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

jwp00 | ACSJCA | JCA11.1.4300/W Library-x64 | research.3f (R4.0.i9 HF05:4883 | 2.

1) 2018/07/18 12:44:00 | PROD-WS-116 | rq_3716194 | 9/17/2018 09:58:42 | 12 | JCA-DEFAULT

Article
pubs.acs.org/Macromolecules

1 Benzoxazine Atropisomers: Intrinsic Atropisomerization Mechanism


2 and Conversion to High Performance Thermosets

3 Kan Zhang,† Zhikun Shang,† Corey J. Evans,‡ Lu Han,§ Hatsuo Ishida,*,§ and Shengfu Yang*,‡

4 School of Materials Science and Engineering, Jiangsu University, Zhenjiang 212013, China

5 Department of Chemistry, University of Leicester, Leicester LE1 7RH, United Kingdom
§
6 Department of Macromolecular Science and Engineering, Case Western Reserve University, Cleveland, Ohio 44106, United States
7 *
S Supporting Information

8 ABSTRACT: Atropisomers have inspired chemists and biologists for decades due to their chiral structures and associated
9 biological properties. However, most of atropisomers reported so far arise in highly substituted biaryls and related compounds,
10 and other types have been rarely observed. Here we report a new type of atropisomerism in ortho-tetrahydrophthalimide
11 functional 1,3-benzoxazine family, where the atropisomerism is evident from NMR spectra, with the branching ratio of the
12 atropisomeric configurations invariant with the measurement temperatures. Density functional theory calculations suggested
13 that the reaction intermediate, ortho-tetrahydrophthalimide phenol, is key to the atropisomerism, which creates a large energy
14 barrier after deprotonation and thus determines the branching ratios of the benzoxazine atropisomers. In addition, the ring-
15 opening polymerization of benzoxazine atropisomers has also been investigated. The benzoxazine atropisomers bearing
16 acetylene exhibit unexpectedly low polymerization temperature in the absence of catalysts, suggesting a self-catalyzed
17 polymerization process. Despite the absence of antiflammable additives, the corresponding polybenzoxazine deriving from
18 benzoxazine atropisomers containing acetylene shows exceptionally low heat release capacity (67.2 J g1−K−1) and excellent char
19 residue value (62%). With this work we demonstrate atropisomerism in the 1,3-benzoxazine family for the first time, and
20 provide molecular-level insights to the mechanism, which can open up possibilities for new applications of atropisomers
21 spanning from the microelectronic to the aerospace industries.

22

23
■ INTRODUCTION
Axial chirality, as a stereogenic element in rotationally hindered
are strongly dependent on the temperature, the steric demand
of the substituents, and the length and rigidity of bridges. In
34
35

24 biaryl compounds, has been extensively investigated and addition to a merely physical rotation about the axis, the 36

25 reviewed.1−6 This special type of stereochemistry featured by atropisomerism in biaryls can also result from the photo- 37

26 hindered rotation is termed as atropisomerism,7 where the chemical or chemically induced processes.2 However, so far the 38

27 restricted rotation about a single chemical bond results in molecular-level interpretation for the atropisomerization 39

28 stereoisomers. Since the first observation of atropisomerism in mechanism in biaryls is still lacking, which poses a great 40

29 1922,8 natural products equipped with a rotationally hindered challenge to developing atropselective approach to single 41

30 biaryl axis have been found to be widespread and structurally isomers. Moreover, atropisomerism has rarely been reported in 42

31 diverse, which can exhibit unique bioactivities9,10 and can


32 enable biological stereoselectivity for molecular targets.11,12 In Received: September 6, 2018
33 general, the atropisomerism in axially chiral biaryl compounds

© XXXX American Chemical Society A DOI: 10.1021/acs.macromol.8b01924


Macromolecules XXXX, XXX, XXX−XXX
Macromolecules Article

43 other types of chemical structures beyond biaryls and related magic-angle spinning (CP-MAS) solid-state 13C NMR spectra were 106
44 compounds.13,14 collected using a Bruker Avance III 400 MHz instrument, using 107
45 Benzoxazine has been widely investigated and reported due adamantance as reference. Powder samples were packed into a 4 mm 108
46 to its outstanding properties for applications.15−21 Benzoxazine zirconia rotor with a spinning rate of 10 000 Hz, and the number of 109
transients was 6000 scans. Fourier transform infrared (FT-IR)
47 resin is known to have the highest molecular-design flexibility 110
spectrophotometer was used to acquire FT-IR spectra. The 111
48 among all the polymeric materials and can be tailored to spectrometer was equipped with a deuterated triglycine sulfate 112
49 desired properties. Besides, some unique stereochemistry can (DTGS) detector and dry air purge unit and was operated at a 113
50 also be observed in benzoxazine monomers.22−24 1,3- resolution of 4 cm−1 in the frequency range of 4000−400 cm−1. 114
51 Benzoxazine compounds consisting of the ortho-substituted Elemental analysis was characterized by an elementary analyzer 115
52 phenolic components possess significant advantages over the (Elementar Vario EL-III). A TA Instruments differential scanning 116
53 para-phenolic counterparts in terms of monomer synthesis, as calorimeter (DSC) model 2920 was used with a temperature ramp 117
54 well as the mechanical and physical properties of the rate of 10 °C/min and a nitrogen flow rate of 60 mL/min for all tests 118
55 corresponding thermosets, due to intramolecular hydrogen of DSC study. All samples were sealed in hermetic aluminum pans 119
56 bonding in phenolic precursor and further reactions of the with lids. Viscosity was tested by Anton Paar GmbH Rheometer MCR 120
57 ortho-substituted phenolic structure after ring-opening poly- 301 with constant shear rate at 10 s−1 with a temperature ramp rate of 121

58 merization. The superior properties of the ortho-benzoxazines 5 °C/min. Thermogravimetric analyses (TGA) were performed on a 122
TA Instruments Q500 thermogravimetric analyzer that was purged 123
59 to their para-substituted counterparts were unexpected since with nitrogen at a flow rate of 40 mL/min. A heating rate of 10 °C/ 124
60 all nonbenzoxazine polymers reported to data exhibit the min was applied. Dynamic mechanical analysis (DMA) was 125
61 opposite trend.25−30 conducted on a TA Instruments Model Q800 DMA applying 126
62 Ortho-imide substituted 1,3-benzoxazine molecules contain controlled strain tension mode with amplitude of 10 μm and a 127
63 three oxygen atoms in the oxazine ring and imide temperature ramp rate of 3 °C/min. Strain sweep was first performed 128
64 functionalities, which are expected to be negatively charged. to determine the linear viscoelastic limit. Specific heat release rate 129
65 If the two groups rotate about the C−N bond so the oxygen (HRR, W/g), heat release capacity (HRC, J/(gK)) and total heat 130
66 atoms approach each other, then strong ionic repulsion will release (THR, kJ/g) were carried out on a microscale combustion 131
67 occur, which may hinder the rotation and thus deliver calorimeter (MCC) from PHINIX. MCC was measured from 100 to 132
68 atropisomerism for these molecules. To prove the concept, 900 °C at a heating rate of 1 °C/s in an 80 cm3/min stream of 133

69 in this work we synthesized a few ortho-tetrahydrophthalimide nitrogen. The anaerobic thermal degradation products in the nitrogen 134
gas stream were mixed with a 20 cm3/min stream of oxygen prior to 135
70 substituted 1,3-benzoxazines and investigated the chemical
entering the combustion furnace (900 °C). Heat release is quantified 136
71 structures using NMR techniques, including correlated spec- by standard oxygen consumption,31 and HRR is obtained by dividing 137
72 troscopy (COSY) and heteronuclear multiple quantum dQ/dt at each time interval, by the initial sample mass, and HRC is 138
73 coherence spectroscopy (HMQC). Density functional theory obtained by dividing the maximum value of HRR by the heating rate. 139
74 (DFT) calculations were then performed to interpret the Computational Methods. Computational chemistry calculations 140
75 experimental observations. In addition, we also investigated the were carried out using both the Gaussian03 and CASTEP suite of 141
76 ring-opening polymerization of this class of benzoxazine programs. For the G03 calculations the B3LYP hybrid function was 142
77 atropisomers. To our knowledge, there are no previous reports employed using a 6-31+G(d) basis set, with tight SCF convergence 143
78 on the polymerization of atropisomers, neither on the and an ultrafine pruning grid.32 Structures were then optimized and 144
79 properties of their resulting polymers. Thus, another purpose the harmonic vibrational frequencies determined to show that a true 145
80 of this paper is to develop novel thermosets based on structural minima had been found. For the 13C NMR calculation step 146
the B3LYP/6-31+G(d) optimized structures were used but the 147
81 benzoxazine atropisomers, aiming to achieve very high
shielding tensors were calculated at the MPW1PW91/6-311+G(2d,p) 148
82 performance thermosets with high thermally stability and low level of theory which has been found to give accurate values of the 13C 149
83 flammability from mono-oxazine benzoxazines, despite the chemical shift.33,34 The infrared spectrum were generated by 150
84 majority of mono-oxazine benzoxazines cannot exhibit performing frequency calculations. 151
85 excellent thermal properties due to their inability to form Synthesis of 2-(2-Hydroxyphenyl)-3a,4,7,7a-tetrahydro-1H-iso- 152
86 large molecular weight polymers which is caused by indole-1,3(2H)-dione (Abbreviated as oHTI). Into a 250 mL round 153
87 termination of the propagating species by the competing side flask were placed cis-1,2,3,6-tetrahydrophthalic anhydride (4.92 g, 154
88 reactions. 0.03 mol), o-aminophenol (3.27 g, 0.03 mol), and 80 mL of acetic 155


acid. The mixture was stirred and refluxed for 6 h. After cooling to 156
room temperature, the precipitate was filtered and washed with 500 157
89 EXPERIMENTAL SECTION mL of deionized water. Removal of water by evaporation afforded a 158
90 Materials. o-Aminophenol (>98%), p-aminophenol (>98%), cis- white powder. The resulting white product was recrystallized from 159
91 1,2,3,6-tetrahydrophthalic anhydride, 4-chloroaniline (98%), 4- isopropanol. (yield ca. 88%). 1H NMR (400 MHz, DMSO-d6), ppm: 160
92 ethynylaniline (98%), 4-aminobenzonitrile (98%), and paraformalde- δ = 2.26−2.29 (d, 2H, −CH2−), 2.42−2.45 (d, 2H, −CH2−), 3.29− 161
93 hyde (99%) were used as-received from Sigma-Aldrich. Aniline was 3.40 (d, 2H), 5.94 (s, 2H, C−H), 6.83−7.27 (4H, Ar), 9.74 (s, 1H, 162
94 purchased from Aldrich and purified by distillation. Acetic acid, −OH). Anal. Calcd. for C14H13NO3: C, 69.12; H, 5.39; N, 5.76. 163
95 hexanes, and xylenes were obtained from Fisher Scientific and used as- Found: C, 69.08; H, 5.42; N, 5.75. 164
96 received. Synthesis of 2-(3-Phenyl-3,4-dihydro-2H-benzo[e][1,3]oxazin-8- 165
97 Characterization. 1D 1H NMR spectra were recorded on a yl)-3a,4,7,7a-tetrahydro-1H-isoindole-1,3(2H)-dione (Abbreviated 166
98 Bruker AVANCE II 400 MHz nuclear magnetic resonance (NMR) as oHTI-a). Into a 100 mL round flask were added 30 mL of xylenes, 167
99 spectrometer at a proton frequency of 400 MHz in CDCl3/DMSO-d6 aniline (0.93 g, 0.01 mol), oHTI (2.55 g, 0.01 mol), and 168
100 using tetramethyl silane as internal standard. 2D 1H−1H Correlated paraformaldehyde (0.61 g, 0.02 mol). The mixture was stirred at 169
101 spectroscopy (COSY) and 1H−13C Heteronuclear multiple quantum 120 °C for 6 h. The mixture was cooled to room temperature and 170
102 coherence (HMQC) were also examined. In order to elucidate the white crystal was obtained. (yield ca. 96%). 1H NMR (CDCl3), ppm: 171
103 difference in isomerism, NMR spectra were also acquired using a δ = 2.31−2.43 (m, 2H, −CH2−), 2.68−2.74 (m, 2H, −CH2−), 3.24− 172
104 Varian Oxford AS300 (300 MHz), a Bruker Ascend III 500 MHz, and 3.32 (m, 2H), 4.64−4.66 (d, 2H, Ar−CH2−N, oxazine), 5.31−5.35 173
105 a Varian Inova 600 MHz NMR sepctrometers. Cross-polarization (d, 2H, O−CH2−N, oxazine), 6.00−6.06 (m, 2H, C−H), 6.91− 174

B DOI: 10.1021/acs.macromol.8b01924
Macromolecules XXXX, XXX, XXX−XXX
Macromolecules Article

Scheme 1. Synthesis of ortho-Tetrahydrophthalimide Benzoxazine Monomers

Figure 1. HMQC NMR spectra of oHTI-a in CDCl3. The expanded view shows that the oxazine protons of oHTI-a are correlated to four different
carbon resonances.

175 7.31 (8H, Ar). Anal. Calcd. for C22H20N2O3: C, 73.32; H, 5.59; N, Synthesis of 4-(8-(1,3-Dioxo-3a,4-dihydro-1H-isoindol-2- 202
176 7.77. Found: C, 73.29%; H, 5.61%; N, 7.75%. (3H,7H,7aH)-yl)-2H-benzo[e][1,3]oxazin-3(4H)-yl)benzonitrile (Ab- 203
177 Synthesis of 2-(3-(4-Ethynylphenyl)-3,4-dihydro-2H-benzo[e]- breviated as oHTI-cy). Into a 100 mL round flask were added 40 204
178 [1,3]oxazin-8-yl)-3a,4,7,7a-tetrahydro-1H-isoindole-1,3(2H)-dione mL of xylenes, 4-aminobenzonitrile (1.18 g, 0.01 mol), oHTI (2.55 g, 205
179 (Abbreviated as oHTI-ac). Into a 100 mL round flask were added 40 0.01 mol), and paraformaldehyde (0.60 g, 0.02 mol). The mixture was 206
180 mL of xylenes, 4-ethynylaniline (1.17 g, 0.01 mol), oHTI (2.43 g, 0.01 stirred at 120 °C for 12 h. The mixture was cooled to room 207
181 mol), and paraformaldehyde (0.60 g, 0.02 mol). The mixture was temperature and white crystal was obtained. (yield ca. 91%). 1H NMR 208
182 stirred at 120 °C for 8 h. The mixture was cooled to room (CDCl3), ppm: δ = 2.28−2.43 (m, 2H, −CH2−), 2.65−2.74 (m, 2H, 209
183 temperature and white crystal was obtained. (yield ca. 92%). 1H NMR −CH2−), 3.24−3.35 (m, 2H), 4.69−4.71 (d, 2H, Ar−CH2−N, 210
184 (CDCl3), ppm: δ = 2.26−2.43 (m, 2H, −CH2−), 2.63−2.75 (m, 2H, oxazine), 5.32−5.36 (d, 2H, O−CH2−N, oxazine), 5.99−6.03 (m, 211
185 −CH2−), 3.01 (d, 1H, CH), 3.23−3.35 (m, 2H), 4.65−4.67 (d, 2H, C−H), 6.93−7.58 (7H, Ar). Anal. Calcd. for C23H19N3O3: C, 212
186 2H, Ar−CH2−N, oxazine), 5.30−5.34 (d, 2H, O−CH2−N, oxazine), 71.67; H, 4.97; N, 10.90. Found: C, 71.61%; H, 5.01%; N, 10.87%. 213
187 5.99−6.03 (m, 2H, C−H), 6.91−7.43 (7H, Ar). Anal. Calcd. for Synthesis of 3-(4-Ethynylphenyl)-3,4-dihydro-2H-benzo[e][1,3]- 214
188 C24H20N2O3: C, 74.98; H, 5.24; N, 7.29. Found: C, 74.92%; H, oxazine (Abbreviated as Ph-ac). Into a 100 mL round flask were 215
189 5.26%; N, 7.24%. added 30 mL of toluene, 4-ethynylaniline (1.17 g, 0.01 mol), phenol 216
190 Synthesis of 2-(3-(4-Chlorophenyl)-3,4-dihydro-2H-benzo[e]- (0.94 g, 0.01 mol), and paraformaldehyde (0.60 g, 0.02 mol). The 217
191 [1,3]oxazin-8-yl)-3a,4,7,7a-tetrahydro-1H-isoindole-1,3(2H)-dione mixture was stirred at 100 °C for 8 h. Column chromatography was 218
192 (Abbreviated as oHTI-ch). Into a 100 mL round flask were added 40 used afterward to purify benzoxazine monomer with mixed solvents of 219
193 mL of xylenes, 4-chloroaniline (1.28 g, 0.01 mol), oHTI (2.43 g, 0.01 hexane and ethyl acetate in 4:1 ratio. The purified products were 220
194 mol), and paraformaldehyde (0.60 g, 0.02 mol). The mixture was further recrystallized from acetone/toluene mixtures (1:1). After 221
195 stirred at 120 °C for 6 h. The mixture was cooled to room recrystallization, needle like crystals were obtained. (yield ca. 38%). 222
196 temperature and white crystal was obtained. (yield ca. 96%). 1H NMR 1
H NMR (CDCl3), ppm: δ = 3.01 (d, 1H, CH), 4.67 (s, 2H, Ar− 223
197 (CDCl3), ppm: δ = 2.30−2.43 (m, 2H, −CH2−), 2.66−2.75 (m, 2H, CH2−N, oxazine), 5.38 (s, 2H, O−CH2−N, oxazine), 6.83−7.43 224
198 −CH2−), 3.23−3.34 (m, 2H), 4.60−4.62 (d, 2H, Ar−CH2−N, (8H, Ar). Anal. Calcd. for C16H13NO: C, 81.68; H, 5.57; N, 5.95. 225
199 oxazine), 5.27−5.31 (d, 2H, O−CH2−N, oxazine), 5.97−6.04 (m, Found: C, 81.61%; H, 5.59%; N, 5.94%. 226
200 2H, C−H), 6.92−7.31 (7H, Ar). Anal. Calcd. for C22H19ClN2O3: Polymerization of Benzoxazine Monomers. Each benzoxazine 227
201 C, 66.92; H, 4.85; N, 7.09. Found: C, 66.89%; H, 4.88%; N, 7.08%. monomer was polymerized using different polymerization cycles. 228

C DOI: 10.1021/acs.macromol.8b01924
Macromolecules XXXX, XXX, XXX−XXX
Macromolecules Article

229 Polymerization of oHTI-ac was carried out by heating at 160 °C, 180
230 °C, 200 °C, and 220 °C for 2 h each step, thus obtaining poly(oHTI-
231 ac). Polymerizations of oHTI-a and oHTI-ch were carried out by
232 heating at 160 °C, 180 °C, 200 °C, 220 °C, and 240 °C for 2 h each
233 step, thus obtaining poly(oHTI-a) and poly(oHTI-ch). In addition,
234 polymerization of oHTI-cy was done by heating at 160 °C, 180 °C,
235 200 °C, 220 °C, 240 °C, and 260 °C for 2 h each step, thus forming
236 poly(oHTI-cy).

237

238
■ RESULTS AND DISCUSSION
Synthesis and Atropisomerism Analysis of Benzox-
239 azine Atropisomers. 1,3-Benzoxazine bearing tetrahydroph-
240 thalimide group as ortho-functionality, 2-(3-phenyl-3,4-dihy-
241 dro-2H-benzo[e][1,3]oxazin-8-yl)-3a,4,7,7a-tetrahydro-1H-iso-
s1 242 indole-1,3(2H)-dione, denoted as oHTI-a (Scheme 1), was
243 synthesized using primary amine (aniline), formaldehyde, and
244 ortho-tetrahydrophthalimide functional phenol (oHTI). 1,3-
245 Benzoxazine compounds are expected to have two equal
246 intensity singlet resonances in 1H NMR spectra due to the two
f1 247 methylenes in the oxazine ring.35,36 However, as seen in Figure
f1 248 1, the characteristic proton resonance due to the O−CH2−N
249 and the Ar−CH2−N groups in the oxazine ring of oHTI-a
250 exhibits two sets of doublets at 5.31 and 5.35 ppm and 4.64
251 and 4.66 ppm, respectively. In addition, the oxazine protons of
252 oHTI-a are correlated to four different carbon resonances in
253 the 1H−13C HMQC spectrum, suggesting the distinct
254 subspectra for benzoxazine compounds. All these observations
255 suggest the coexistence of two isomers. The branching ratio
256 can be integrated as 1:1.72 for the two products according to
257 the relative intensities. Further measurements were performed
258 at different temperatures (15 °C, 20 °C, 25 °C, 30 °C, and 35
259 °C) and at different frequencies (300, 400, 500, and 600 MHz)
260 but no change in chemical shifts, line widths, J-couplings,
261 integration values of each proton in both products and the
262 branching ratio was observed (Figure S1 of the Supporting
263 Information, SI). Furthermore, the ratio persists over a period
264 of three months and even after multiple rounds of drying or
265 thermal treatments in vacuo (30 °C, 60 °C, 90 °C, and 120
266 °C). These observations confirm that the two structures are
267 not conformers under thermal equilibrium, but are stereo-
268 chemical isomers. Figure 2. Optimized stuctures and 13C NMR spectra of oHTI-a. (a)
269 To provide molecular level interpretation to the isomerism Optimized stuctures of oHTI-a: oHTI-a-A (left) and oHTI-a-B (right)
270 in 1,3-benzoxazine, DFT calculations were performed using by DFT calculations. (b) Comparison between the experimental and
calculated 13C NMR spectra of oHTI-a. The NMR spectra are first
271 B3LYP/6-31+G(d) method. For oHTI-a, two configurations
calculated for oHTI-a-A and oHTI-a-B, respectively, and the overall
f2 272 were identified (Figure 2a). In both structures, the oxazine ring spectrum is generated by adopting the experimental branching ratio of
273 is perpendicular to the plane of imide group; while the 1:1.72.
274 cyclohexene end-cap is located on either side of oxazine ring,
275 resulting in two different configurations (denoted as oHTI-a-A not determined by the reaction products under thermal 290
276 and oHTI-a-B, respectively). Besides, the interconversion equilibrium. 291
277 barrier between oHTI-a-A and oHTI-a-B was calculated as To unveil the determining factor for the branching ratio of 292
278 high as 106.8 kJ/mol, which makes the two configurations oHTI-a, we performed calculations on a reaction intermediate, 293
279 thermally stable and hinders the rotation about the C−N bond. oHTI, which were found to have four conformations (denoted 294
280 Both experimental and calculated 13C NMR spectra are as oHTI-1, oHTI-2, oHTI-3, and oHTI-4, see Figure 3). oHTI- 295 f3
281 presented in Figure 2b, which show excellent agreement (see 3 and oHTI-4 both have an intramolecular hydrogen bond, 296
t1 282 Table 1 for the full list of chemical shifts) and confirm that the making it difficult for deprotonation. In contrast, the phenolic 297
283 two isomers seen in the NMR spectra are oHTI-a-A and oHTI- OH bond is perpendicular to the plane of imide group in 298
284 a-B, respectively. However, the energetic difference between oHTI-1 and oHTI-2, and can release a proton and form 299
285 the two isomers is 2.4 kJ/mol, dedicating a branching ratio of anionic products (oHTI-1* and oHTI-2*) in solution. The 300
286 1:2.6 between oHTI-a-A and oHTI-a-B at 298 K, notably energy difference between oHTI-1 and oHTI-2 was found to 301
287 different from the experimental ratio of 1:1.72. Together with be 1.84 kJ/mol in benzene, which gives rise to a branching 302
288 the large interconversion energy barrier between the two ratio of 1.75:1 at the reaction temperature of 393 K. Further 303
289 structures, such a deviation indicates that the branching ratio is calculations on oHTI-1* and oHTI-2* and their transition 304

D DOI: 10.1021/acs.macromol.8b01924
Macromolecules XXXX, XXX, XXX−XXX
Macromolecules Article

13
Table 1. Experimental and Calculated C Chemical Shifts reaction intermediate, ortho-tetrahydrophthalimide phenol, 329
of Each Atropisomer accounts for the atropisomerism and determines the branching 330
ratio of the benzoxazine atropisomers. Potentially, the 331
atropisomerization mechanism discovered in this work can 332
be used as a generic design principle for improving drug 333
development, organic catalysts, molecular electronics, and 334
other fields wherever atropisomers can play important and 335
prevalent roles. 336
Polymerization Behavior of Benzoxazine Atro- 337
pisomers. The polymerization profiles of benzoxazine 338
atropisomers with various functionalities were studied by 339
DSC as depicted in Figure 4a, and the results are summarized 340 f4
in Table 2. The thermograms show that the exothermic 341 t2
maximum of the ring-opening polymerization for oHTI-a 342
shift (ppm) shift (ppm) centers at 252 °C, while the maxima for oHTI-ac, oHTI-ch, 343
expt. calc. atoms expt. calc. atoms and oHTI-cy are at 214 °C, 261 °C, and 273 °C, respectively. 344
178.93 181.16 5 178.89 181.02 18 The polymerization temperatures of benzoxazine atropisomers 345
178.93 180.34 3 178.89 180.56 21 are shifted due to the difference in electronegativity of 346
149.73 154.60 13 149.73 154.02 10 acetylene, chlorine, and nitrile functionalities. 347
148.36 152.83 22 148.34 152.71 2 oHTI-a exhibits an endothermic peak at 129 °C and a very 348
129.32 132.86 7 129.32 133.41 24 small exothermic peak at 131 °C as shown in Figure S14a, 349
129.32 132.49 8 129.32 133.08 23 followed by a very sharp endothermic peak at 154 °C can be 350
127.75 130.30 24 127.47 130.38 4 observed after the exothermic peak, which is due to the phase 351
127.47 129.91 26 127.84 130.31 15 changes of crystal-to-crystal transition. As shown in the DSC 352
127.47 129.91 17 127.84 130.17 6 profiles of Figure S14b, this phase transition disappears after 353
127.36 127.32 15 127.39 127.40 13 the first scanning by heating to 137 °C at a heating rate of 10 354
122.10 123.62 27 122.21 123.59 1 °C/min, suggesting that the monotropic characteristic between 355
122.10 123.62 10 122.21 123.44 16 these two crystal forms in oHTI-a according to the Burger- 356
121.99 123.17 25 122.06 123.31 5 Ramberger heat-of-fusion rule.37 Besides, the value of the 357
120.37 122.88 14 120.39 122.81 11 exothermic peak of oHTI-a shows no variations as shown in 358
120.27 120.93 23 120.13 120.75 3 Figure S14, suggesting that the crystal phase has no influence 359
118.80 119.72 16 118.94 120.07 14 on the ring-opening polymerization of benzoxazines. 360
79.63 81.94 19 80.01 82.15 8 Generally, one expects an increase on the polymerization 361
50.98 47.27 21 50.71 47.21 12 temperature due to the electron-withdrawing groups in the 362
39.51 40.15 1 39.37 39.73 19 para-position of the phenyl in amine substituent.38 According 363
39.51 39.73 2 39.37 39.43 20 to the electronegativity order −H < −CCH < −Cl < CN, 364
23.67 23.36 9 23.76 23.30 22 the electronegativity in the para-position of the phenyl should 365
23.67 22.96 6 23.76 23.09 25 follow the order oHTI-a < oHTI-ac < oHTI-ch < oHTI-cy but 366
the polymerization temperature for these benzoxazines has a 367
305 state suggested a large energy barrier between the two anionic different order as oHTI-ac < oHTI-a < oHTI-ch < oHTI-cy. 368
306 precursors, i.e., as high as 102.2 and 95.2 kJ/mol for oHTI-1* The values for the heat of benzoxazine monomers are 369
307 and oHTI-2*, respectively, which prevents any further summarized in Table 2. The acetylene containing monomer, 370
308 interexchange between the two structures under thermal oHTI-ac, shows the highest heat of polymerization with a value 371
309 conditions. The emergence of such an energy barrier can be of 563 J/g, and the heat of polymerization were 226, 176, and 372
310 attributed to the repulsion between negatively charged oxygen 186 J/g for oHTI-a, oHTI-ch and oHTI-cy, respectively. The 373
311 atoms in the oxazine ring and imide functionalities when they significantly higher heat of polymerization for oHTI-ac 374
312 approach the transition state structure via rotation, and a large suggests the involvement of other events in the ring-opening 375
313 configurational change will be necessary for rotational switch reaction. Since the presence of acetylene is the unique feature 376
314 about the C−N bond. As a result, no conformational switch via of oHTI-ac, one possibility is the polymerization of acetylene, 377
315 rotation is possible in the follow-on reactions and thus the which takes place in advance of oxazine ring, lowering 378
316 branching ratio maintains. The mechanism for the formation of polymerization temperature of oHTI-ac. To compare, we 379
s2 317 atropisomers of oHTI-a is illustrated in Scheme 2. synthesized another benzoxazine monomer containing acety- 380
318 To further confirm the reaction mechanism, three ortho- lene group, Ph-ac (a benzoxazine monomer derived from 381
319 tetrahydrophthalimide benzoxazine monomers with different phenol and 4-ethynylaniline), which showed the maximum of 382
320 functionalities in amine (Scheme 1) were also designed and polymerization at 226 °C. In contrast, the maximum 383
321 synthesized. Similar to oHTI-a, all of the three ortho- temperature of polymerization for oHTI-a is 38 °C higher 384
322 tetrahydrophthalimide benzoxazines (oHTI-ac, oHTI-ch, and than that of oHTI-ac in absence of acetylene functionality. 385
323 oHTI-cy) have a very high yield (>90%). Although monomers These DSC results indicate the existence of self-catalytic 386
324 of these three molecules have different functionalities in amine, polymerization mechanism possibly caused by the interaction 387
325 the branching ratio between two atropisomers in each between acetylene and benzoxazine when acetylene function- 388
326 monomer was found to be similar to that of oHTI-a. This ality is incorporated into benzoxazine atropisomers. 389
327 strongly supports the intrinsic atropisomerization mechanism In order to qualitatively study the structural evolution of 390
328 in ortho-tetrahydrophthalimide benzoxazines, where the oHTI-ac during heating, in situ FT-IR analyses were carried 391

E DOI: 10.1021/acs.macromol.8b01924
Macromolecules XXXX, XXX, XXX−XXX
Macromolecules Article

Figure 3. (a) 1H NMR spectra of oHTI-a, oHTI-ac, oHTI-ch, and oHTI-cy. (b) The expanded region between 4.5 and 5.5 ppm shows two sets of
doublet resonances belonging to −CH2− of the oxazine ring, where the well-separated doublets of −O−CH2−N− are used to integrate for
determine the branching ratio of atropisomers in each monomer.

Scheme 2. Atropisomeric Mechanism of ortho-Tetrahydrophthalimide Functional Benzoxazinea

a
(a) 2-D (dimension) scheme for the simplified mechanism of benzoxazine formation. (b) 3-D scheme for the atropisomerism mechanism. The
energy barrier between oHTI-1* and oHTI-2* is calculated as 102.2 and 95.2 kJ/mol, respectively, making it difficult for inter-conversion of the
structures.

392 out. As shown in Figure 4c, the characteristic absorption bands 933 cm−1 (benzoxazine related mode) can be used to monitor 394
393 at 1239 cm−1 (C−O−C antisymmetric stretching mode) and the ring-opening polymerization of the benzoxazine resin.39,40 395

F DOI: 10.1021/acs.macromol.8b01924
Macromolecules XXXX, XXX, XXX−XXX
Macromolecules Article

Figure 4. (a) DSC thermograms of oHTI-a, oHTI-ac, oHTI-ch, and oHTI-cy. (b) DSC comparison of oHTI-ac with oHTI-a and Ph-ac reveals the
presence of self-catalyzed polymerization between oxazine ring and acetylene in oHTI-ac. (c) FT-IR spectra of oHTI-ac after various thermal
treatments. (d) FT-IR spectra of Ph-ac after various thermal treatments.

Table 2. DSC Thermograms of Benzoxazine Monomers polymerization in oHTI-ac. This cyclotrimerization is reported 414
to be preceded by an intermediate trimer which is formed at 415
onset temp max temp heat of polymerization
monomer (°C) (°C) (J/g) much lower temperature than the cyclotrimerization temper- 416

oHTI-a 231 252 226


ature. This intermediate structure forces three monomeric 417

oHTI-ac 179 214 563


molecules to form a rigid structure that resembling a 418

oHTI-ch 245 261 176


cyclotrimer. The intermediate structure thus increases the 419

oHTI-cy 246 273 186 likelihood of packing with neighboring cyclotrimer intermedi- 420
ates forcing the oxazine ring in the vicinity of each other. This 421

396 Both bands decrease as the temperature increase, and fully two cyclotrimer intermediates can be formed in the melt, 422

397 disappear at 220 °C. Meanwhile, the OH band in the region of allowing the benzoxazine polymerization to take place at lower 423

398 3500−3200 cm−1 emerges alongside the decrease of both the temperature than the completely random orientation of the 424

399 characteristic bands of the internal arynes at 2100 cm−1 and benzoxazine monomers. Much more detailed study is required 425

400 the C−H stretching vibration of the monosubstituted to verify this hypothesis, though it is not the main goal of the 426

401 acetylene at 3267 cm−1, and the broadening of the absorption current paper. 427

402 band at 1605 cm−1 attributed to the aromatic ring. These The structural conversion from benzoxazine monomer to 428

403 variations are likely related to the benzene ring formation by polybenzoxazine of oHTI-ac was also investigated by solid- 429
404 the cyclotrimerization of acetylenes.41 The above results state 13C NMR with cross-polarization magic-angle spinning 430
405 therefore suggest that the exotherm in the DSC thermogram (CP-MAS). In accordance with the 13C NMR spectrum of 431
406 of oHTI-ac has two undergoing processes, including the oHTI-ac in solution, the characteristic solid 13C resonances of 432
407 polymerization of acetylene and ring-opening polymerization the oxazine ring are assigned at around 47 and 77 ppm for Ar− 433
408 of oxazine group. Extra evidence is available from the complete CH2−N− and −O−CH2−N−, respectively (Figure 5).18 434 f5
409 disappearance of the characteristic absorption bands of oxazine Besides, the characteristic resonances of the acetylene carbons 435
410 ring for Ph-ac at 1239 cm−1 (C−O−C antisymmetric of CH and −C are centered at around 70 and 84 ppm. In 436
411 stretching mode) and 933 cm−1 (benzoxazine related mode) addition, the spectrum for poly(oHTI-ac) showed drastic 437
412 complete the disappearance at a higher temperature of 240 °C differences compared with that of oHTI-ac as expected. The 438
413 (Figure 4d), further suggesting the existence of self-catalyzed carbon resonances of oxazine ring and acetylene of oHTI-ac 439

G DOI: 10.1021/acs.macromol.8b01924
Macromolecules XXXX, XXX, XXX−XXX
Macromolecules Article

Figure 5. Comparison among solution-state 13C NMR spectrum, solid-state 13C NMR spectrum for oHTI-ac, and solid-state CP-MAS 13C NMR
spectrum for poly(oHTI-ac).

Scheme 3. Representation of Self-Catalyzed Polymerization of oHTI-ac

440 completely disappeared, and a broad signal corresponding to in Figure 6a. As expected, degradation in the temperature 455 f6
441 the −CH2− of Mannich bridge around 47 ppm can be range of 250 °C to 400 °C was observed for poly(oHTI-a), 456
442 observed in the spectrum of poly(oHTI-ac). Moreover, a very poly(oHTI-ch), and poly(oHTI-cy), which can be attributed to 457
443 broad peak attributed to the −CH of benzene ring are the partial decomposition of terminal groups as defects of 458
444 formed after polymerization. These spectral changes suggest polybenzoxazines from monofunctional benzoxazines.42 On 459
445 the completion of ring-opening polymerization and cyclo- the contrary, poly(oHTI-ac) exhibits excellent thermal stability 460
446 trimerization for oHTI-ac can be achieved at the temperature
at this temperature range with the highest Td5 (5% weight loss 461
447 as low as 220 °C, and formation of the thermoset with highly
temperature) of 396 °C among this class of polybenzoxazines. 462
448 aromatic cross-linking network. Therefore, the results of solid-
449 state 13C NMR analysis fully agree with in situ FTIR results, The highly aromatic cross-linked networks formed from the 463

450 and we can now propose a structural transformation cyclotrimerization of acetylene along with the polybenzoxazine 464
s3 451 mechanism of oHTI-ac, as being described in Scheme 3. network significantly decrease the fraction of the terminal 465
452 Thermal and Heat Release Properties of Thermosets. defects in poly(oHTI-ac). In addition, poly(oHTI-ac) also 466
453 The thermal stability of poly(oHTI-a), poly(oHTI-ac), poly- showed a char yield of as high as 62%. The TGA data therefore 467
f6 454 (oHTI-ch), and poly(oHTI-cy) were studied by TGA as seen demonstrates the high thermal stability of poly(oHTI-ac). 468

H DOI: 10.1021/acs.macromol.8b01924
Macromolecules XXXX, XXX, XXX−XXX
Macromolecules Article

Figure 6. (a) Thermogravimetric analysis of polybenzoxazines. (b) DSC thermograms of polybenzoxazines recorded under nitrogen at a heating
rate of 10 °C/min. (c) Dynamic mechanical spectra of poly(oHTI-ac). (d) Viscosity versus temperature of oHTI-ac.

469 Figure 6b compares DSC thermograms of polybenzoxazines. which exhibits the advantage of being monofucntional 495
470 In general, the glass transition temperature (T g ) of benzoxazine with very low viscosity. All the result above 496
471 polybenzoxazines based on monofunctional benzoxazine highlights the advantage of the presence of acetylene in 497
472 monomers are much lower than those derived from difunc- benzoxazine atropisomers. 498
473 tional or main-chain type benzoxazines because of the small Microscale combustion calorimetry (MCC) is typically a 499
474 molecular weight and/or low cross-link density of monofunc- constant heating rate pyrolysis test in which the evolved gases 500
475 tional benzoxazines, despite some expected reactivity on the are combusted in excess oxygen and the specific heat release 501
476 aromatic ring of the phenolic and aniline moiety.43 As shown rate (HRR) is determined for a milligram-sized specimen by 502
477 in Figure 6b, the Tg of poly(oHTI-a), poly(oHTI-ch) and oxygen consumption calorimetry. HRR is obtained by dividing 503
478 poly(oHTI-cy) are observed at 233 °C, 198 °C, and 217 °C, dQ/Dt at each time interval, by the initial sample mass. 504
479 which are much lower than that of poly(oHTI-ac) (Tg = 276 Besides, the heat release capacity (HRC) is obtained by 505
480 °C). Moreover, dynamic mechanical analysis (DMA) was dividing the maximum value of HRR by the heating rate. 506
481 utilized to further confirm the Tg value of poly(oHTI-ac). As Generally, HRC is an effective evaluation of thermal 507
482 shown in Figure 6c, the Tg determined by the peak combustion that is one of the best single predictors of the 508
483 temperature of both E″ and tan δ is as high as 298 °C, flammability of a material.44 In addition, HRC is a material 509
484 which is in relatively good agreement with that measured by property that is rooted in the chemical structure of the polymer 510
485 DSC with the expected discrepancy between the two different and can be calculated from additive molar group contributions 511
486 techniques. The significant improvement of Tg for poly(oHTI- for simple materials.44 512
487 ac) is attributed to the formation of highly aromatic cross- Milligram size samples of poly(oHTI-a), poly(oHTI-ac), 513
488 linked networks endowed by cyclotrimerization of acetylene poly(oHTI-ch), and poly(oHTI-cy) were tested at a constant 514
489 along with phenolic Mannich bridge networks. In general, the heating rate of 1 K/s over the temperature range 100−900 °C. 515
490 required viscosity for resin transfer molding (RTM) processing As shown in Figure 7, MCC characterization of poly(oHTI-a), 516 f7
491 technique is roughly below 1 Pa·s at the processing poly(oHTI-ac), poly(oHTI-ch), and poly(oHTI-cy) reveals 517
492 temperature. Interestingly, the viscosity of oHTI-ac is less HRC values of 107.8, 67.2, 109.0, and 113.2 J g1−K−1, 518
493 than 0.1 Pa.s in the temperature range between 169 °C and respectively. Besides, poly(oHTI-a), poly(oHTI-ch), and 519
494 200 °C at a heating rate of 5 °C/min as shown in Figure 6d, poly(oHTI-cy) exhibit THR values of 17.8, 17.6, and 16.4 520

I DOI: 10.1021/acs.macromol.8b01924
Macromolecules XXXX, XXX, XXX−XXX
Macromolecules Article

between experiments and calculations suggests that the 541


intrinsic mechanism for the formation of this new type 542
atropisomerism, i.e., hindering rotation about the C−N bond 543
by intramolecular repulsion between negatively charged 544
oxygen atoms in the oxazine ring and imide functionalities 545
and the consequent structural change needed for rotation. 546
Such detailed structural and mechanistic analysis delivers a 547
general mechanism as well as the basis for better understanding 548
of the factors that are critical for the atropisomerization in this 549
and related fields, stemming from a variety of new types of 550
atropisomers to be discovered utilizing the mechanism 551
discovered in this work. In addition, polymerization of 552
benzoxazine atropisomers was investigated in this study. The 553
benzoxazine atropisomers bearing acetylene functionality 554
exhibited to undergo self-catalyzed polymerization between 555
oxazine ring and acetylene. The resultant thermoset derived 556
from the acetylene containing benzoxazine atropisomers 557
showed exceptionally high thermal stability with high char 558
yield value (62%) and extraordinarily low flammability with 559
low heat release capacity (HRC of 67.2 J g1−K−1) as well as the 560
very low total heat release (THR of 11.2 kJ g −1 ). 561
Consequently, the benzoxazine atropisomers discovered in 562
this study also offers great potential for designing high- 563
performance materials with very low molecular weight 564
precursors.


565

ASSOCIATED CONTENT 566


*
S Supporting Information 567
The Supporting Information is available free of charge on the 568
ACS Publications website at DOI: 10.1021/acs.macro- 569
mol.8b01924. 570

NMR spectra, FT-IR spectra, as well as DSC scans of 571


benzoxazines, and structural coordinates from DFT 572
calculations (PDF) 573

Figure 7. Microscale combustion calorimetric results of poly(oHTI-


a), poly(oHTI-ac), poly(oHTI-ch), and poly(oHTI-cy) as a function
■ AUTHOR INFORMATION
Corresponding Authors
574
575

of temperature: (a) heat release rate and (b) total heat release. Heat *E-mail: hxi3@cwru.edu (H.I.). 576
release rate is the same as heat release capacity at the temperature *E-mail: sfy1@le.ac.uk (S.Y.). 577
ramp rate (1 K/s) used in this experiment. ORCID 578
Hatsuo Ishida: 0000-0002-2590-2700 579
521 KJg−1, while poly(oHTI-ac) shows a much lower THR value of Shengfu Yang: 0000-0001-5684-6388 580
522 11.2 KJg−1. Generally, the lower the values of HRC and THR, Notes 581
523 the higher the flame resistance becomes. The HRC value of The authors declare no competing financial interest.


582
524 poly(oHTI-ac) is lower than the other 47 polymers reported
525 by Lyon et al.45 Furthermore, the HRC value of poly(oHTI-ac) ACKNOWLEDGMENTS 583
526 is even lower than the benzoxazole resins derived from ortho-
Financial support for this work was provided by the National
amide benzoxazines, which shows HRC of ∼92 J g1−K−1.9 The
584
527
Natural Science Foundation of China (No. 51603093), the 585
528 highly aromatic cross-linked networks formed from the
Science and Technology Agency of Jiangsu Province (No. BK 586
529 cyclotrimerization of acetylene along with the polybenzoxazine
20160515), China Postdoctoral Science Foundation (No. 587
530 network leads to a substantial HRC reduction. Therefore, this
2016M600369) and Jiangsu Postdoctoral Science Foundation 588
531 newly developed benzoxazine atropisomers containing acety-
(No. 1601015A). S.Y. wishes to thank the UK Engineering and 589
532 lene functionality opens opportunities for the applications that
Physical Sciences Research Council (EPSRC) and the 590
benefit from low flammability.


533
Leverhulme Trust for grants in support of this work.


591

534 CONCLUSIONS
REFERENCES 592
535 In summary, we reported a new type of atropisomerism in (1) Clayden, J.; Moran, W. J.; Edwards, P. J.; LaPlante, S. R. The 593
536 ortho-tetrahydrophthalimide substituted 1,3-benzoxazine fam- challenge of atropisomerism in drug discovery. Angew. Chem., Int. Ed. 594
537 ily based on stereochemical difference rather than traditionally 2009, 48, 6398−6401. 595
538 observed conformational difference for the first time. The (2) Bringmann, G.; Mortimer, A. J. P.; Keller, P. A.; Gresser, M. J.; 596
539 atropisomerism is evident from NMR measurements and has Garner, J.; Breuning, M. Atroposelectrive synthesis of axially chiral 597
540 been supported by DFT calculations. The excellent agreement biaryl compounds. Angew. Chem., Int. Ed. 2005, 44, 5384−5427. 598

J DOI: 10.1021/acs.macromol.8b01924
Macromolecules XXXX, XXX, XXX−XXX
Macromolecules Article

599 (3) Bringmann, G.; Gulder, T.; Gulder, T. A. M.; Breuning, M. (25) Ghosh, M. K.; Mittal, K. L. Polyimides: Fundamentals and 668
600 Atroposelective total synthesis of axially chiral biaryl natural products. Applications; Marcel Dekker, 1996. 669
601 Chem. Rev. 2011, 111, 563−639. (26) Liu, X. Q.; Jikei, M.; Kakimoto, M. A. Synthesis and properties 670
602 (4) Kozlowski, M. C.; Morgan, B. J.; Linton, E. C. Total synthesis of of AB-type semicrystalline polyimides prepared from polyamic acid 671
603 chiral biaryl natural products by asymmetric biaryl coupling. Chem. ethyl ester precursors. Macromolecules 2001, 34, 3146−3154. 672
604 Soc. Rev. 2009, 38, 3193−3207. (27) Dingemans, T. J.; Mendes, E.; Hinkley, J. J.; Weiser, E. S.; 673
605 (5) Qin, T.; Skraba-Joiner, S. L.; Khalil, Z. G.; Johnson, R. P.; StClair, T. L. Poly (ether imide) s from diamines with para-, meta-, 674
606 Capon, R. J.; Porco, J. A. Atropselective syntheses of (−) and (+) and ortho-arylene substitutions: Synthesis, characterization, and liquid 675
607 rugulotrosin A utilizing point-to-axial chirality transfer. Nat. Chem. crystalline properties. Macromolecules 2008, 41, 2474−2483. 676
608 2015, 7, 234−240. (28) Campo, C. J.; Anastasiou, T.; Uhrich, K. E. Polyanhydrides: the 677
609 (6) Jolliffe, J. D.; Armstrong, R. J.; Smith, M. D. Catalytic effects of ring substitution changes on polymer properties. Polym. Bull. 678
610 enantioselective synthesis of atropisomeric biaryls by a cation-directed 1999, 42, 61−68. 679
611 O-alkylation. Nat. Chem. 2017, 9, 558−562. (29) Yang, C. A.; Wang, Q.; Xie, H.; Zhong, G.; Zhang, H. Synthesis 680
612 (7) Moss, G. P. Basic terminology of stereochemistry. Pure Appl. and characterisation of polymethacrylates containing para-, meta-and 681
613 Chem. 1996, 68, 2193−2222. ortho-monosubstituted azobenzene moieties in the side chain. Liq. 682
614 (8) Christie, G. H.; Kenner, J. The molecular configurations of Cryst. 2010, 37, 1339−1346. 683
615 polynuclear aromatic compounds. Part I. The resolution of 6:6’- (30) Attwood, T. E.; Cinderey, M. B.; Rose, J. B. Poly(arylene ether 684
616 dinitro-and 4:6:4’:6’-tetranitrodiphenic acids into optically active sulfone)s by polyetherification: 5. Effects of molecular structure on 685
617 components. J. Chem. Soc., Trans. 1922, 121, 614−620. toughness. Polymer 1993, 34, 2155−2161. 686
618 (9) Williams, D. H.; Bardsley, B. The vancomycin group of (31) Lyon, R. E.; Walters, R. N. Pyrolysis combustion flow 687
619 antibiotics and the fight against resistant bacteria. Angew. Chem., Int. calorimetry. J. Anal. Appl. Pyrolysis 2004, 71, 27−46. 688
620 Ed. 1999, 38, 1172−1193. (32) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; 689
621 (10) Hubbard, B. K.; Walsh, C. T. Vancomycin assembly: nature’s Robb, M. A.; Cheeseman, J. R.; Montgomery, J. A., Jr., Vreven, T.; 690
622 way. Angew. Chem., Int. Ed. 2003, 42, 730−765. Kudin, K. N.; Burant, J. C.; Millam, J. M.; Iyengar, S. S.; Tomasi, J.; 691
623 (11) Zask, A.; Murphy, J.; Ellestad, G. A. Biological stereoselectivity Barone, V.; Mennucci, B.; Cossi, M.; Scalmani, G.; Rega, N.; 692
624 of atropisomeric natural products and drugs. Chirality 2013, 25, 265− Petersson, G. A.; Nakatsuji, H.; Hada, M.; Ehara, M.; Toyota, K.; 693
625 274. Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, 694
626 (12) LaPlante, S. R.; Edwards, P. J.; Fader, L. D.; Jakalian, A.; Hucke, O.; Nakai, H.; Klene, M.; Li, X.; Knox, J. E.; Hratchian, H. P.; Cross, J. 695
627 O. Revealing atropisomer axial chirality in drug discovery. B.; Bakken, V.; Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R. 696
628 ChemMedChem 2011, 6, 505−513. E.; Yazyev, O.; Austin, A. J.; Cammi, R.; Pomelli, C.; Ochterski, J. W.; 697
629 (13) Casarini, D.; Rosini, C.; Grilli, S.; Lunazzi, L.; Mazzanti, A.
Ayala, P. Y.; Morokuma, K.; Voth, G. A.; Salvador, P.; Dannenberg, J. 698
630 Conformational Studies by Dynamic NMR. 93.1 Stereomutation,
J.; Zakrzewski, V. G.; Dapprich, S.; Daniels, A. D.; Strain, M. C.; 699
631 Enantioseparation, and Absolute Configuration of the Atropisomers
Farkas, O.; Malick, D. K.; Rabuck, A. D.; Raghavachari, K.; Foresman, 700
632 of Diarylbicyclononanes. J. Org. Chem. 2003, 68, 1815−1820.
J. B.; Ortiz, J. V.; Cui, Q.; Baboul, A. G.; Clifford, S.; Cioslowski, J.; 701
633 (14) Mancinelli, M.; Perticarari, S.; Prati, L.; Mazzanti, A.
Stefanov, B. B.; Liu, G.; Liashenko, A.; Piskorz, P.; Komaromi, I.; 702
634 Conformational analysis and absolute configuration of axially chiral
Martin, R. L.; Fox, D. J.; Keith, T.; M. A., Al-Laham, Peng, C. Y.; 703
635 1-aryl and 1,3-bisaryl-xanthines. J. Org. Chem. 2017, 82, 6874−6885.
636 (15) Holly, F. W.; Cope, A. C. Condensation products of aldehydes Nanayakkara, A.; Challacombe, M.; Gill, P. M. W.; Johnson, B.; Chen, 704

637 and ketones with o-aminobenzyl alcohol and o-hydroxybenzylamine. W.; Wong, M. W.; Gonzalez, C.; Pople, J. A. Gaussian 03, Revision 705
638 J. Am. Chem. Soc. 1944, 66, 1875−1879. E.01; Gaussian, Inc.: Wallingford CT, 2004. 706

639 (16) Ning, X.; Ishida, H. Phenolic materials via ring-opening (33) Adamo, C.; Barone, V. Exchanging functionals with improved 707
640 polymerization: Synthesis and characterization of bisphenol-A based long-range behavior and adiabatic connection methods without 708
641 benzoxazines and their polymers. J. Polym. Sci., Part A: Polym. Chem. adjustable parameters: The mPW and mPW1PW models. J. Chem. 709
642 1994, 32, 1121−1129. Phys. 1998, 108, 664−675. 710
643 (17) Ishida, H.; Froimowicz, P. Advanced and Emerging Polybenzox- (34) Lodewyk, M. W.; Siebert, M. R.; Tantillo, D. J. Computational 711
644 azine Science and Technology; Elsevier: Amsterdam, 2017. prediction of 1H and 13C chemical shifts: A useful tool for natural 712
645 (18) Ishida, H.; Agag, T. Handbook of Benzoxazine Resins; Elsevier: products, mechanitic, and synthestic orgainic chemistry. Chem. Rev. 713
646 Amsterdam, 2011. 2012, 112, 1839−1862. 714
647 (19) Nair, C. R. Advances in addition-cure phenolic resins. Prog. (35) Laobuthee, A.; Chirachanchai, S.; Ishida, H.; Tashiro, K. 715
648 Polym. Sci. 2004, 29, 401−498. Asymmetric mono-oxazine: An inevitable product from Mannich 716
649 (20) Ghosh, N. N.; Kiskan, B.; Yagci, Y. Polybenzoxazines-new high reaction of benzoxazine dimers. J. Am. Chem. Soc. 2001, 123, 9947− 717
650 performance thermosetting resins: synthesis and properties. Prog. 9955. 718
651 Polym. Sci. 2007, 32, 1344−1391. (36) Chirachanchai, S.; Laobuthee, A.; Phongtamrug, S. Self 719
652 (21) Zhang, K.; Han, L.; Froimowicz, P.; Ishida, H. A smart latent termination of ring opening reaction of p-substituted phenol-based 720
653 catalyst containing o-trifluoroacetamide functional benzoxazine: benzoxazines: An obstructive effect via intramolecular hydrogen bond. 721
654 precursor for low temperature formation of very high Performance J. Heterocycl. Chem. 2009, 46, 714−721. 722
655 polybenzoxazole with low dielectric constant and high thermal (37) Burger, A.; Ramberger, R. On the polymerization of 723
656 stability. Macromolecules 2017, 50, 6552−6560. pharmaceuticals and other molecular ctrystals. Microchim. Acta 724
657 (22) Ejfler, J.; Krauzy-Dziedzic, K.; Szafert, S.; Lis, T.; Sobota, P. 1979, 72, 259−271. 725
658 Novel chiral and achiral benzoxazine monomers and their thermal (38) Andreu, R.; Reina, J. A.; Ronda, J. C. Studies on the thermal 726
659 polymerization. Macromolecules 2009, 42 (42), 4008−4015. polymerization of substituted benzoxazine monomers: electronic 727
660 (23) Wang, J.; Wang, H.; Liu, J.; Liu, W.; Shen, X. Synthesis, curing effects. J. Polym. Sci., Part A: Polym. Chem. 2008, 46, 3353−3366. 728
661 kinetics and thermal properties of novel difunctional chiral and achiral (39) Agag, T.; Takeichi, T. Synthesis and characterization of novel 729
662 benzoxazines with double chiral centers. J. Therm. Anal. Calorim. benzoxazine monomers containing allyl groups and their high 730
663 2013, 114, 1255−1264. performance thermosets. Macromolecules 2003, 36, 6010. 731
664 (24) Wang, J.; Ren, T.; Wang, Y.; He, W.; Liu, W.; Shen, X.-d. (40) Han, L.; Iguchi, D.; Gil, P.; Heyl, T. R.; Sedwick, V. M.; Arza, 732
665 Synthesis, curing behavior and thermal properties of fluorene- C. R.; Ohashi, S.; Lacks, D. J.; Ishida, H. Oxazine ring-related 733
666 containing benzoxazines based on linear and branched butylamines. vibrational modes of benzoxazine monomers using fully aromatically 734
667 React. Funct. Polym. 2014, 74, 22−30. substituted, deuterated, 15N isotope exchanged, and oxazine-ring- 735

K DOI: 10.1021/acs.macromol.8b01924
Macromolecules XXXX, XXX, XXX−XXX
Macromolecules Article

736 substituted compounds and theoretical calculations. J. Phys. Chem. A


737 2017, 121, 6269−6282.
738 (41) Nguyen, H. X.; Ishida, H.; Hurwitz, F. I.; Gorecki, J. P.;
739 D’Amore, L.; Hyatt, L. H. Polymerization of diethynyldiphenyl-
740 methane: DSC, FTIR, NMR and dielectric characterization. J. Polym.
741 Sci., Part B: Polym. Phys. 1989, 27, 1611−1627.
742 (42) Hamerton, I.; Thompson, S.; Howlin, B. J.; Stone, C. A. New
743 method to predict the thermal degradation behavior of polybenzox-
744 azines from empirical data using structure property relationships.
745 Macromolecules 2013, 46, 7605−7615.
746 (43) Hemvichian, K.; Ishida, H. Thermal decomposition process in
747 aromatic amine-based polybenzoxazines investigated by TGA and
748 GC-MS. Polymer 2002, 43, 4391−4402.
749 (44) Lyon, R. E.; Safronava, N.; Quintiere, J. G.; Stoliarov, S. I.;
750 Walters, R. N.; Crowley, S. Materials properties and fir test results.
751 Fire Mater. 2014, 38, 264−278.
752 (45) Lyon, R. E.; Janssens, M. L. Polymer Flammability; U.S.
753 Department of Transportation, Federal Aviation Administration,
754 Report #DOT/FAA/AR-05/14.

L DOI: 10.1021/acs.macromol.8b01924
Macromolecules XXXX, XXX, XXX−XXX

You might also like