You are on page 1of 10

Reactive and Functional Polymers 143 (2019) 104344

Contents lists available at ScienceDirect

Reactive and Functional Polymers


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

Synthesis of novel multi-functional fluorene-based benzoxazine resins: T


Polymerization behaviour, curing kinetics, and thermal properties
⁎ ⁎
Ting Wanga, Xuan-Yu Heb, Abdul Qadeer Dayoa,c, Jun-Yi Wanga, Jun Wanga, , Wen-bin Liua,
a
Key Laboratory of Superlight Material and Surface Technology of Ministry of Education, College of Materials Science and Chemical Engineering, Harbin Engineering
University, Harbin 150001, China
b
Xi'an Aerospace Composites Research Institute,Xi'an 710025, China
c
Department of Chemical Engineering, Balochistan University of Information Technology, Engineering and Management Sciences, Quetta 87300, Pakistan

A R T I C LE I N FO A B S T R A C T

Keywords: In this paper, a new series of novel aromatic multi-functional fluorene-based benzoxazine monomers (AMFB)
Fluorene-based benzoxazine were synthesized from the reaction of 9,9-bis(4-aminophenyl)-2,7-dihydroxy-fluorene, aniline, paraformalde-
Curing behavior hyde, and unsubstituted or substituted phenol including phenol, p-(tert-octyl)phenol, and nonylphenol, by two-
Curing kinetics stage Mannich condensation reactions. The structures of intermediates and the final benzoxazine monomers
Thermal properties
were identified via Fourier transform infrared (FTIR) and Hydrogen and Carbon nuclear magnetic resonance
(1HNMR and 13CNMR) analyses. The rheology, polymerization behaviour, curing kinetics, and thermal stability
of the obtained novel benzoxazines thermosets were measured by using rheometer, differential scanning ca-
lorimetry (DSC), FTIR, and thermogravimetric analysis (TGA), respectively. The o-nonyl phenol containing
fluorene-based benzoxazine monomer (t-BF-a-n) showed the best processability with the lowest curing energy
values. Moreover, the prepared polybenzoxazine thermosets showed excellent thermal stabilities and with
higher glass transition temperature (258–333 °C) than the traditional difunctional fluorene-based benzoxazine
resins.

1. Introduction on the properties of the thermosets [6–8]. Polybenzoxazine matrices


have broad application prospects in composite manufacturing, elec-
During the 20th century, phenolic resins were widely used in the tronic packaging, insulation, and other fields [9–11].
construction, electrical, electronic, and aerospace industries due to Recently, the synthesis of multi-functional benzoxazine monomers
their excellent mechanical strength, electrical properties, dimensional with three or more oxazine rings has attracted the industries due to the
stability, chemical corrosion resistance, less smoke emission, thermal excellent mechanical strength and thermal stabilities of the thermosets
insulation, and flame retardancy. However, the phenolic resins have [12,13]. Multicomponent star-shaped benzoxazine compounds with
practical application issues due to their inherent polymerization me- inorganic constituents are also reported in the literature [14–16]. The
chanism. For example, strong acidic or alkaline based catalysts are used fluorene-based benzoxazine resins are one of the easy and economical
during the production process, which corrodes the equipment, and the ways to achieve multi-functional benzoxazine. A series of epoxy,
final products have obvious defects due to the volatile molecules pro- polyimide, polycarbonate, and benzoxazine resins having fluorene
duced during the curing process [1]. Therefore, polybenzoxazine, a new molecule in the structure are prepared. Thermosets showed good heat
type of polymer materials; was developed which not only has the in- resistance, excellent thermal stabilities, dielectric properties, good
herent physical and mechanical properties of traditional phenolic resins mechanical properties, high transparency, and refractive index
but can be cured without adding the catalyst [2,3]. Apart from these, [17–26]. However, the fluorene molecule exists as an overhang side
polybenzoxazine thermosets also have other advantages, such as low group in these resins, which increases the polymer chain rigidity and
porosity, low broadband dielectric loss, approximate zero shrinkage, curing temperature, the ultimate thermosets showed very poor tough-
and so on [4,5]. Moreover, the polybenzoxazine resins have a very ness with higher brittleness. Therefore, it is necessary to redesign the
flexible molecular design; several kinds of monomers can be obtained molecular structure to improve the properties of polymers.
by various combinations of phenols and amines; which ultimately effect In this paper, asymmetric fluorene-based benzoxazine monomers


Corresponding authors.
E-mail addresses: wj6267@hrbeu.edu.cn (J. Wang), liuwenbin@hrbeu.edu.cn (W.-b. Liu).

https://doi.org/10.1016/j.reactfunctpolym.2019.104344
Received 17 May 2019; Received in revised form 31 July 2019; Accepted 22 August 2019
Available online 23 August 2019
1381-5148/ © 2019 Elsevier B.V. All rights reserved.
T. Wang, et al. Reactive and Functional Polymers 143 (2019) 104344

were prepared via two-stage Mannich condensation methods. The several times with saturated NaCl solution (100 mL), and the solvents
phenolic compounds used for the synthesis of the monomers contain p- were removed by water washing and rotary evaporation. The light
tert-octyl or o-nonyl chain on the phenol. The detailed synthesis, yellow solid withan81% yield was obtained.
polymerization behaviour, curing kinetics, and thermal properties of
the polymers are investigated. 2.2.5. Preparation ofAMFB monomers
Phenol (0.01 mol), p-formaldehyde (0.03 mol), b-ABF-a (0.005 mol),
2. Experimental and xylene (20 mL) were added to a 250 mL three-neck flask and re-
acted at 130 °C for 4 h, and cooled to room temperature. The pre-
2.1. Materials cipitates were produced after adding n-hexane. The produced pre-
cipitates were filtered, several times washed with ethanol and dried in a
2,7-dihydroxy-9-fluorenone, methylsulfonic acid, sodium borohy- vacuum dryer at 60 °C for 24 h to obtain the phenol-based aromatic
dride, and p-formaldehyde were obtained from Shanghai Jingchun multi-functional fluorene-based benzoxazine monomer (t-BF-a-p). The
Reagent Co., Ltd. (Shanghai, China). Xylene and n-hexane were pro- yield of the t-BF-a-p monomer was 61%.
cured from Shandong Haobo Biomaterials Co., Ltd. (Shandong, China). The same procedure was followed for the synthesis of other phenolic
Aniline, formic acid, acetic acid, ethanol, tetrahydrofuran (THF), di- compounds only phenol was substituted by p-tert-octyl phenol or o-
chloromethane, and trichloromethane were purchased from Tianjin nonylphenol. The prepared monomers were represented as t-BF-a-o, and
Kermel Chemical Reagent Co., Ltd. (Tianjin, China). Phenol, p-tert-octyl t-BF-a-n, respectively.
phenol, and o-nonylphenol were obtained from Tianjin Komeo
Chemical Reagent Co., Ltd.(Tianjin, China). Tianjin Fuyu Fine Chemical 2.3. Curing of benzoxazine resins
Co., Ltd. (Tianjin, China) supplied the trifluoroacetic anhydride, ethyl
acetate, sodium carbonate, and sodium bicarbonate. All the chemicals AMFB monomers were melted and shifted to steel molds having
used in the study were of AR grade and used as received without pur- definite test shape; subsequently, the samples were isothermally cured
ification. in an air-dry oven. The samples were cured at 150, 180, 210, and 240 °C
for 2 h at each stage. Finally, the cured thermosets were cooled to room
2.2. Preparation ofAMFB monomers temperature and polished to the appropriate test specifications using
the sandpaper. After each curing stage, the DSC and FTIR analyses were
The route followed for the synthesis of AMFB monomers is shown in performed to evaluate the curing behaviour. In the meanwhile, the
Scheme 1. completely cured thermoset samples were used to evaluate their
thermal stability resistance using the TGA test.
2.2.1. Preparation of9,9-bis(4-aminophenyl)-2,7-dihydroxy-fluorene
The 9,9-bis(4-aminophenyl)-2,7-dihydroxy-fluorene (BADHF) was 2.4. Characterization methods
prepared by the reaction of 2,7-dihydroxy-9-fluorenone with aniline.
The followed method for the synthesis method is already discussed in Fourier transform infrared spectroscopy (FTIR) spectra from 500 to
detail in our earlier published article [27]. 4000 cm−1 wavenumber were recorded on a PE Instruments spotlight
400 FTIR spectrometer. Proton and carbon nuclear magnetic resonance
2.2.2. Preparation of fluorenyl monomer having trifluoroimide protective (1HNMR and 13CNMR) spectra were produced after dissolving the
group (b-PAHF) samples in DMSO-d6 solution on a Bruker AVANCE 500 spectrometer at
BADHF (0.05 mol, 17.4 g) and THF (100 mL) were poured into a 500 MHz. Differential scanning calorimetry (DSC) test was performed at
250 mL three-neck flask and stirred for 15 min in an ice-water bath. a 10 °C·min−1 heating rate on a Q-200 differential scanning calori-
Then, the trifluoroacetic anhydride (0.15 mol, 10.5 g) was added meter, TA Instruments, USA, from 20 to 380 °C in 50 mL min−1N2 flow.
dropwise, and the solution was reacted at room temperature for 2.5 h. The rheological behaviors of the monomers and prepolymer were in-
After the reaction, the THF was evaporated under vacuum, and ethyl vestigated at 10 s relaxation time and 3.259 Pa pressure by using a
acetate (200 mL) was added. The solution was washed five times with dynamic rheometer AR-2000ex, TA Instruments, USA, with a 1 Hz
saturated aqueous sodium bicarbonate solution and three times with frequency, and 4 °C min−1as the heating rate. Thermogravimetric
distilled water. Finally, the organic phase and solvent were removed by analysis (TGA) was performed using a thermogravimetric analyzer Q50,
rotary evaporation to obtain the pale red powder and dried under va- TA Instruments, USA, at a heating rate of 10 °C min−1 under N2 or air
cuum at 80 °C. The yield of b-PAHF powder was 92%. supply at 50 mL min−1 flow rate.

2.2.3. Preparation of bisphenol benzoxazine monomer containing 3. Results and discussion


trifluoroimide protective group (b-PABF-a)
A mixture was prepared by mixing b-PAHF (0.01 mol, g), aniline 3.1. Structural characterization of intermediates and AMFB monomers
(0.02 mol), p-formaldehyde (0.06 mol, g), chlorobenzene (15 mL), and
xylene (15 mL) into a 250 mL three-neck flask and reacted at 130 °C for The FTIR spectra of intermediates and AMFB monomers are pro-
4 h, and cooled to room temperature. The n-hexane (100 mL) was added duced in Fig. 1. From the BADHF spectrum in Fig. 1(A), we can easily
and the produced precipitates were filtered and dissolved in di- observe the benzene ring skeleton absorption at 1611 and 1511 cm−1
chloromethane. Hereafter, the dichloromethane was removed by the peaks. The bending absorption of CeN was observed at 1182 cm−1. In
alkali washing with 0.5 mol/L sodium carbonate solution, water addition, the absence of 1700 cm−1peak assigned to C]O stretching
washing, and rotary evaporation. Finally, the light yellow powder (b- vibration peak of 2,7-dihydroxyfluorenone indicates that the reaction
PABF-a) was obtained (yield 86%) after vacuum drying at 80 °C. was completed. The spectrum of b-PAHF intermediate shows the broad
overlapping peak in 3100–3600 cm−1, this peak indicates the of NeH
2.2.4. Preparation of bisphenol benzoxazine monomer containing primary and OeH stretching vibration peaks of the trifluoroimide group, and
amine group (b-ABF-a) the absorption signal at 1704 cm−1 is ascribed to the stretching vibra-
A mixture of b-PABF-a (0.09 mol), sodium borohydride (0.45 mol), tion of carbonyl (C]O) groups, which indicates that the amino group in
methanol (2 mL), and ethyl acetate (200 mL) was prepared in a 250 mL BADHF has been replaced by trifluoroimide group.
three-neck flask, and the solution was stirred for 5 h at room tem- The spectrum of b-PABF-a shows the NeH stretching vibration peak
perature in nitrogen (N2) atmosphere. Later, the solution was washed at 3297 cm−1. The characteristic absorption at 1711 cm−1 was ascribed

2
T. Wang, et al. Reactive and Functional Polymers 143 (2019) 104344

H 2N NH2
O NH2
CH3SO3H
+ o
HO OH 150 C, 14h
HO OH

H H
F3C N N CF3
H H NH2 C C
H 2N NH2 F3C N N CF3
C C O O
O O + (CH2O)n
TFAA O O
HO OH THF 0 oC, 2.5 h HO OH CB/Xylene 130 C, 4 h
o
N N

BADHF
b-PAHF b-PABF-a

O O
H 2N NH2 R R
OH N N
R
+ (CH2O)n
NaBH4
O O o
EtOAc/EtOH rt, 5 h Xylene 130 C, 4 h O O
N N
N N

b-ABF-a
AMFB

C9H19 C9H19
O O O O O O

N N N N N N

O O O O O O

N N N N N N

t-BF-a-p t-BF-a-o t-BF-a-n

Scheme 1. The route followed for the synthesis of AMFB monomers.

to C]O stretching vibration. The characteristic absorption peak of confirmed by the detection of CeH out-of-plane bending vibration in
1,2,4,5- tetrasubstituted benzene was observed at 1478 cm−1. In ad- the phenyl at 928–943 cm−1 peak. These results confirm that the
dition to this, the stretching vibration of ether bonds (CeOeC) was monomers contain a benzoxazine structure [28]. The FTIR spectrum of
noticed at 1244 cm−1 absorption peak, and the CeH out-of-plane t-BF-a-o monomer was nearly similar to the FTIR spectrum of t-BF-a-p,
bending vibration in the benzene ring was recorded at 950 cm−1. These only have broad peaks of the saturated CeH stretching vibration from
peaks confirm that the benzoxazine ring in the b-PABF-a monomer. The 3100 to 2800 cm−1. In addition to this, the FTIR spectrum of t-BF-a-n
NeH stretching vibration peak of the primary amine was spotted at monomer was similar to the FTIR spectrum of t-BF-a-o monomer but has
3363 cm−1 in the b-ABF-a monomer sample. Moreover, the absence of a very broad –OH stretching vibration peak of adsorbed water mole-
carbonyl (C]O) stretching vibration peak at 1700 cm−1 indicates that cules at 3416 cm−1.
the trifluoroimide protective group was completely removed after the The 1H and 13C NMR spectra of intermediates monomers are plotted
sodium borohydride reaction. in Fig. S1. In Fig. S1(A), the hydroxyl proton on the phenyl ring was
In Fig. 1(B), the absorption peak at 3032 cm−1 corresponds to the detected at a 9.26 ppm. The chemical shift of the phenyl proton was
stretching vibration of the unsaturated CeH bond in the benzene ring. observed at 6.40–7.45 ppm. The 4.93 ppm peak is attributed to the
The stretching vibration peak at 2910 cm−1 belongs to the saturated amino group proton in the BADHF. The b-PAHF intermediate shows the
alkyl CeH bond. The oscillating peaks of methylene on the oxazine ring chemical shift of the secondary amine protons of the trifluoride imide
were observed at 1368 and 1338 cm−1, and the peaks at 1224 and was observed at 11.24 ppm. In addition to this, the hydroxyl proton on
1253 cm−1 showed the asymmetric stretching vibration of CeOeC the benzene ring was observed at 9.44 ppm, and protons of the benzene
bond, while the symmetrical stretching vibration of CeOeC bond was ring were observed by the chemical shift at 6.71–7.58 ppm.
detected at 1110 cm−1. The asymmetric stretching vibration peak of After the first Mannich condensation reaction, 9.44 ppm peak re-
CeNeC bond was located at 1158 cm−1. The benzoxazine ring was presenting the phenolic hydroxyl proton disappear. In addition, the

3
T. Wang, et al. Reactive and Functional Polymers 143 (2019) 104344

Fig. 2. The 1HNMR (A) and 13


CNMR (B) spectra of AMFB monomers.
Fig. 1. The FTIR spectra of intermediates (A) and AMFB monomers (B).

atom in fluorene ring.


oxazine ring characteristic peaks of OeCH2eN and AreCH2eN were
The 1H and 13CNMR spectra of t-BF-a-p and other new benzoxazine
observed at 5.44 and 4.74 ppm, respectively, which confirm the suc-
monomers are displayed in Fig. 2. The phenyl ring hydrogen atoms
cessful synthesis of b-PABF-a monomer. Moreover, the proton peaks
were detected at 6.60–7.43 ppm, while the oxazine ring characteristic
(δ = 11.24 ppm) of secondary amine in trifluoride imide do not change
absorption peaks of NeCH2eAr and OeCH2eN were observed at
significantly. After the removal of the trifluoride imide group from b-
4.56–4.70 and 5.33–5.40 ppm, respectively. The NeCH2eAr and
PABF-a monomer via the action of sodium borohydride, the proton
OeCH2eN characteristic absorption peaks indicate that there are two
peak (δ = 11.24 ppm) of secondary amine on trifluoride imide was
kinds of oxazine ring structures in the monomer, indicating that the
completely disappeared. The proton of oxazine ring characteristic peaks
target product t-BF-a-p was obtained. In addition to this, additional
of OeCH2eN and AreCH2eN were observed at 5.41 and 4.70 ppm,
peaks were observed depending on the phenol, such as the t-BF-a-o
respectively. Moreover, a new proton peak at 4.89 ppm, which corre-
monomer shows the protons of tert-octyl from 0.67 to 1.63 ppm. The t-
sponds to the amino group on the phenyl ring, was observed, this
BF-a-n monomer produced the peaks of hydrogen protons of nonyl
confirmed that the b-ABF-a monomer is formed.
13 group from 0.40 to 0.81 ppm.
CNMR spectra of all the intermediates are displayed in Fig. S1(B).
The 13CNMR spectra of t-BF-a-p and other new benzoxazine
For BADHF, the single peaks located at 147.8 and 152.3 ppm are as-
monomers are displayed in Fig. 2(B). The chemical shifts at 93.2–93.6
signed to the carbon atom resonances of CeNH2 and CeOH, respec-
and 59.4–59.7 ppm are assigned to the carbon atom resonances of
tively. For b-PABF, the multiple peaks around 115.6 ppm are attributed
OeCH2eN and AreCH2eN of oxazine rings, respectively. In particular,
to the carbon atom resonances of eCF3 and the single peak located at
there are four kinds of carbon atom resonances corresponding to bi-
153.2 ppm is assigned to the carbon atom resonances of C]O. For b-
sphenol-type oxazine rings (59.55 and 93. 2 ppm) and diamine-type
PABF-a, the peak at 63.5 ppm is described to the quaternary carbon
oxazine rings (59.7 and 93.6 ppm) existing in t-BF-a-p monomer

4
T. Wang, et al. Reactive and Functional Polymers 143 (2019) 104344

Fig. 3. The relationship between dynamic viscosity and temperature of AMFB


monomers. Fig. 4. DSC thermograms of AMFB polymers.

backbone. The peaks around 63.5 ppm are ascribed to the quaternary recorded at 140 °C. Furthermore, the t-BF-a-p monomer showed the
carbon atom in fluorene ring. The analysis of 13CNMR spectra of t-BF-a- increased in the viscosity value (gelation point) to early when relatively
o and t-BF-a-n monomer were nearly similar to the 13CNMR spectrum of compared to other monomers. The t-BF-a-p monomer does not have any
t-BF-a-p. substituted alkyl groups in the molecule structure, and as the tem-
The obtained FTIR,1H and 13C NMR spectra are in good agreement perature was increased, the cross-linking reaction simultaneously oc-
with the proposed molecular structure. In connection with the analysis, curred on ortho and para position of phenolic hydroxyl. Therefore, the
we can suggest that all the intermediates and AMFB monomers are t-BF-a-p monomer has the highest reaction activity and can be cured at
successfully synthesized. a slightly lower temperature.
As we know that the octyl group of t-BF-a-o monomer and nonyl
group of t-BF-a-n monomer are attached at para and ortho positions of
3.2. Rheological properties phenolic hydroxyl, respectively. The reaction activity of the ortho po-
sition is much higher than that of the para position, so, the t-BF-a-o
The viscosity variation of AMFB was determined as a function of monomer has higher reactivity and reaches the gel point temperature
temperature from 50 to 250 °C, and the produced curves are illustrated firstly. From the above plot, the processing window of the t-BF-a-p, t-BF-
in Fig. 3. In order to compare the technological properties of different a-n, and t-BF-a-o monomers was calculated to be 10, 21, and 65 °C,
AMFB monomers, the softening point and gel point temperatures were respectively.
determined at 1000 Pa·s viscosity. The monomer's processing window
was calculated from the difference of gel point and softening point 3.3. Polymerization behavior
temperatures [29,30].
We can easily observe from the plot that all the AMFB monomers 3.3.1. DSC study
were in a solid state at room temperature and gradually melted as the The dynamic DSC thermograms of AMFB polymers are shown in
temperature was increased. Among the studied AMFB monomers, the t- Fig. 4. The dynamic DSC of pristine AMFB monomers before and after
BF-a-n monomer has the lowest softening point temperature (123 °C). each curing stage was evaluated, and the produced curves are shown in
This is due to the fact that the melting points of benzoxazine monomers Fig. 5. In addition to this, the curing degree of the AMFB monomer after
decrease with an increase in the alkyl chain length in the monomer each curing stage is shown in Table 1.
structure. From the DSC curves of the neat monomers, we can easily analyze
Generally speaking, the branched polymer has two effects on visc- that all the AMFB monomers showed a sharp exothermic peak. These
osity. First, when the branched-chain is quite short, the viscosity is exothermic peaks correspond to the ring-opening polymerization of the
much lower than that of the straight-chain molecules with the same monomer. The exothermic peak temperatures of t-BF-a-p, t-BF-a-o, and
molecular weight, and it is easy to flow. If the branched-chain is very t-BF-a-n monomers were observed at 264, 267, and 270 °C, respectively.
long, the viscosity is much higher than that of the straight-chain mo- In addition to this, the initial polymerization temperatures were in the
lecules with the same molecular weight. Secondly, the effect of the range of 210–220 °C, which show that the synthesized AMFB monomers
number of branched chains on the viscosity is similar to the above. follow the thermal ring-opening mechanism of 1,3-benzoxazine. The
When the molecular weight is equal, the shorter the branched-chain is, lowest exothermic peak temperature was observed for t-BF-a-p
the smaller the steric resistance is, which leads to high resistance, low monomer, while the t-BF-a-n monomer showed the highest exothermic
viscosity, and easy flow. If there are fewer branched chains, the longer peak temperature. This shows that the t-BF-a-p monomer has higher
branched chains shows greater the steric resistance, resulting in smaller reactivity, as it does not contain any substituent group on the phenolic
resistance, higher viscosity, and less flow [37]. So, the t-BF-a-o hydroxyl, both ortho and para position can participate in the cross-
monomer has the octyl group with branched structure, which creates linking reaction during the thermal ring-opening polymerization. While
the steric hindrance and greatly restricts the molecular movement, and the t-BF-a-n monomer has lower reactivity as the ortho position con-
in the result, the monomer has higher viscosity and softening point tains long-chain alkyl, the reaction can only take place on the para
temperature (151 °C). The t-BF-a-p monomer did not reach to softening position of phenolic hydroxyl, and the para position has lower activity
point temperature and the lowest viscosity value of 1678 Pa·s was [31]. However, the t-BF-a-o monomer has moderate value even it

5
T. Wang, et al. Reactive and Functional Polymers 143 (2019) 104344

Table 1
The curing degree of monomers after isothermal curing at different tempera-
tures.
Monomer Curing degree (%)

150 °C 180 °C 210 °C

t-BF-a-p 18.3 56.4 83.7


t-BF-a-o 3.9 63.0 87.1
t-BF-a-n 34.7 63.1 84.9

contains the branched octyl structure on the para position, and the
branched octyl structure may hinder the formation of cross-linking
network structure [25]. Moreover, form the embedded DSC data in
Fig. 5, we can observe that as the length of the substituted alkyl chain
length was increased, a decrease in the exothermic enthalpy value was
recorded. These declines in the enthalpy values can be dedicated to the
dilution behavior of the alkyl chain on the oxazine ring concentration.
The t-BF-a-o monomer also shows a small heat absorption peak corre-
sponding to the melting point at 153 °C.
It can be seen that the exothermic enthalpies of all the AMFB
monomers decrease and percentage of the curing degree increase as the
curing stage proceeds to the next temperature level. This confirms that
the number of non-reacted functional groups in benzoxazine monomers
resin decreases continuously with the increasing curing degree. After
the isothermal curing at 210 °C, the curing degree of all the monomers
exceeds 83%. After the final curing stage at 240 °C, the exothermic peak
was not observed in all the synthesized monomers.This confirms that
the polymer does not contain any uncured benzoxazine groups.
In addition, the DSC curves showed that, after each curing stage, the
exothermic peak temperature moved to higher temperatures due to the
formed network structures and increased curing degree. The activity of
the functional group in the system is significantly hindered by the
formed cross-linking network and reduced the polymerization rate.
Moreover, an increase in the temperature reduced the system viscosity
and increased the reactivity of functional groups, and ultimately, the
complete curing can be achieved at higher temperatures [31,32].
The glass transition temperatures (Tg) of the fluorene-based poly-
benzoxazines were observed after complete curing procedure, the va-
lues for poly(t-BF-a-p), poly(t-BF-a-o) and poly(t-BF-a-n) were recorded
at 338, 328 and 256 °C, respectively. From the Tg, we can see that the
long nonyl chain containing poly(t-BF-a-n) has the lowest values, while
the simple phenol-based poly(t-BF-a-p) has the highest Tg value. The
fluorine molecule is rigid and the addition of the molecule in molecular
backbone restrains the internal rotations of the polymer segments. In
addition to this, the Tg value of all the prepared thermosets are much
higher than the typical bifunctional fluorene-based polybenzoxazine
(poly(BF-p), 242 °C by the DSC test) [25], and the Tg values were much
lower than the furan-containing tetrafunctional fluorene-based ben-
zoxazine (poly(t-BF-f), 440 °C by the DMA test) [27].

3.3.2. FTIR study


The FTIR spectra of all the synthesized AMFB monomers after iso-
thermal heating at different curing temperatures are shown in Fig. 6.
The FTIR spectra from 4000 to 2000 cm−1 of all the AMFB mono-
mers after different curing stages are nearly similar to each other. As
can be seen, the oxazine ring characteristic peak at 928–943 cm−1of t-
BF-a-p monomer has been greatly reduced after heating at 180 °C. The
strength of the oscillating vibration peaks of methylene on the oxazine
ring at 1368 and 1338 cm−1, the CeOeC stretching peaks at 1253 and
1224 cm−1, and the CeNeC vibration peaks at 1158 and
1110 cm−1gradually decreased as the curing proceeds, and can com-
pletely disappear after curing at 240 °C. In addition to this, after the
Fig. 5. DSC thermograms of AMFB monomers after isothermal curing at dif- oxazine ring-opening, the AreOH characteristic absorption peak ap-
ferent temperatures, t-BF-a-p (A), t-BF-a-o (B) and t-BF-a-n (C). peared at 3371 cm−1, which indicated that the oxazine ring was opened
and formed CeO bond at higher temperatures. After curing at 240 °C,

6
T. Wang, et al. Reactive and Functional Polymers 143 (2019) 104344

the benzoxazine monomers were completely cured, these results are


inconsistent with the DSC results. Moreover, the intermolecular and
intramolecular hydrogen bond in the polybenzoxazine network struc-
tures were observed by the broad peak at 3200–3600 cm−1. The ex-
pected intermolecular and intramolecular hydrogen bond include
eOH⋯O (3320–3400 cm−1), eOH⋯π (3550–3600 cm−1), eOH⋯O
(3430–3470 cm−1), and eOH⋯N (3140–3190 cm−1), and so on [31].

3.4. Curing kinetics of AMFB monomers

As can be seen, the DSC curves of curing reactions are conducted at


different heating rates of 5, 10, 15, and 20 °C·min−1, with exothermic
peaks appearing at different temperatures. In this study, the activation
energy (Ek, Eo) is calculated by the Kissinger and Ozawa methods using
Eqs. (1) and (2) [33].

⎛ β ⎞ AR E
− k
ln ⎜ 2 ⎟ = ln
T
⎝ ⎠p Ek RT p (1)

AE E
ln(β ) = ln ⎛ o ⎞ − ln F (α ) − 1.052 ⎛ o ⎞
⎝ R ⎠ ⎝ RT ⎠ (2)
here, the Ek and Eo show the activation energy calculated by Kissinger
and Ozawa methods, respectively, while the β, Tp, R, F(α) represent the
heating rate, peak exotherm temperature (K), the universal gas con-
stant, and constant function, respectively.
The value of activation energy can be calculated from the ln (β/Tp2)
versus 1/Tp plot, where the slope of the plot show (−Ek/R) in the case
of Kissinger method. While the Ozawa method calculation needs lnβ
versus 1/Tp plot, and the slope of the plots represents the (−1.052(Eo/
R)).
The DSC scanning curves of all the synthesized AMFB monomers at
different heating rates are shown in Fig. 7.
From the dynamic DSC test curves, we can easily evaluate that the
curing exothermic peaks of the all the benzoxazine monomer moved
towards the higher temperatures and the peak height increased as the
heating rate of the DSC was increased. On the lower heating rate, the
monomers have sufficient time for curing reaction, as the DSC scanning
rate was increased, the time of curing reaction was reduced, and the
thermal effect of the reaction increased in unit time, the thermal inertia
increases, and the exothermic peak temperature moved towards the
higher temperatures.

3.4.1. Kissinger and Ozawa methods


The linear regression curves of Kissinger and Ozawa methods for all
the synthesized AMFB benzoxazine monomers are shown in Fig. S2, and
the apparent activation energy values obtained from Kissinger and
Ozawa methods are summarized in Table 2.
Both Kissinger and Ozawa linear regression curves for all the AMFB
monomers showed a good linear relationship. The calculated apparent
activation energies for the benzoxazine monomer's curing reactions
were in the range of 132.6–145.4 kJ/mol by Kissinger method and
134.5–146.5 kJ/mol by Ozawa method. The activation energy values of
studied benzoxazine monomers decrease significantly with the increase
in the alkyl chain length. This is because of the longer alkyl chain
monomer has lower viscosity, which improves the curing reaction.

3.4.2. Starink method


Starink proposes an accurate method after modifying the
Flynn–Wall–Ozawa and Kissinger–Akahira–Sunose methods for the
calculation of the conversion rate (α). The Starink equation is as follow
[34].

βi ⎞ Ea
Fig. 6. The FTIR spectra of AMFB monomers after isothermal curing at different ln ⎛⎜ 1.92 ⎟ = Const − 1.0008
temperatures, t-BF-a-p (A), t-BF-a-o (B) and t-BF-a-n (C). T
⎝ a, i ⎠ RTa, i (3)
The relationship between conversion (α) of AMFB monomer against

7
T. Wang, et al. Reactive and Functional Polymers 143 (2019) 104344

Table 2
The apparent activation energies calculated using the Kissinger and Ozawa
methods.
Samples Ek (kJ/mol) Eo (kJ/mol)

t-BF-a-p 145.4 146.5


t-BF-a-o 138.9 140.4
t-BF-a-n 135.3 137.0

Fig. 8. TGA thermograms of AMFB polymers under nitrogen atmosphere.

Table 3
Thermal stability parameters of cured AMFB thermosets.
Samples T5 (°C) T10 (°C) Yc (%, 800 °C) Reference

Poly(t-BF-a-p) 347 395 58.1 Current work


Poly(t-BF-a-o) 344 379 52.9 Current work
Poly(t-BF-a-n) 336 377 48.0 Current work
Poly(B-bbf) 329 351 36.3 [36]
Poly(B-obf) 316 331 18.5 [36]
Poly(BABPF-p) 380 399 47.0 [31]

temperature is shown in Fig. S3.


From Fig. S3 plots, we can easily observe that all the curves of AMFB
monomers, even on different scanning rate; show an “S” shape, sug-
gesting that the curing reaction of the monomer is an autocatalytic
reaction. In the initial curing reaction stage, the conversion rate
changed slowly with the rise in the temperature, latera sharp en-
hancement in the conversion rate was observed as the temperature of
the system was increased to higher temperatures. This forms the cross-
linking network in the polymer structure and increases the viscosity,
which reduced the molecular movement and the collision between
functional groups became extremely difficult. After a certain extent of
conversion rate, the curing reaction gradually changes from chemical
control to diffusion control, and the conversion curves become flatter.
The corresponding temperature values of α (0.05, 0.10, 0.15,
0.20……0.90, 0.95) at different heating rates can be obtained from the
relationship between α and T of AMFB monomers. The fitting curve of
ln(β/Tf1.92) and 1/Tf at different heating rates are devised by Starink
equation. Fig. S4 illustrates the linear relationship of ln(β/Tf1.92) versus
1/Tf with a slope of −1.0008(Ea/R). The value of activation energy can
be calculated from the slope. The relationship between Ea and α is
shown in Fig. S5.
Fig. 7. Dynamic DSC thermograms of t-BF-a-p (A), t-BF-a-o (B) and t-BF-a-n (C)
From the Ea v/s α plot, we can easily observe that the Ea of t-BF-a-p
monomers at different heating rates.
monomer decreases with the increase of conversion at the initial

8
T. Wang, et al. Reactive and Functional Polymers 143 (2019) 104344

reaction stage (α < 0.2), mainly due to the viscosity of monomer was Acknowledgments
reduced as the temperature was increased. However, the behaviors oft-
BF-a-o and t-BF-a-n monomers were opposite to the behavior oft-BF-a-p This work is financially sponsored by the National Natural Science
monomer. This is due to the fact that these two kinds of monomers were Foundation of China (Project No.51773048), Natural Science
in the complete melting state before polymerization, and their viscos- Foundation of Heilongjiang Province (Project No.E2017022), and
ities were lowest as discussed earlier in the rheological properties sec- Fundamental Research Funds for the Central Universities (Project
tion. When the polymerization starts, the system viscosities increase No.HEUCFP201724).
suddenly due to the gelation. At the intermediate reaction stage, the Ea
of all the curing systems changes little. On α > 0.6, the Ea of all the Appendix A. Supplementary data
systems increases with the increase of conversion. This is due to the
formation of large three-dimensional network structures after the ox- Supplementary data to this article can be found online at https://
azine ring-opening, which reduced the functional groups and chain doi.org/10.1016/j.reactfunctpolym.2019.104344.
segments activities. The polymerization process gradually changes from
chemical control to diffusion control, which obviously increases the Ea. References

[1] C.P.R. Nair, Advances in addition-cure phenolic resins, Prog. Polym. Sci. 29 (5)
3.5. Thermal stabilities (2004) 401–498.
[2] X. Ning, H. Ishida, Phenolic materials via ring-opening polymerization: synthesis
and characterization of bisphenol-A based benzoxazines and their polymers, J.
The thermogravimetric curves of AMFB polymers are plotted in Polym. Sci. A Polym. Chem. 32 (6) (1994) 1121–1129.
Fig. 8. The extracted data from the TGA plot, such as the 5 and 10 [3] X. Ning, H. Ishida, Phenolic materials via ring-opening polymerization of benzox-
azines effect of molecular structure on mechanical and dynamic mechanical prop-
weight percentage loss temperatures (T5 and T10) and carbon residue erties, J. Polym. Sci. B Polym. Phys. 32 (1994) 921–927.
rate (Yc) at 800 °C, are summarized in Table 3. [4] Y.P. Chen, A.Q. Dayo, H.Y. Zhang, A.R. Wang, J. Wang, W.B. Liu, Y. Yang, Q.R. Qin,
The evaluation of TGA curves showed that the initial thermal de- Y.G. Yang, Synthesis of cardanol-based phthalonitrile monomer and its copoly-
merization with phenol–aniline-based benzoxazine, J. Appl. Polym. Sci. 136 (20)
composition temperatures (T5 and T10) of AMFB polymers were in the
(2019) 47505.
range of 336–360 and 377–405 °C, respectively, and the Yc were re- [5] A.Q. Dayo, S. Ullah, S. Kiran, J. Wang, A.H. Shah, A. Zegaoui, Y.B. Arse, W.B. Liu,
corded in 34.9–58.1% range. These thermal stability results are sig- Tensile and water absorption behaviour of polybenzoxazine / hemp fibres compo-
sites: experimental analysis and theoretical validation, Dig. J. Nanomater. Biostruct.
nificantly higher than that of bisphenol A-aniline based poly-
14 (1) (2019) 231–241.
benzoxazine resin (i.e., T5: 277 °C, T10: 317 °C, and Yc: 33.2%) [35]. [6] Y.L. Kobzar, I.M. Tkachenko, V.N. Bliznyuk, V.V. Shevchenko, Fluorinated poly-
These enhancements in the thermal stability values are due to the in- benzoxazines as advanced phenolic resins for leading-edge applications, React.
troduction of heat resistance fluorene ring in the polymers. The increase Funct. Polym. 133 (2018) 71–92.
[7] A.R. Wang, A.Q. Dayo, L.W. Zu, Y.L. Xu, D. Lv, S. Song, T. Tang, W.B. Liu, J. Wang,
in the number of functional sights in the monomer can improve the B.C. Gao, Bio-based phthalonitrile compounds: synthesis, curing behavior, ther-
cross-linking and effectively reduce the share of unstable end groups in momechanical and thermal properties, React. Funct. Polym. 127 (2018) 1–9.
polymers. [8] K. Zhang, L. Han, P. Froirnowicz, H. Ishida, Synthesis, polymerization kinetics and
thermal properties of para-methylol functional benzoxazine, React. Funct. Polym.
The T5, and T10 thermal parameters of the poly(t-BF-a-p) were best 129 (2018) 23–28.
among the studied polybenzoxazine thermosets and recorded as 347 [9] B. Kiskan, Adapting benzoxazine chemistry for unconventional applications, React.
and 395 °C, respectively. These values are higher than the corre- Funct. Polym. 129 (2018) 76–88.
[10] A.Q. Dayo, A. Zegaoui, A.A. Nizamani, S. Kiran, J. Wang, M. Derradji, W.A. Cai,
sponding T5 and T10 values of bisphenol fluorene-aniline based ben- W.B. Liu, The influence of different chemical treatments on the hemp fiber/poly-
zoxazine (poly(B-bbf) and poly(B-obf)) [36], and lower than the dia- benzoxazine based green composites: mechanical, thermal and water absorption
mine-fluorene based benzoxazine resin (poly(BABPF-p)) [31], however, properties, Mater. Chem. Phys. 217 (2018) 270–277.
[11] V. García-Martínez, M.R. Gude, A. Ureña, Understanding the curing kinetics and
the Yc value was much higher. These enhancements can be dedicated to rheological behaviour of a new benzoxazine resin for carbon fibre composites,
the increased number of the oxazine ring, which significantly improved React. Funct. Polym. 129 (2018) 103–110.
the crosslinking density of polybenzoxazine thermosets. However, the [12] R.P. Subrayan, F.N. Jones, Condensation of substituted phenols with hexakis
(methoxymethyl)melamine: synthesis, characterization, and properties of sub-
excessive introduction leads to the generation of more CH2eNeCH2
stituted 2,4,6-Tris[3,4-dihydro-1,3-(2H)-benzoxazin-3-yl]-s-triazine derivatives,
flexible segments in the network structures, which resulting in the de- Chem. Mater. 10 (11) (1998) 3506–3512.
cline of initial thermal decomposition temperatures [31,36]. A large [13] D.X. Wang, B. Li, Y.H. Zhang, Z.J. Lu, Triazine-containing benzoxazine and its high-
number of aromatic rings produce a higher value of carbon residue. performance polymer, J. Appl. Polym. Sci. 127 (1) (2013) 516–522.
[14] B. Kiskan, A.L. Demirel, O. Kamer, Y. Yagci, Synthesis and characterization of na-
nomagnetite thermosets based on benzoxazines, J. Polym. Sci. A Polym. Chem. 46
(20) (2008) 6780–6788.
4. Conclusions [15] X. Wu, et al., Well-defined organic–inorganic hybrid benzoxazine monomers based
on cyclotriphosphazene: synthesis, properties of the monomers and poly-
benzoxazines, Polymer 52 (19) (2011) 4235–4245.
In this current study, three different fluorine-based benzoxazine [16] X. Wu, S.Z. Liu, D.T. Tian, J.J. Qiu, C.M. Liu, Highly branched benzoxazine
monomers were prepared with different aliphatic chains lengths and monomer based on cyclotriphosphazene: synthesis and properties of the monomer
and polybenzoxazines, Polymer 52 (4) (2011) 1004–1012.
substitution sites. The effects of aliphatic chains lengths and substitu- [17] W. Liu, Q.H. Qiu, J. Wang, Z.C. Huo, H. Sun, Curing kinetics and properties of epoxy
tion sites on the curing kinetics, behavior, and rheological and thermal resin–fluorenyl diamine systems, Polymer 49 (20) (2008) 4399–4405.
properties were studied. The synthesized AMFB benzoxazine monomers [18] T.A. Reddy, M. Srinivasan, Preparation and properties of cardopolyimides con-
taining phenoxaphosphine units, J. Polym. Sci. A Polym. Chem. 27 (4) (1989)
have higher softening point temperature and narrower processing 1419–1424.
window. The activation energies calculated by Kissinger and Ozawa [19] S.C. Lin, E.M. Pearce, Epoxy resins. II. The preparation, characterization, and curing
methods showed that the value was significantly reduced as the alkyl of epoxy resins and their copolymers, J. Polym. Sci.: Polym. Chem. Ed. 17 (10)
(1979) 3095–3119.
chain length was increased, and recorded in the range of 108.1–145.4
[20] S.-H. Hsiao, C.-T. Li, Synthesis and characterization of new fluorene-based poly
and 110.8–146.5 kJ/mol, respectively. The glass transition tempera- (ether imide)s, J. Polym. Sci. A Polym. Chem. 37 (10) (1999) 1403–1412.
tures of the synthesized thermosets were higher than the bifunctional [21] P.R. Srinivasan, V. Mahadevan, M. Srinivasan, Preparation and properties of some
fluorene-based benzoxazines. All the AMFB monomers have excellent cardopolyamides, J. Polym. Sci.: Polym. Chem. Ed. 19 (9) (1981) 2275–2285.
[22] Y.L. Liu, C.Y. Chang, C.Y. Hsu, M.C. Tseng, C.I. Chou, Preparation, characterization,
thermal stability. The initial thermal decomposition temperature and and properties of fluorene-containing benzoxazine and its corresponding cross-
char yield at 800 °C are higher than those of bisphenol A-aniline and linked polymer, J. Polym. Sci. A Polym. Chem. 48 (18) (2010) 4020–4026.
bisphenol fluorenyl polybenzoxazine resins. [23] Y. Lu, M.M. Li, Y.J. Zhang, D. Hu, L.L. Me, W.J. Xu, Synthesis and curing kinetics of
benzoxazine containing fluorene and furan groups, Thermochim. Acta 515 (1–2)

9
T. Wang, et al. Reactive and Functional Polymers 143 (2019) 104344

(2011) 32–37. [31] X.Y. He, et al., Synthesis, thermal properties and curing kinetics of fluorene dia-
[24] H.C. Chang, C.H. Lin, Y.W. Tian, Y.R. Feng, L.H. Chan, Synthesis of 9,9-bis(4- mine-based benzoxazine containing ester groups, Eur. Polym. J. 49 (9) (2013)
aminophenyl)fluorene-based benzoxazine and properties of its high-performance 2759–2768.
thermoset, J. Polym. Sci. A Polym. Chem. 50 (11) (2012) 2201–2210. [32] X.Y. He, et al., Investigation of synthesis, thermal properties and curing kinetics of
[25] J. Wang, X.Y. He, J.T. Liu, W.B. Liu, L. Yang, Investigation of the polymerization fluorene diamine-based benzoxazine by using two curing kinetic methods,
behavior and regioselectivity of fluorene diamine-based benzoxazines, Macromol. Thermochim. Acta 564 (2013) 51–58.
Chem. Phys. 214 (5) (2013) 617–628. [33] P. Kasemsiri, S. Hiziroglu, S. Rimdusit, Effect of cashew nut shell liquid on gelation,
[26] J. Wang, T.T. Ren, Y.D. Wang, X.Y. He, W.B. Liu, X.D. Shen, Synthesis, curing be- cure kinetics, and thermomechanical properties of benzoxazine resin, Thermochim.
havior and thermal properties of fluorene-containing benzoxazines based on linear Acta 520 (1–2) (2011) 84–92.
and branched butylamines, React. Funct. Polym. 74 (0) (2014) 22–30. [34] M.J. Starink, The determination of activation energy from linear heating rate ex-
[27] H. Wang, J. Wang, X.Y. He, T.T. Feng, N. Ramdani, M.J. Luan, W.B. Liu, X.D. Xu, periments: a comparison of the accuracy of isoconversion methods, Thermochim.
Synthesis of novel furan-containing tetrafunctional fluorene-based benzoxazine Acta 404 (1) (2003) 163–176.
monomer and its high performance thermoset, RSC Adv. 4 (110) (2014) [35] H.A. Ghouti, A. Zegaoui, M. Derradji, W.A. Cai, J. Wang, W.B. Liu, A.Q. Dayo,
64798–64801. Multifunctional hybrid composites with enhanced mechanical and thermal prop-
[28] J. Dunkers, H. Ishida, Vibrational assignments of N, N-bis (3, 5-dimethyl-2-hydro- erties based on polybenzoxazine and chopped kevlar/carbon hybrid fibers,
xybenzyl)methylamine in the fingerprint region, Spectrochim. Acta 51 (5) (1995) Polymers 10 (2018) 1308.
855–867. [36] J. Wang, M.Q. Wu, W.B. Liu, S.W. Yang, J.W. Bai, Q.Q. Ding, Y. Li, Synthesis, curing
[29] S. Rimdusit, P. Kunopast, I. Dueramae, Thermomechanical properties of arylamine- behavior and thermal properties of fluorene containing benzoxazines, Eur. Polym.
based benzoxazine resins alloyed with epoxy resin, Polym. Eng. Sci. 51 (9) (2011) J. 46 (5) (2010) 1024–1031.
1797–1807. [37] A. Jabbarzadeh, J.D. Atkinson, R.I. Tanner, Effect of molecular shape on rheological
[30] S. Rimdusit, W. Bangsen, P. Kasemsiri, Chemorheology and thermomechanical properties in molecular dynamics simulation of star, H, comb, and linear polymer
characteristics of benzoxazine-urethane copolymers, J. Appl. Polym. Sci. 121 (6) melts, Macromolecules 36 (2003) 5020–5031.
(2011) 3669–3678.

10

You might also like