, M.; Guadagno, L.; Naddeo, C.; Longo, P.; Mariconda, A.; Agovino, A., New structure of
5
diamine curing agent for epoxy resins with self-restoration ability: Synthesis and spectroscopy
6
7 characterization. J. Mol. Struct. 2017, 1130, 400-407.
8 4. Ren, R.; Xiong, X.; Ma, X.; Liu, S.; Wang, J.; Chen, P.; Zeng, Y., Isothermal curing kinetics and
9 mechanism of DGEBA epoxy resin with phthalide-containing aromatic diamine. Thermochim. Acta 2016,
10 623, 15-21.
11 5. Tan, Y.; Shao, Z.-B.; Chen, X.-F.; Long, J.-W.; Chen, L.; Wang, Y.-Z., Novel Multifunctional Organic
12 Inorganic Hybrid Curing Agent with High Flame-Retardant Efficiency for Epoxy Resin. ACS Appl. Mater.
13
Interfaces 2015, 7 (32), 17919-17928.
14
15 6. Luo, Q.; Yuan, Y.; Dong, C.; Huang, H.; Liu, S.; Zhao, J., Highly Effective Flame Retardancy of a
16 Novel DPPA-Based Curing Agent for DGEBA Epoxy Resin. Ind. Eng. Chem. Res. 2016, 55 (41), 10880-
17 10888.
18 7. Nowicki, J.; Bereska, B.; Dziwiniski, E.; Organek, M.; Ilowska, J.; Bereska, A., Cyanoethylated
19 Isophoronediamine-New Potential Hardener for Epoxy Resins. Ind. Eng. Chem. Res. 2016, 55 (7), 1827-
20 1832.
21 8. Ma, S.; Webster, D. C., Naturally Occurring Acids as Cross-Linkers To Yield VOC-Free, High-
22
23
Performance, Fully Bio-Based, Degradable Thermosets. Macromolecules 2015, 48 (19), 7127-7137.
24 9. Ding, C.; Matharu, A. S., Recent Developments on Biobased Curing Agents: A Review of Their
25 Preparation and Use. ACS Sustainable Chem. Eng. 2014, 2 (10), 2217-2236.
26 10. Auvergne, R.; Caillol, S.; David, G.; Boutevin, B.; Pascault, J.-P., Biobased Thermosetting Epoxy:
27 Present and Future. Chem. Rev. 2014, 114 (2), 1082-1115.
28 11. Baroncini, E. A.; Yadav, S. K.; Palmese, G. R.; Stanzione, J. F., Recent advances in bio-based epoxy
29 resins and bio-based epoxy curing agents. J. Appl. Polym. Sci. 2016, 133 (45), 19.
30
12. Fache, M.; Monteremal, C.; Boutevin, B.; Caillol, S., Amine hardeners and epoxy cross-linker from
31
32 aromatic renewable resources. Eur. Polym. J. 2015, 73, 344-362.
33 13. May, C. A., Epoxy Resins Chemistry and Technology (2nd Edition). Marcel Dekker, Inc: New York,
34 1988.
35 14. Wan, J.; Li, C.; Bu, Z.-Y.; Fan, H.; Li, B.-G., Acrylonitrile-capped poly(propyleneimine) dendrimer
36 curing agent for epoxy resins: Model-free isoconversional curing kinetics, thermal decomposition and
37 mechanical properties. Mater. Chem. Phys. 2013, 138 (1), 303-312.
38
15. Perrin, F. X.; Nguyen, T. M. H.; Vernet, J. L., Chemico-diffusion kinetics and TTT cure diagrams of
39
40 DGEBA-DGEBF/amine resins cured with phenol catalysts. Eur. Polym. J. 2007, 43 (12), 5107-5120.
41 16. Wan, J.; Bu, Z.-Y.; Xu, C.-J.; Li, B.-G.; Fan, H., Preparation, curing kinetics, and properties of a
42 novel low-volatile starlike aliphatic-polyamine curing agent for epoxy resins. Chem. Eng. J. 2011, 171, 357-
43 367.
44 17. Wan, J.; Li, C.; Bu, Z.-Y.; Xu, C.-J.; Li, B.-G.; Fan, H., A comparative study of epoxy resin cured with
45 a linear diamine and a branched polyamine. Chem. Eng. J. 2012, 188 (0), 160-172.
46
18. Wan, J.; Li, C.; Bu, Z.-Y.; Fan, H.; Li, B.-G., Evaluating a four-directional benzene-centered aliphatic
47
48 polyamine curing agent for epoxy resins. J. Therm. Anal. Cal. 2013, 114 (1), 365-375.
49 19. Wan, J.; Li, B.-G.; Fan, H.; Bu, Z.-Y.; Xu, C.-J., Nonisothermal reaction kinetics of DGEBA with four-
50 armed starlike polyamine with benzene core (MXBDP) as novel curing agent. Thermochim. Acta 2010, 510
51 (1-2), 46-52.
52 20. http://www.sustainablebrands.com/news_and_views/green_chemistry/rennovia-develops-process-
53 produce-100-bio-based-nylon.
54 21. Wan, J.; Li, B.-G.; Fan, H.; Bu, Z.-Y.; Xu, C.-J., Nonisothermal reaction, thermal stability and
55
56
dynamic mechanical properties of epoxy system with novel nonlinear multifunctional polyamine hardener.
57 Thermochim. Acta 2010, 511 (1-2), 51-58.
58
59 17
60
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 18 of 37

1 22. Rennovia produces 100% biobased nylon-6,6 polymer. International Sugar Journal 2013, 115
2 (1379), 757-757.
3 23. Le Notre, J.; Scott, E. L.; Franssen, M. C. R.; Sanders, J. P. M., Biobased synthesis of acrylonitrile
4 from glutamic acid. Green Chem. 2011, 13 (4), 807-809.
5
24. Opaliki, M.; Kenny, J. M.; Nicolais, L., Cure kinetics of neat and carbon-fiber-reinforced TGDDM/DDS
6
7 epoxy systems. J. Appl. Polym. Sci. 1996, 61 (6), 1025-1037.
8 25. Sourour, S.; Kamal, M. R., Differential scanning calorimetry of epoxy cure: isothermal cure kinetics.
9 Thermochim. Acta 1976, 14 (1-2), 41-59.
10 26. Musa, R. K., Thermoset characterization for moldability analysis. Polym. Eng. Sci. 1974, 14 (3), 231-
11 239.
12 27. Vyazovkin, S., Isoconversional Kinetics of Thermally Stimulated Processes. Springer International
13
Publishing AG. Part of Springer Nature: 2015.
14
15 28. Vyazovkin, S., Isoconversional Kinetics of Polymers: The Decade Past. Macromol. Rapid Commun.
16 2017, 38 (3), 1600615.
17 29. Friedman, H. L., Kinetics of thermal degradation of char-forming plastics from thermogravimetry.
18 Application to a phenolic plastic. J. Polym. Sci.: Part C 1964, 6 (1), 183-195.
19 30. Ozawa, T., A new method of analyzing thermogravimetric data. Bull. Chem. Soc. Japan. 1965, 38,
20 1881-1886.
21 31. Flynn, J. H.; Wall, L. A., General treatment of the thermogravimetry of polymers. J. Res. Nat. Bur.
22
23
Stand., A 1966, 70A (6), 487-523.
24 32. T.Akahira; T.Sunose, Method of determining activation deterioration constant of electrical insulating
25 materials. Res. Report ChibaInst. Technol. (Sci. Technol.) 1971, 16, 22-31.
26 33. Vyazovkin, S., Evaluation of activation energy of thermally stimulated solid-state reactions under
27 arbitrary variation of temperature. J. Comput. Chem. 1997, 18 (3), 393-402.
28 34. Vyazovkin, S., Modification of the integral isoconversional method to account for variation in the
29 activation energy. J. Comput. Chem. 2001, 22 (2), 178-183.
30
35. Sbirrazzuoli, N., Is the Friedman Method Applicable to Transformations with Temperature Dependent
31
32 Reaction Heat? Macromol. Chem. Phys. 2007, 208 (14), 1592-1597.
33 36. Vyazovkin, S.; Sbirrazzuoli, N., Isoconversional kinetic analysis of thermally stimulated processes in
34 polymers. Macromol. Rapid Commun. 2006, 27 (18), 1515-1532.
35 37. Gillham, J. K.; Noshay, J. A. B. A., Isothermal transitions of a thermosetting system. J. Appl. Polym.
36 Sci. 1974, 18 (4), 951-961.
37 38. Smith, I. T., The mechanism of the crosslinking of epoxide resins by amines. Polymer 1961, 2, 95-
38
108.
39
40 39. Rozenberg, B. A., Kinetics, thermodynamics and mechanism of reactions of epoxy oligomers with
41 amines. Adv. Polym. Sci. 1986, 75, 113-165.
42 40. Ehlers, J. E.; Rondan, N. G.; Huynh, L. K.; Pham, H.; Marks, M.; Truong, T. N., Theoretical study on
43 mechanisms of the epoxy-amine curing reaction. Macromolecules 2007, 40 (12), 4370-4377.
44 41. Cole, K. C.; Hechler, J. J.; Noel, D., A new approach to modeling the cure kinetics of epoxy/amine
45 thermosetting resins. 2. Application to a typical system based on bis[4-(diglycidylamino)phenyl]methane
46
and bis(4-aminophenyl) sulfone. Macromolecules 1991, 24 (11), 3098-3110.
47
48 42. Schab-Balcerzak, E.; Janeczek, H.; Kaczmarczyk, B.; Bednarski, H.; Sek, D.; Miniewicz, A., Epoxy
49 resin cured with diamine bearing azobenzene group. Polymer 2004, 45 (8), 2483-2493.
50 43. Ramrez, C.; Rico, M.; J. Lpez, B.; Montero, R. M., Study of an epoxy system cured with different
51 diamines by differential scanning calorimetry. J. Appl. Polym. Sci. 2007, 103 (3), 1759-1768.
52 44. Vyazovkin, S.; Sbirrazzuoli, N., Mechanism and kinetics of epoxy-amine cure studied by differential
53 scanning calorimetry. Macromolecules 1996, 29 (6), 1867-1873.
54 45. Vyazovkin, S., A time to search: finding the meaning of variable activation energy. Physical
55
56
Chemistry Chemical Physics 2016, 18 (28), 18643-18656.
57 46. Stutz, H.; Mertes, J., Influence of the structure on thermoset cure kinetics. J. Polym. Sci. Part A:
58 Polym. Chem. 1993, 31 (8), 2031-2037.
59 18
60
ACS Paragon Plus Environment
Page 19 of 37 Industrial & Engineering Chemistry Research

1 47. Stutz, H.; Mertes, J.; Neubecker, K., Kinetics of thermoset cure and polymerization in the glass
2 transition region. J. Polym. Sci. Part A: Polym. Chem. 1993, 31 (7), 1879-1886.
3 48. Allegra, G.; Bignotti, F.; Gargani, L.; Cociani, M., Glass transition in polymers: freezing of rotational
4 oscillations Macromolecules 1990, 23, 5326-5334.
5
49. Cukierman, S.; Halary, J.-L.; Monnerie, L., Dynamic mechanical response of model epoxy networks
6
7 in the glassy state. Polym. Eng. Sci. 1991, 31 (20), 1476-1482.
8 50. Wan, J.; Gan, B.; Li, C.; Molina-Aldareguia, J.; Kalali, E. N.; Wang, X.; Wang, D.-Y., A sustainable,
9 eugenol-derived epoxy resin with high biobased content, modulus, hardness and low flammability: Synthesis,
10 curing kinetics and structureproperty relationship. Chem. Eng. J. 2016, 284, 1080-1093.
11 51. Wan, J.; Gan, B.; Li, C.; Molina-Aldareguia, J. M.; Li, Z.; Wang, X.; Wang, D.-Y., A novel biobased
12 epoxy resin with high mechanical stiffness and low flammability: synthesis, characterization and properties.
13
J. Mater. Chem. A 2015, 3, 21907-21921.
14
15 52. Wan, J.; Zhao, J.; Gan, B.; Li, C.; Molina-Aldareguia, J.; Zhao, Y.; Pan, Y.-T.; Wang, D.-Y., Ultrastiff
16 Biobased Epoxy Resin with High Tg and Low Permittivity: From Synthesis to Properties. ACS Sustainable
17 Chem. Eng. 2016, 4 (5), 2869-2880.
18 53. Delatycki, O.; Shaw, J. C.; Williams, J. G., Viscoelastic properties of epoxy-diamine networks. J.
19 Polym. Sci. Part B: Polym. Phys. 1969, 7 (5), 753-762.
20 54. Kamon, T.; Furukawa, H., Curing mechanisms and mechanical properties of cured epoxy resins. Adv.
21 Polym. Sci. 1986, 80, 173-202.
22
23
55. Amdouni, N.; Sautereau, H.; Gerard, J. F., Epoxy composites based on glass beads. I. Viscoelastic
24 properties. J. Appl. Polym. Sci. 1992, 45 (10), 1799-1810.
25 56. Nez, L.; Fraga, F.; Castro, A.; Fraga, L., Elastic Moduli and Activation Energies for an Epoxy/m-
26 XDA System by DMA and DSC. J. Therm. Anal. Cal. 1998, 52 (3), 1013-1022.
27 57. Li, G.; Lee-Sullivan, P.; Thring, R., Determination of activation energy for glass transition of an
28 epoxy adhesive using dynamic mechanical analysis. J. Therm. Anal. Cal. 2000, 60 (2), 377-390.
29 58. Goyanes, S. N.; Knig, P. G.; Marconi, J. D., Dynamic mechanical analysis of particulate-filled epoxy
30
resin. J. Appl. Polym. Sci. 2003, 88 (4), 883-892.
31
32 59. Starkweather, H. W., Simple and complex relaxations. Macromolecules 1981, 14 (5), 1277-1281.
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59 19
60
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 20 of 37

1 Figures
2
3 Fig. 1. Heat flow-time curves of curing reaction of DGEBA/TADH at 40, 50, 60 and 70 from isothermal DSC
4
5 experiments.
6
7
8 Fig. 2. (A) Fractional conversion-time curves of DGEBA/TADH at 40, 50, 60 and 70 , (B) non-isothermal DSC trace of
9
10 completely cured DGEBA/TADH with heating rate of 10 oC/min.
11
12
13 Fig. 3. Variation of curing rate (d/dt) with t for DGEBA/TADH at 40, 50, 60, 70 .
14
15
16 Fig. 4 Experimental d/dt (line) and predicated values (dot) from Kamal model at 40, 50, 60 and 70 .
17
18
19 Fig. 5. Arrhenius plots of ln k 1 vs. 1/T and ln k 2 vs. 1/T.
20
21
22 Fig. 6. f() as a function of for the isothermal cure reaction of DGEBA/TADH at 40, 50, 60 and 70 .
23
24
25 Fig. 7. The experimental rate (line) and the predicated (dot) from the extended Kamal model coupled with a diffusion term.
26
27
28 Fig. 8. Effective activation energy (E ) as a function of for the isothermal curing reaction of DGEBA/TADH.
29
30
31 Fig. 9. E and tan vs. temperatures for DGEBA/TADH and DGEBA/(1,6-diaminohexane) DAH at 1Hz (3 /min).
32
33
34 Fig. 10. Dependency of frequency on relaxation of DGEBA/DAH (A and B) and DGEBA/TADH (C and D) networks.
35
36
37
Fig. 11. Arrhenius plots of ln f against 1/T of DGEBA/DAH (A) and DGEBA/TADH (B) networks.
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59 20
60
ACS Paragon Plus Environment
Page 21 of 37 Industrial & Engineering Chemistry Research

1
2 Tables
3
Table 1. Parameters of Kamal model for 40, 50, 60 and 70 for isothermal DGEBA/TADH cure.
4
5 T c () k 1 (min-1) k 2 (min-1) m n m+n R2
6 40 0.00688 0.0294 1.06 1.19 2.26 0.99924
7 50 0.0141 0.0473 1.01 1.04 2.05 0.99912
8 60 0.0277 0.0822 0.965 0.978 1.94 0.99957
9 70 0.0467 0.138 0.934 0.976 1.91 0.99928
10
11 Table 2. Values of critical conversion c and constant C for different curing temperatures T c .
12 T c () 40 50 60 70
13 c 0.709 0.811 0.898 0.948
14
C 61 60.4 72.6 95.4
15
16 Table 3. Relaxation temperatures, characteristic modulus and crosslink density of DGEBA/TADH and DGEBA/DAH
17 networks (1Hz).
18
DGEBA/TADH DGEBA/DAH
19
20 relaxation temperature (T )/ 136.7* 120.5
21 relaxation temperature (T )/ -38.4* -42.6
22 relaxation temperature (T )// 64.6 None
23
24 Modulus at -100 /MPa 5865* 5469
25 Modulus at 20 /MPa 2617* 2435
26 Modulus at 80 /MPa 1649* 1748
27 Rubber Modulus (T g +30) /MPa 56* 35
28 Crosslink density (g/mol) 5195 3128
29
30 * Data quoted from Reference 21
31 Table 4 - and -relaxation temperatures and peak height (Tan ) of DGEBA/DAH and DGEBA/TADH networks at
32 frequency of 1, 3, 6, 12, and 24 Hz.
33 relaxation relaxation
34 f(Hz)
35 T Peak height T Peak height
36 DGEBA/DAH 1 120.5 0.69 -41.8 0.063
37 3 122.1 0.73 -35.3 0.065
38 6 123.7 0.68 -31.2 0.066
39 12 125.6 0.70 -26.5 0.068
40 24 127.2 0.71 -21.8 0.068
41 DGEBA/TADH 1 136.7 0.50 -40.4 0.059
42
3 140.4 0.52 -28.1 0.058
43
44 6 141.3 0.53 -25.0 0.060
45 12 142.0 0.52 -20.0 0.063
46 24 143.7 0.50 -14.0 0.064
47
Table 5 Comparison of ultimate mechanical properties and gel time of DGEBA/DAH and DGEBA/TADH.
48
49 DGEBA/DAH DGEBA/TADH
50 Flexural strength (MPa) 91.7 3.6 94.6 3.0
51 Flexural Modulus (MPa) 1960 50 2440 80
52
Impact strength (kJ m-2) 4.4 1.8 2.5 0.4
53
54 Shear strength (MPa) Not applicable 8.0 0.4
55 Gel time at 25 (min) Not applicable 130 - 150
56
57
58
59 21
60
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 22 of 37

1 TOC
2
3
4
5
6
Aliphatic amine curing agent
NH2
7 H 2N
8
9 N
N
10
11 TADH NH2
12 H 2N
13 CH3
14 CH2 CH CH2 O O CH2 CH CH2
15 DGEBA
O CH3 O
16
17
18 High curing reactivity towards DGEBA.
Easy to process at room temperature.
19
20
21
22 Potentially renewable raw materials.
23
24
Good thermomechanical properties.
25
26
Good for fast-cure epoxy adhesives.
27
28
High flexible strength and modulus.
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59 22
60
ACS Paragon Plus Environment
Page 23 of 37 Industrial & Engineering Chemistry Research

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34 Fig. 1. Heat flow-time curves of curing reaction of DGEBA/TADH at 40, 50, 60 and 70 oC from isothermal
35 DSC experiments.
36
37 222x179mm (300 x 300 DPI)
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 24 of 37

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34 Fig. 2. Fractional conversion-time curves of DGEBA/TADH at 40, 50, 60 and 70 oC.
35
214x171mm (300 x 300 DPI)
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
Page 25 of 37 Industrial & Engineering Chemistry Research

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34 Fig. 3. Variation of curing rate (d/dt) with t for DGEBA/TADH at 40, 50, 60, 70 oC.
35
36 214x175mm (300 x 300 DPI)
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 26 of 37

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34 Fig. 4 Experimental d/dt (line) and predicated values (dot) from Kamal model at 40, 50, 60 and 70 oC.
35
36 220x179mm (300 x 300 DPI)
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
Page 27 of 37 Industrial & Engineering Chemistry Research

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33 Fig. 5. Arrhenius plots of ln k1 vs. 1/T and ln k2 vs. 1/T.
34
35 221x174mm (300 x 300 DPI)
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 28 of 37

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
Fig. 6. f() as a function of for the isothermal cure reaction of DGEBA/TADH at 40, 50, 60 and 70 oC.
34
35 213x169mm (300 x 300 DPI)
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
Page 29 of 37 Industrial & Engineering Chemistry Research

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33 Fig. 7. The experimental rate (line) and the predicated (dot) from the extended Kamal model coupled with a
34 diffusion term.
35
36 219x170mm (300 x 300 DPI)
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 30 of 37

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33 Fig. 8. Effective activation energy (E) as a function of for the isothermal curing reaction of DGEBA/TADH.
34
35 209x164mm (300 x 300 DPI)
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
Page 31 of 37 Industrial & Engineering Chemistry Research

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33 Fig. 9. E and tan vs. temperatures for DGEBA/TADH and DGEBA/(1,6-diaminohexane) DAH at 1Hz (3
oC/min).
34
35 225x173mm (300 x 300 DPI)
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 32 of 37

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34 Fig. 10. Dependency of frequency on relaxation of DGEBA/DAH (A and B) and DGEBA/TADH (C and D)
35 networks.
36
224x180mm (300 x 300 DPI)
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
Page 33 of 37 Industrial & Engineering Chemistry Research

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35 Fig. 10. Dependency of frequency on relaxation of DGEBA/DAH (A and B) and DGEBA/TADH (C and D)
networks.
36
37 226x187mm (300 x 300 DPI)
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 34 of 37

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
Fig. 10. Dependency of frequency on relaxation of DGEBA/DAH (A and B) and DGEBA/TADH (C and D)
34 networks.
35
36 221x175mm (300 x 300 DPI)
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
Page 35 of 37 Industrial & Engineering Chemistry Research

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34 Fig. 10. Dependency of frequency on relaxation of DGEBA/DAH (A and B) and DGEBA/TADH (C and D)
35 networks.
36
37 221x179mm (300 x 300 DPI)
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 36 of 37

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34 Fig. 11. Arrhenius plots of ln f against 1/T of DGEBA/DAH (A) and DGEBA/TADH (B) networks.
35
36 219x176mm (300 x 300 DPI)
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
Page 37 of 37 Industrial & Engineering Chemistry Research

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34 Fig. 11. Arrhenius plots of ln f against 1/T of DGEBA/DAH (A) and DGEBA/TADH (B) networks.
35
36 220x180mm (300 x 300 DPI)
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment

Das könnte Ihnen auch gefallen