Synthesis, Crystal Structures, and Solid-State Polymerization of 8-[4

CC.C. EEE. III.... JJJ............. LL. MM. NN. OO.. loading. To add this web app to the home screen open the browser option menu and tap on Add to ho...
1 downloads 0 Views 3MB Size
Article pubs.acs.org/crystal

Cite This: Cryst. Growth Des. XXXX, XXX, XXX−XXX

Synthesis, Crystal Structures, and Solid-State Polymerization of 8‑[4(Dimethylamino)phenyl]octa-5,7-diynyl Carbamates Masataka Ikeshima,† Hiroshi Katagiri,†,‡ Wataru Fujiwara,† Shizuo Tokito,†,‡ and Shuji Okada*,†,‡ †

Graduate School of Science and Engineering, Yamagata University, 4-3-16 Jonan, Yonezawa 992-8510, Japan Graduate School of Organic Materials Science, Yamagata University, 4-3-16 Jonan, Yonezawa 992-8510, Japan



Crystal Growth & Design Downloaded from pubs.acs.org by UNIV OF SOUTH DAKOTA on 09/18/18. For personal use only.

S Supporting Information *

ABSTRACT: Six butadiyne monomers with (dimethylamino)phenyl and N-substituted urethane substituents were synthesized and their crystal structures and solid-state polymerization were investigated. In five of the six monomers, directions of the butadiyne stacking and the intermolecular hydrogen bonding between urethane groups coincided, and translational distance d between adjacent butadiyne moieties was around 5 Å. Among them, four monomers showed angle θ between the translation axis of the monomers and the butadiyne moiety of around 45°, which is appropriated for regular 1,4-addition polymerization. These geometries resulted in formation of polydiacetylenes with characteristic excitonic absorption and relatively high conversion. One monomer was found to have large θ, and the regular polymerization was not recognized to be quite low conversion. However, in the remaining one monomer among six, the directions of butadiyne stacking and intermolecular hydrogen bonding were different and d was much longer than 5 Å. However, distance between reacting carbons for 1,4-addition was short enough resulting in comparable conversion without regular polymerization. Although introduction of urethane group is effective to increase probability of the solid-state polymerization of butadiynes, hydrogen bonding direction, methylene chain conformation, and disorder of the aromatic rings were found to afford the variation from the appropriate conditions.



INTRODUCTION π-Conjugated polymers have been investigated as materials for a variety of electronic and optical functions such as electrical conduction,1−3 optical nonlinearlity,4−6 electroluminescence,7−9 electric amplification by field-effect transistors,10−13 photovoltaics,14−17 etc. Among various π-conjugated polymers, polydiacetylenes (PDAs) are quite unique in their preparation process, i.e., topochemical polymerization of butadiyne derivatives.18 Except for butadiyne derivatives, limited monomers can be polymerized in the solid state, e.g., compounds with two CC bonds to show [2 + 2] photocycloaddition reactions,19−24 butadienes,25−30 hexatriene,31 hexatriyne,32 and quinodimethanes.33−35 In the crystals, all these monomers should take the proper polymerizable arrangement in which distances between intermolecular reacting atoms are within the mobile ranges, and the case of butadiynes is as well. As shown in Figure 1, the appropriate conditions of the monomer arrangement for the polymerization have been reported as follows: The distance d between © XXXX American Chemical Society

two adjacent butadiyne monomers along the translation direction, i.e., the polymerization direction, is about 5 Å, and the angle θ between the translation direction and the linear butadiyne moiety is about 45°.36,37 As materials for electronics and photonics, control of the electronic state is an important issue, and introduction of conjugated substituents is one of the general strategies. However, butadiyne derivatives with directly bound conjugated substituents often do not polymerize in their crystalline states because many of the monomer alignments in the crystals are out of polymerizable conditions. In order to increase probability of topochemical polymerization of these derivatives, we have introduced specific substituents taken from symmetrically substituted polymerizable monomers as one of the butadiyne substituents.38,39 For example, polymerizable butadiyne monomers have been obtained by introducing Received: May 29, 2018 Revised: August 17, 2018 Published: September 7, 2018 A

DOI: 10.1021/acs.cgd.8b00815 Cryst. Growth Des. XXXX, XXX, XXX−XXX

Crystal Growth & Design

Article

was investigated in relation to the crystal structure. Optical and electronic properties of the polymers were also examined.



EXPERIMENTAL SECTION

Starting materials, which are not specifically stated, were commercially available and used without further purifications. DMAP butadiyne derivatives were prepared according to the scheme in Figure 3. Details of the synthesis procedures are described below. 2-Methyl-4-[4-(dimethylamino)phenyl]-3-butyn-2-ol 2. 4Bromo-N,N-dimethylaniline 1 (4.00 g, 20 mmol) and 2-methyl-3butyn-1-ol (2.02 g, 24 mmol) were added to triethylamine (60 mL) and dissolved at 60 °C. Then, tetrakis(triphenylphosphine)palladium(0) (462 mg, 0.4 mmol) and copper(I) iodide (152 mg, 0.8 mmol) were added. After stirring the mixture for 24 h, it was concentrated by the solvent evaporation, and the residue was purified by column chromatography (silica gel, dichloromethane) to give 1.95 g (48%) of 2 as a pale green solid. IR (KBr) 3338, 2981, 2223, 1612, 1525, 813 cm−1; 1H NMR (400 MHz, CDCl3) δ = 1.60 (6H, s), 2.05 (1H, s), 2.96 (6H, s), 6.61 (2H, d, J = 7.9 Hz), 7.28 (2H, d, J = 7.9 Hz); 13C NMR (100 MHz, CDCl3) δ = 31.63, 40.18, 65.67, 82.91, 91.46, 109.45, 111.70, 132.65, 150.00. 4-Ethynyl-N,N-dimethylaniline 3. Compound 2 (2.03g, 10 mmol) and pulverized potassium hydroxide (561 mg, 10 mmol) were added to toluene (50 mL), and the mixture was refluxed for 2 h under a nitrogen atmosphere. The reactant was filtered, and solvent in the filtrate was removed under reduced pressure. The residue was purified by column chromatography (silica gel, hexane−dichloromethane (4:1)) to give 950 mg (65%) of 3 as a pale yellowish green solid. IR (KBr) 3291, 2952, 2123, 1604, 1517, 817 cm−1; 1H NMR (400 MHz, CDCl3) δ = 1.56 (1H, s) 2.97 (6H, s), 6.60 (2H, d, J = 7.9 Hz), 7.38 (2H, d, J = 7.9 Hz); 13C NMR (100 MHz, CDCl3) δ = 40.14, 74.74, 84.81, 108.64, 111.61, 133.15, 150.31. 8-[4-(N,N-Dimethylamino)phenyl]octa-5,7-diyn-1-ol 4. To the mixture of 3 (1.02 g, 7 mmol), copper(I) chloride (35 mg, 0.35 mmol), butylamine (30 mL), and methanol (15 mL), a small amount of hydroxylamine hydrochloride was added so as that the color of the mixture became yellow under a nitrogen atmosphere. To this mixture, 6-bromohex-5-yn-1-ol50 (1.24 g, 7 mmol) in butylamine (20 mL) and methanol (10 mL) was added dropwise at ambient temperature for 1.5 h. When the color of the mixture became dark, an appropriate amount of hydroxylamine hydrochloride was added until the color turned to yellow. After addition completion, the mixture was further stirred overnight. Solvent in the mixture was removed under reduced pressure, and the residue was purified by column chromatography (silica gel, dichloromethane) to give 1.15 g (51%) of 4 as a pale yellow solid. IR (KBr) 3232, 2937, 2861, 2235, 2132, 1602, 1521, 813 cm−1; 1H NMR (500 MHz, CDCl3) δ = 1.61−1.76 (4H, m), 2.40 (2H, t, J = 6.8 Hz), 2.97 (6H, s), 3.69 (2H, t, J = 6.2 Hz), 6.59 (2H, d, J = 9.1 Hz), 7.35 (2H, d, J = 9.1 Hz), the hydroxyl proton was not clearly observed; 13C NMR (100 MHz, CDCl3) δ = 19.40, 24.64, 31.74, 40.07, 62.32, 65.95, 72.20, 76.47, 83.06, 108.11, 111.59, 133.70, 150.31 General Synthetic Procedure of 8-[4-(N,N-Dimethylamino)phenyl]octa-5,7-diynyl (N-Substituted Carbamate) 5a−5f. Compound 4 (1.0 mmol) was dissolved in toluene (20 mL) at 60 °C under a nitrogen atmosphere, and dibutyltin dilaurate (3 drops) and then isocyanate (1.1 mmol) were added dropwise. After stirring the mixture for 1 day, its solvent was evaporated and the concentrated residue was purified by a silica-gel column chromatography (silica gel, dichloromethane or dichloromethane-hexane mixture depending on compounds) and finally recrystallized from acetone−hexane mixture to give the corresponding carbamate as colorless crystals. 8-[4-(N,N-Dimethylamino)phenyl]octa-5,7-diynyl N-Phenylcarbamate 5a. Yield 90%; mp 108 °C; IR (KBr) 3336, 2962, 2857, 2857, 2238, 2152, 1706, 1606, 1245, 815 cm−1; 1H NMR (500 MHz, CDCl3) δ = 1.68 (2H, m), 1.83 (2H, m), 2.43 (2H, t, J = 7.1 Hz), 2.98 (6H, s), 4.20 (2H, t, J = 6.5 Hz), 6.58 (1H, br s), 6.59 (2H, d, J = 9.1 Hz,), 7.06 (1H, t, J = 7.1 Hz), 7.27−7.41 (6H, m); 13C NMR (100 MHz, CDCl3) δ = 18.82, 24.35, 27.53, 39.57, 65.33, 66.86,

Figure 1. Solid-state polymerization scheme of butadiyne derivatives. Geometric parameters of optimal monomer alignment for the polymerization are d ≈ 5 Å and θ ≈ 45°.

urethane group forming intermolecular hydrogen bonding in both substituents40,41 or even in one of two substituents.42−44 In our previous studies, we synthesized butadiyne derivatives substituted by 4-(carbazol-9-yl)phenyl (CP) or (9-phenyl)carbazol-3-yl (PC) group to obtain PDAs with small ionization potentials,45,46 which were potential materials for holeinjection electrodes. Since hole injection and transport properties of the PDAs seems to be strongly affected by the highest occupied molecular orbital (HOMO) levels of the corresponding monomers, monomers with a smaller ionization potential are worth investigating. Thus, calculation of the HOMO levels of some model compounds using the Gaussian 09 program,47 in which the molecular structure optimization and HOMO energy level calculation were carried out at the B3LYP/6-31G(d) and B3LYP/6-311+G(d,p) levels, respectively, were conducted. For the three compounds (R = Me) in Figure 2, i.e., penta-1,3-diyne derivatives substituted by (a) CP,

Figure 2. Butadiyne derivatives with an electron-donating aromatic ring: (a) CP, (b) PC, and (c) DMAP derivatives. For calculations, substituent R of the model compounds was set to methyl group. In (c), acetylenic carbons are numbered for explanation of Table 3.

(b) PC, and (c) 4-(dimethylamino)phenyl (DMAP) groups, the HOMO energies obtained were −5.70, −5.55, and −5.24 eV, respectively. As was expected from the chemical structures, the DMAP derivative has higher HOMO energy because of the substitution by the dimethylamino group with a stronger electron donating feature. Although some butadiyne derivatives with a DMAP group have been prepared and polymerized in the solid state,48,49 the relationship between crystal structures and polymerizability has never been reported. In the present study, we synthesized six DMAP butadiyne derivatives with a urethane group. We succeeded to analyze their crystal structures, and polymerizability in the solid state B

DOI: 10.1021/acs.cgd.8b00815 Cryst. Growth Des. XXXX, XXX, XXX−XXX

Crystal Growth & Design

Article

Figure 3. Synthesis scheme of [4-(dimethylamino)phenyl]butadiyne derivatives 5a−5f. 8-[4-(N,N-Dimethylamino)phenyl]octa-5,7-diynyl N-Phenethylcarbamate 5f. Yield 82%; mp 116 °C; IR (KBr) 3343, 3027, 2938, 2237, 2134, 1687, 1604, 1536, 1444, 1361, 1284, 1253, 1201, 1143, 1095, 1029, 946, 817, 748, 698, 528 cm−1; 1H NMR (500 MHz, CDCl3) δ = 1.61 (2H, m), 1.74 (2H, m), 2.38 (2H, t, J = 6.8 Hz), 2.81 (2H, t, J = 6.8 Hz), 2.97 (6H, s), 3.44 (2H, dt, J = 6.1, 6.1 Hz), 4.08 (2H, t, J = 6.1 Hz), 4.69 (1H, br t), 6.58 (2H, d, J = 8.5 Hz), 7.13−7.27 (3H, m), 7.28−7.37 (4H, m); 13C NMR (125 MHz, CDCl3) δ = 19.31, 24.85, 28.15, 36.12, 40.05, 42.09, 64.16, 66.07, 72.21, 76.52, 82.77, 108.13, 111.60, 126.45, 128.59, 128.75, 133.70, 138.76, 150.34, 156.50. Found: C, 77.11; H, 7.28; N, 7.38%. Calcd for C25H28N2O2: C, 77.29; H, 7.26; N, 7.21%. Melting points were determined by differential scanning calorimetry (DSC) using an SII DSC 6220 calorimeter with a heating rate of 5 °C min−1. FT-IR spectra were measured using a Horiba FT-720 spectrometer. 1H and 13C NMR spectra were recorded on a JEOL JNM-ECX-400 or -500 instrument. Chemical shifts were calibrated by TMS for 1H NMR and by CDCl3 (77 ppm) for 13C NMR. The elemental analyses were performed with a PerkinElmer 2400 series II CHNS/O elemental analyzer. UV−vis diffuse reflectance spectra were measured using a JASCO V-570 spectrophotometer with an integrated sphere (JASCO ILN-472). X-ray single crystal analysis was performed using Rigaku Saturn 724 with a MoKα source. Data collection at 263 K, cell refinement, and data reduction were performed using the CrystalClear-SM software.51 The structures were solved by the direct method using SHELXT52 and refined using SHELXL2014.53 All materials for publication were prepared by Yadokari-XG 2009 software.54,55 Single crystals for the analysis were prepared from the acetone−hexane solutions by the slow evaporation method. Solid-state polymerization was stimulated by irradiation of UV at 254 nm using a 16 W lamp (UVP R-52G). When the diffuse reflectance spectra in the course of photopolymerization were measured, UV was irradiated through the quartz window of the holder of the samples, in which the monomers were mixed with potassium bromide and ground. Conversions to the corresponding polymers were obtained by the gravimetric method. The samples after UV irradiation for 350 h were weighed (w0) and put into acetone to be the dispersions with the concentration of 1 g/L. They were sonicated for 10 min and stored for more than 12 h. Then, they were filtered using membrane filters (Durapore 1.0 μm, VVHP). The filtered materials were dried for 1 day and weighed (w). The conversions were calculated by the following equation: 100w/w0 (%). The ionization potentials of the monomers and the corresponding polymers were examined using photoelectron yield spectroscopy (PYS) in air obtained by a Riken Keiki AC-3 spectrometer. For the monomer samples, the powdered crystals were pressed on glass substrates, and the light intensity for the measurements was set to 10 nW. For the UV-irradiated samples, the samples on membrane filters,

72.34, 75.76, 82.30, 108.56, 111.10, 118.39, 122.87, 128.53, 133.23, 139.34, 149.85, 153.23. Found: C, 76.47; H, 6.74; N, 7.78%. Calcd for C23H24N2O2: C, 76.64; H, 6.71; N, 7.77%. 8-[4-(N,N-Dimethylamino)phenyl]octa-5,7-diynyl N-[4(Trifluoromethyl)phenyl]carbamate 5b. Yield 58%; mp 163 °C; IR (KBr) 3330, 2964, 2896, 2237, 2132, 1697, 1606, 1540, 1365, 1328, 1245, 1155, 1108, 840, 811, 773, 653 cm−1; 1H NMR (500 MHz, CDCl3) δ = 1.68 (2H, m), 1.84 (2H, m), 2.43 (2H, t, J = 6.8 Hz), 2.97 (6H, s), 4.22 (2H, t, J = 6.5 Hz), 6.59 (2H, d, J = 9.1 Hz), 6.80 (1H, br s), 7.34 (2H, d, J = 9.1 Hz), 7.51 (2H, d, J = 9.1 Hz), 7.55 (2H, d, J = 9.1 Hz); 13C NMR (125 MHz, CDCl3) δ = 19.32, 24.83, 27.96, 40.05, 65.03, 66.30, 72.16, 82.54, 108.02, 111.62, 118.02, 124.15 (q, JCF = 204 Hz), 125.15 (q, JCF = 25 Hz), 126.32 (q, JCF = 3 Hz), 133.73, 141.03, 150.40, 153.15, one of the acetylenic peaks overlapped with a solvent peak. Found: C, 67.17; H, 5.52; N, 6.46%. Calcd for C24H23F3N2O2: C, 67.28: H, 5.41: N, 6.54%. 8-[4-(N,N-Dimethylamino)phenyl]octa-5,7-diynyl N-(4Chlorophenyl)carbamate 5c. Yield 88%; mp 131 °C; IR (KBr) 3288, 2925, 2857, 2360, 1697, 1592, 1245, 823 cm−1; 1H NMR (500 MHz, CDCl3) δ = 1.68 (2H, m), 1.83 (2H, m), 2.43 (2H, t, J = 6.8 Hz), 2.98 (6H, s), 4.20 (2H, t, J = 6.3 Hz), 6.59 (2H, d, J = 8.8 Hz), 7.25−7.28 (4H, m), 7.33 (1H, br s), 7.34 (2H, d, J = 8.8 Hz); 13C NMR (100 MHz, CDCl3) δ = 19.03, 24.52, 27.75, 39.81, 64.58, 66.00, 71.95, 82.39, 107,77, 111.38, 119.59, 128.44, 128.77, 133.50, 136.38, 150.13, 153.21, one of the acetylenic peaks overlapped with a solvent peak. Found: C, 69.95; H, 5.87; N, 7.09%. Calcd for C23H23N2O2: C, 69.92; H, 5.73; N, 7.04%. 8-[4-(N,N-Dimethylamino)phenyl]octa-5,7-diynyl N-(4Methoxyphenyl)carbamate 5d. Yield 69%; mp 105 °C; IR (KBr) 3316, 2965, 2938, 2235, 2130, 1702, 1606, 1536, 1415, 1363, 1243, 1180, 1074, 1025, 817, 781, 734, 634, 522 cm−1; 1H NMR (500 MHz, CDCl3) δ = 1.66 (2H, m), 1.82 (2H, m), 2.41 (2H, t, J = 6.8 Hz), 2.97 (6H, s), 3.77 (3H, s), 4.18 (2H, t, J = 6.3 Hz), 6.54 (1H, br s), 6.59 (2H, d, J = 9.1 Hz), 6.85 (2H, d, J = 9.1 Hz), 7.28 (2H, br d), 7.34 (2H, d, J = 9.1 Hz); 13C NMR (125 MHz, CDCl3) δ = 19.31, 24.86, 28.06, 40.03, 55.45, 64.51, 66.15, 72.20, 76.59, 82.70, 108.11, 111.60, 114.20, 120.64, 130.91, 133.70, 150.35, 153.94, 155.90. Found: C, 73.61; H, 7.03; N, 7.15%. Calcd for C24H26N2O3: C, 73.82; H, 6.71; N, 7.17%. 8-[4-(N,N-Dimethylamino)phenyl]octa-5,7-diynyl N-Benzylcarbamate 5e. Yield 80%; mp 86 °C; IR (KBr) 3286, 3031, 2948, 2896, 2233, 2130, 1681, 1604, 1521, 1454, 1365, 1267, 1232, 1193, 1135, 1091, 1027, 944, 815, 750, 698, 530 cm−1; 1H NMR (500 MHz, CDCl3) δ = 1.64 (2H, m), 1.77 (2H, m), 2.40 (2H, t, J = 6.8 Hz), 2.98 (6H, s), 4.13 (2H, t, J = 6.6 Hz), 4.37 (2H, d, J = 5.7 Hz), 4.96 (1H, br t), 6.59 (2H, d, J = 9.0 Hz), 7.26−7.31 (3H, m), 7.32−7.37 (4H, m); 13C NMR (125 MHz, CDCl3) δ = 19.29, 24.84, 28.07, 40.05, 45.00, 64.40, 66.08, 72.21, 76.53, 82.76, 108.11, 111.59, 127.42 (overlapping two peaks), 128.62, 133.70, 138.49, 150.33, 156.62. Found: C, 76.85; H, 7.16; N, 7.52%. Calcd for C24H26N2O2: C, 76.98; H, 7.00; N, 7.48%. C

DOI: 10.1021/acs.cgd.8b00815 Cryst. Growth Des. XXXX, XXX, XXX−XXX

Crystal Growth & Design

Article

Table 1. Crystallographic Data of 5a−5f compound

5a

5b

5c

5d

5e

5f

formula formula weight crystal system space group a/Å b/Å c/Å α/deg β/deg γ/deg V/Å Z Dx/Å3 R wR2 GOF

C23H24N2O2 360.44 triclinic P1̅ 5.1870(14) 12.317(4) 15.634(5) 102.428(9) 95.580(11) 93.238(10) 967.7(5) 2 1.237 0.0527 0.1490 1.046

C24H23F3N2O2 428.44 triclinic P1̅ 4.9855(2) 13.0325(5) 18.1787(8) 109.430(4) 91.960(4) 100.825(4) 1088.07(8) 2 1.308 0.0458 0.1446 1.027

C23H23ClN2O2 394.88 triclinic P1̅ 5.159(3) 11.221(8) 18.366(11) 101.280(18) 95.30(3) 97.728(15) 1025.5(11) 2 1.279 0.0573 0.1465 1.098

C24H26N2O2 390.47 triclinic P1̅ 10.0211(4) 18.6993(6) 23.8075(8) 75.710(3) 87.998(3) 83.592(3) 4296.1(3) 8 1.207 0.0556 0.1656 1.027

C24H26N2O2 374.47 triclinic P1̅ 5.9428(3) 8.6694(5) 42.721(3) 88.9940(10) 89.4990(10) 75.168(6) 2127.3(2) 4 1.169 0.0636 0.1734 1.066

C25H28N2O2 388.49 triclinic P1̅ 5.113(3) 12.199(6) 18.218(9) 74.023(11) 86.068(14) 85.288(15) 1087.5(10) 2 1.186 0.0522 0.1502 1.036

which were obtained in the experiments to evaluate the conversions mentioned above, were used, and the light intensity was 100 nW.

the monomer alignment in the crystals. For 5a, 5b, 5c, 5e, and 5f, the adjacent molecules have only the translation relation in the crystals (Figure 5a−c and Figure 6a,b,d). Thus, the polymerization direction must be along one of three crystallographic axes. The shortest side lengths of their unit cells are around 5 Å of the a axis, indicating that the a axis can be the solid-state polymerization direction. For 5e, two conformers form two monomer arrays, each of which is composed of one conformer. Figure 6a,b show the monomer arrays compose of conformer A (A1 and A2 in Figure 4e) and conformer B (B1−B4 in Figure 4e). However, for 5d, two of four conformers form one monomer array and the other two form another monomer array along a axis. In each array, different conformers alternately align (··ABAB·· or ··CDCD·· of conformers in Figure 4d), and one of the array is shown in Figure 5d. Except for 5e, the monomer alignment motif mentioned above is all based on intermolecular hydrogen bonding of the urethane groups between adjacent molecules, i.e., directions of the hydrogen bonding and the monomer stacking array are the same. However, the monomer alignment motif of 5e is different. Although there is intermolecular hydrogen bonding of urethane groups between adjacent molecules even in 5e along the b axis (Figure 6c), the different conformers with opposite molecular directions are alternately aligned and the butadiyne moieties are not stacked along this direction. Thus, polymerization cannot progress along the b axis, and the nearest stacking distance between the monomers in the same orientation was found to be along the a axis. Table 2 summarizes structural parameters and characteristics in the crystals. As mentioned above, directions of intermolecular hydrogen bonding and butadiyne stacking are the same except for 5e. Thus, the following discussion is related for the five compounds except 5e. When we pay attention to hydrogen bonding along butadiyne stacking, distance between nitrogen and oxygen atoms of adjacent urethane groups is from 2.88 to 3.03 Å, and the angle between O···H−N intermolecular hydrogen bond and H−N covalent bond of urethane groups is from 152.2° to 163.8°. These relatively small variations seem to cause a basic motif of molecular stacking. As a result, distance d becomes around 5 Å from 4.99 to 5.19 Å. Focusing on the conformation of the butylene chain between butadiyne and urethane groups, three compounds have all-trans structures (ttt in Table 2), and two have gauche-trans-gauche



RESULTS AND DISCUSSION Single crystals appropriated for X-ray crystallographic analysis were obtained for all compounds in this study. Crystallographic data of monomers 5a−5f are shown in Table 1. Monomer structures in the crystals are displayed in Figure 4. All crystals belongs to triclinic space group P1̅. However, the number of molecules in a unit cell (Z) is different. Compounds 5a, 5b, 5c, and 5f have two molecules related by the center of symmetry in a unit cell. Molecules of 5a showed disorder in the butylene chain (A1 and A2 in Figure 4a), and their occupancy factors were 0.865 and 0.135, respectively. For 5b, rotation disorder was found in the trifluoromethyl group (A1 and A2 in Figure 4b), and their occupancy factors were 0.663 and 0.337, respectively. Monomer 5c has a single molecular structure as shown in Figure 4c. For 5f, disorder was recognized in the phenethyl group (A1 and A2 in Figure 4f), and their occupancy factors were 0.510 and 0.490, respectively. However, 5d and 5e were analyzed to have molecules with different conformations. For 5d, four conformers A−D were found as shown in Figure 4d. Structures of A and D are similar and also those of B and C are similar. However, within each pair, the details of the structures are different. Molecular structures of 5e were much complicated. There are two conformers A and B, and each conformer showed disorder. For conformer A, disorder was observed in the benzyl group (A1 and A2 in Figure 4e), and their occupancy factors were 0.5 each. Meanwhile, conformer B has disorder around both aromatic rings. The structures of B1 and B2 in Figure 4e are shown with disorder in the benzyl group, and the DMAP group is fixed to one structure. Occupancy factors of the benzyl group in B1 and B2 were 0.5 each. In the structures of B1, B3, and B4, the benzyl group is fixed, and disorder around the DMAP group is shown. Occupancy factors of the DMAP group for these structures were 0.34, 0.34, and 0.32, respectively. As mentioned above, monomer alignment is more important from the point of view of the solid-state polymerization, and adjacent monomer distance d should be around 5 Å. Thus, we picked up monomer stacking arrays along one direction with d of around 5 Å from the crystal structures. Figures 5 and 6 show D

DOI: 10.1021/acs.cgd.8b00815 Cryst. Growth Des. XXXX, XXX, XXX−XXX

Crystal Growth & Design

Article

Figure 5. Monomer arrays in the crystals of N-phenylcarbamates: (a) 5a, (b) 5b, (c) 5c, and (d) 5d. Hydrogen atoms are omitted. Intermolecular hydrogen bonds are indicated by red dashed lines.

Figure 6. Monomer arrays in the crystals of N-benzylcarbamates and N-phenethylcarbamates: (a−c) 5e and (d) 5f. Hydrogen atoms are omitted. Intermolecular hydrogen bonds are indicated by red dashed lines. Figure 4. Monomer structures in the crystals: (a) 5a, (b) 5b, (c) 5c, (d) 5d, (e) 5e, and (f) 5f. In the symbols of the structures, alphabets indicate conformers and numbers indicate disordered structures of the conformers.

structures (gtg in Table 2). From the point of view of alkyl chain packing, the latter structures are apt to be loose. Actually, E

DOI: 10.1021/acs.cgd.8b00815 Cryst. Growth Des. XXXX, XXX, XXX−XXX

Crystal Growth & Design

Article

Table 2. Structural Parameters and Characteristics of 5a−5f in Crystals packing parameters

hydrogen-bonding parameters

compound

da/Å

θa/deg

ϕb/deg

O···N/Å

∠O···H−N/deg

conformation of butylene groupc

5a 5b 5c 5d

5.19 4.99 5.16 4.97, 5.07d 5.03, 5.01g 5.94 5.94 5.11

47.4 51.6 48.1 71.0, 70.5e 71.3, 71.0h 41.1 53.4, 55.7, 53.3j 45.7

70.3 67.6 75.4 17.8, 79.4e 89.2, 26.3h 66.8 28.8, 47.4, 27.1j 80.0

3.03 2.92 3.00 2.89, 2.88f 2.91, 2.90i (2.87)k (2.82)k 2.93

163.8 155.3 156.0 158.8, 164.1f 153.6, 152.2i (178.0)k (156.1)k 156.1

gtg ttt gtg ttt, ttt ttt, ttt gtg ttt ttt

5e 5f

Geometric definitions of distance d and angle θ are shown in Figure 1. bGeometric definition of angle ϕ is displayed in Figure 7. cMethylene chain conformation is expressed for three C−C bonds from the carbon attached to the butadiyne moiety to that next to the urethane group. Symbols g and t indicate gauche and trans conformations, respectively. dAverage distances from conformer A to conformer B and from B to A, respectively, in Figure 4d, in the direction from the carbon atom to the oxygen atom of the carbonyl group. eFor conformer A and for conformer B, respectively, in Figure 4d. fFrom conformer A to conformer B and from B to A, respectively, in Figure 4d. gAverage distances from conformer C to conformer D and from D to C, respectively, in Figure 4d, in the direction from the carbon atom to the oxygen atom of the carbonyl group. hFor conformer C and for conformer D, respectively, in Figure 4d. iFrom conformer C to conformer D and from D to C, respectively, in Figure 4d. jAngles for the structures of B1, B3, and B4, respectively, in Figure 4e. kStacking direction of butadiyne moieties and direction of intermolecular hydrogen bonding are not parallel. a

the gtg structures showed longer d than the ttt structures, i.e., d of no more than 5.11 Å for the ttt structures and that of more than 5.16 Å for the gtg structures. Meanwhile, the range of angle θ is spread, although four compounds of five showed θ of around 45° from 45.7° to 51.6°. Large deviation of θ from 45° was found in 5d, i.e., θ of about 71°. However, 5e, whose intermolecular hydrogen bonding is not parallel to the butadiyne stacking direction, showed longer d of 5.94 Å, although θ is around 45°, i.e., 41.1° and 55.7°. According to the criteria for the solid-state 1,4-addition polymerization of butadiyne derivatives, i.e., d ≈ 5 Å and θ ≈ 45°, these results suggested that 5a, 5b, 5c, and 5f would polymerize in the regular 1,4-addition manner, while regular polymerization was not expected for 5d with larger θ of about 70° and 5e with longer d of about 5.9 Å. Angle ϕ in Table 2 indicates the angle between the plane of the butadiyne stacking array and the πconjugated phenyl plane directly attached to the butadiyne moiety (Figure 7). The former plane can be a π-conjugated

The other angle problem is related to the bond angle changes at C1 and C4 in Figure 2c in the course of polymerization, i.e., transformation of sp carbons with the bond angle of 180° to sp2 carbons with the bond angle of 120°. From the model structures of [4-(dimethylamino)phenyl]butadiyne moiety and the corresponding polymer, the atom displacements (r) due to polymerization were calculated (Supporting Information). Large r values were obtained for the nitrogen and carbons in the dimethylamino group to be 1.9 Å and 2.2−2.4 Å, respectively. However, the r values for the phenyl carbons ranged from 0.4−1.2 Å, which were comparable to those of C1 and C4 of 1.1 Å. Thus, in order to realize more efficient solid-state polymerization of phenylbutadiyne groups, introduction of intramolecular or intermolecular mechanisms to compensate relatively large displacement of the substituents of the aromatic ring should be considered, e.g., introduction of flexible or bulky substituents.56−59 Solid-state polymerization of the powder samples were monitored by UV−vis diffuse reflectance spectra (Figure 8). Six spectra in Figure 8 can be classified into two categories. One is the spectra with characteristic absorption due to PDA exciton and the spectra of 5a, 5b, 5c, and 5f belong to this category. These absorption spectra indicate regular 1,4addition polymerization to form an ene-yne sequence of the PDA structure since there are a lot of examples that regularly polymerized PDAs display characteristic absorption bands in the longer wavelength region than 500 nm, resulting in colors from blue to red in general.56−58,60−67 The absorption maxima of these spectra at the initial stage of polymerization were 637, 660, 672, and around 610 nm for 5a, 5b, 5c, and 5f, respectively. As mentioned above, these four compounds were found to have monomer stacks appropriate for solid-state polymerization, and these structural features prove regular 1,4addition polymerization resulting in spectra with the excitonic absorption. The other is spectra without absorption maxima in the visible region and absorbance increased as the wavelength became shorter from the absorption edge, e.g., 5d and 5e. The data of crystallographic structures of 5d and 5e revealed that these monomer arrays were not suitable for regular 1,4-

Figure 7. Dihedral angle ϕ between the phenyl ring plane and the butadiyne-array plane, which is considered to become a π-conjugated plane of PDA backbone after polymerization.

backbone plane after polymerization. Thus, if ϕ is small even after polymerization, the conjugation effect from substituted phenyl groups effectively transferred to the backbone. From the geometric consideration based on d of 5 Å and the phenyl ring thickness of 3.4 Å, the smallest ϕ is calculated to be around 43°. However, 5a, 5b, 5c, and 5f showed much larger ϕ angles from 67° to 80°. This may be due to steric hindrance between dimethylamino groups. For 5d and 5e, smaller angles than 43° were observed. In the case of 5d, conformers with small ϕ of around 20° and large ϕ of near 90° are alternately aligned. Meanwhile, large d can be small ϕ for 5e. F

DOI: 10.1021/acs.cgd.8b00815 Cryst. Growth Des. XXXX, XXX, XXX−XXX

Crystal Growth & Design

Article

Figure 8. UV−visible diffuse reflectance spectra in the course of UV irradiation: (a) 5a, (b) 5b, (c) 5c, (d) 5d, (e) 5e, and (f) 5f.

easier than the regular 1,4-addition. For 5e, both conformer arrays have smaller C1···C4 distances than the corresponding C1···C2 and C3···C4 distances, and the C1···C4 distance between two conformers of A is 3.97 Å, which is similar to that of 5b. Accordingly, 5e can be polymerized in 1,4-addtion manner, and the conversion may become relatively high. However, d of 5.94 Å for 5e is much longer than 5 Å, and regularity of the polymerization is low because of the large structural mismatch between the monomer array and the resulting polymer in the crystals. The ionization potentials of the monomers and the corresponding polymers were evaluated by PYS in air. Among the six compounds, the monomer ionization potentials of 5c, 5e, and 5f could be obtained to be 5.72, 5.68, and 5.66 eV, respectively. In our previous studies, we obtained the ionization potentials of CP- and PC-type monomers (Figure 2) to be about 6.0 and 5.8 eV, respectively.45,46 In comparison with these values, DMAP-type monomers have smaller ionization potentials, which agreed with the results on the HOMO energy calculation mentioned in the Introduction.

addition and that these structural features reflected their absorption spectra without characteristic excitonic bands. Conversion from the monomers to the corresponding polymers were obtained by the gravimetric method. Since UV light irradiated was not able to fully penetrate the pulverized samples, the absolute conversions were not high. However, relative polymerizability can be discussed. For the former class of compounds, conversions of 5a, 5b, 5c, and 5f were 6%, 12%, 9%, and 16%, respectively. Conversions of 5d and 5e in the latter class were ∼0% and 7%, respectively, and 5e was found to show relatively high conversion comparable to 5a and 5c. In order to clarify this reason, intermolecular C···C distances relating to 1,4- and 1,2-addition polymerization were investigated for all structures (Table 3). As was expected from the discussion above, polymerization of 5a, 5b, 5c, and 5f were again confirmed to be 1,4-addtion, i.e., C1···C4 distances of 5a, 5b, 5c, and 5f are 3.36−3.97 Å, which are shorter than the corresponding C1···C2 and C3···C4 distances. However, C1··· C4 distances of 5d are longer than the corresponding C1···C2 and C3···C4 distances. This indicates that 1,2-addition of 5d is G

DOI: 10.1021/acs.cgd.8b00815 Cryst. Growth Des. XXXX, XXX, XXX−XXX

Crystal Growth & Design

Article

molecular hydrogen bonding and the butadiyne stacking, methylene chain conformation including gauche form, and disorder of the aromatic rings at the molecular ends. Thus, if we can suppress such structural variations by introducing another molecular interactions, finer tuning for the crystal structures having regularly polymerizable stacks seems to be possible. In addition, introduction of intramolecular or intermolecular mechanisms to compensate relatively large displacement of the substituents of the aromatic ring seemed to be considered for efficient conversion.

Table 3. Intermolecular Carbon Atom Distance Related to Solid-State Addition Polymerization intermolecular carbon atom distanceb/Å compound 5a 5b 5c 5d

5e 5f

structure

A···B B···A C···D D···C A···A B1···B1

a

C1···C4c

C1···C2d

C3···C4d

3.825 3.966 3.856 5.100 5.272 5.247 5.184 3.973 4.771 3.664

4.467 4.350 4.454 4.644 4.802 4.763 4.734 5.086 5.289 4.354

4.468 4.336 4.449 4.741 4.849 4.805 4.751 5.120 5.450 4.365



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.cgd.8b00815. Estimation of atom displacements due to polymerization (PDF)

a

Molecular structures in the crystals are shown in Figure 4, and their alignments are displayed in Figures 5 and 6. bCarbon numbering is shown in Figure 2c. cBond formation between these two carbons results in 1,4-addition polymerization. dBond formation between these two carbons results in 1,2-addition polymerization.

Accession Codes

CCDC 1844165−1844170 contain the supplementary crystallographic data for this paper. These data can be obtained free of charge via www.ccdc.cam.ac.uk/data_request/cif, or by emailing [email protected], or by contacting The Cambridge Crystallographic Data Centre, 12 Union Road, Cambridge CB2 1EZ, UK; fax: +44 1223 336033.

The ionization potentials of the photopolymerized samples could be obtained for 5c and 5f, and they were 5.92 and 6.12 eV, respectively. These values were quite unexpected because the ionization potentials of PDAs usually decrease compared with the corresponding monomers because of the HOMO energy elevation caused by π-conjugation extension. The reason is not clear, but one of the plausible explanations is as follows: Generally, the dimethylamino group is less stable than the carbazolyl group, and the side reactions other than polymerization, such as photooxidation,68,69 may occur during UV irradiation at the outermost surface of the samples. These side reactions may lower the HOMO energy levels to increase the ionization potentials.



AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. ORCID

Hiroshi Katagiri: 0000-0003-4100-9995 Shuji Okada: 0000-0001-9519-6648 Notes



The authors declare no competing financial interest.



CONCLUSION Six urethane derivatives with a DMAP-substituted butadiyne moiety were synthesized, and their crystal structures and solidstate polymerization behaviors were studied. Five of six compounds 5a, 5b, 5c, 5d, and 5f showed the packing motif that the direction of the intermolecular hydrogen bonding between urethane groups and that of the butadiyne stacking were parallel, and d was around 5 Å. When the butylene chain conformation is gtg, d becomes longer compared with the ttt conformation. Among them, four compounds except 5d showed excitonic absorption after UV-induced polymerization, indicating that the regular 1,4-addition polymerization progressed to form PDAs. In the case of 5d, θ was much larger than 45°, and the conversion to the polymer was quite low. For 5e, the packing motif was not like the other five compounds, and d was much larger than 5 Å. However, the intermolecular carbon distance for 1,4-addition polymerization was short enough, and the conversion to the polymer was comparable to that of the compounds showing excitonic absorption during polymerization, although polymerization of 5e was not so regular. As we have been demonstrated, introduction of intermolecular hydrogen bonding to the arylbutadiyne monomers generally increases probability of the solid-state polymerization. In the present study, this empirical rule was crystallographically confirmed. Partial structures, which caused variation from the suitable condition for 1,4-addition polymerization, were the directional mismatch between the inter-

ACKNOWLEDGMENTS We thank Mr. Shinji Ishii and Mr. Shota Kaneko for their contribution in the synthetic experiments. We also thank Prof. Junji Kido, Prof. Yong-Jin Pu, and Prof. Hisahiro Sasabe for cooperation on calculations using the Gaussian 09 program.



REFERENCES

(1) Shirakawa, H.; Louis, E. J.; MacDiarmid, A. G.; Chiang, C. K.; Heeger, A. J. Synthesis of Electrically Conducting Organic Polymers: Halogen Derivatives of Polyacetylene, (CH)x. J. Chem. Soc., Chem. Commun. 1977, 578−580. (2) Pron, A.; Rannou, P. Processible Conjugated Polymers: From Organic Semiconductors to Organic Metals and Superconductors. Prog. Polym. Sci. 2002, 27, 135−190. (3) Osaka, I.; McCullough, R. D. Advances in Molecular Design and Synthesis of Regioregular Polythiophenes. Acc. Chem. Res. 2008, 41, 1202−1214. (4) Sauteret, C.; Hermann, J.-P.; Frey, R.; Pradère, F.; Ducuing, J.; Baughman, R. H.; Chance, R. R. Optical Nonlinearities in OneDimensional-Conjugated Polymer Crystals. Phys. Rev. Lett. 1976, 36, 956−959. (5) Yamamoto, T.; Yamada, W.; Takagi, M.; Kizu, K.; Maruyama, T.; Ooba, N.; Tomaru, S.; Kurihara, T.; Kaino, T.; Kubota, K. πConjugated Soluble Poly(aryleneethynylene) Type Polymers. Preparation by Palladium-Catalyzed Coupling Reaction, Nonlinear Optical Properties, Doping, and Chemical Reactivity. Macromolecules 1994, 27, 6620−6626. (6) Hegde, P. K.; Adhikari, A. V.; Manjunatha, M. G.; Poornesh, P.; Umesh, G. Third-Order Nonlinear Optical Susceptibilities of New H

DOI: 10.1021/acs.cgd.8b00815 Cryst. Growth Des. XXXX, XXX, XXX−XXX

Crystal Growth & Design

Article

Copolymers Containing Alternate 3,4-Dialkoxythiophene and (1,3,4Oxadiazolyl)pyridine Moieties. Opt. Mater. 2009, 31, 1000−1006. (7) Burroughes, J. H.; Bradley, D. D. C.; Brown, A. R.; Marks, R. N.; Mackay, K.; Friend, R. H.; Burns, P. L.; Holmes, A. B. Light-Emitting Diodes Based on Conjugated Polymers. Nature 1990, 347, 539−541. (8) Kraft, A.; Grimsdale, A. C.; Holmes, A. B. Electroluminescent Conjugated Polymers−Seeing Polymers in a New Light. Angew. Chem., Int. Ed. 1998, 37, 402−428. (9) Zhang, X.; Kale, D. M.; Jenekhe, S. A. Electroluminescence of Multicomponent Conjugated Polymers. 2. Photophysics and Enhancement of Electroluminescence from Blends of Polyquinolines. Macromolecules 2002, 35, 382−393. (10) Fuchigami, H.; Tsumura, A.; Koezuka, H. Polythienylenevinylene Thin-Film Transistor with High Carrier Mobility. Appl. Phys. Lett. 1993, 63, 1372−1374. (11) Hamilton, M. C.; Kanicki, J. Organic Polymer Thin-Film Transistor Photosensors. IEEE J. Sel. Top. Quantum Electron. 2004, 10, 840−848. (12) Lee, J.; Han, A.-R.; Kim, J.; Kim, Y.; Oh, J. H.; Yang, C. Solution-Processable Ambipolar Diketopyrrolopyrrole−Selenophene Polymer with Unprecedentedly High Hole and Electron Mobilities. J. Am. Chem. Soc. 2012, 134, 20713−20721. (13) Kumaki, D.; Fujisaki, Y.; Tokito, S. Reduced Contact Resistance and Highly Stable Operation in Polymer Thin-Film Transistor with Aqueous MoOx Solution Contact Treatment. Org. Electron. 2013, 14, 475−478. (14) Sariciftci, N. S.; Braun, D.; Zhang, C.; Srdanov, V. I.; Heeger, A. J.; Stucky, G.; Wudl, F. Semiconducting Polymer-Buckminsterfullerene Heterojunctions: Diodes, Photodiodes, and Photovoltaic Cells. Appl. Phys. Lett. 1993, 62, 585−587. (15) Günes, S.; Neugebauer, H.; Sariciftci, N. S. Conjugated Polymer-Based Organic Solar Cells. Chem. Rev. 2007, 107, 1324− 1338. (16) Facchetti, A. π-Conjugated Polymers for Organic Electronics and Photovoltaic Cell Applications. Chem. Mater. 2011, 23, 733−758. (17) Adegoke, O. O.; Jung, I. H.; Orr, M.; Yu, L.; Goodson, T., III Effect of Acceptor Strength on Optical and Electronic Properties in Conjugated Polymers for Solar Applications. J. Am. Chem. Soc. 2015, 137, 5759−5769. (18) Wegner, G. Topochemische Reaktionen von Monomeren mit konjugierten Dreifachbindungen I. Mitt.: Polymerisation von Derivaten des 2.4-Hexadiin-1.6-diols im kristallinen Zustand. Z. Naturforsch., B: J. Chem. Sci. 1969, 24B, 824−832. (19) Hasegawa, M.; Suzuki, Y. Four-Center Type Photopolymerization in the Solid State: Poly-2,5-distrylpyrazine. J. Polym. Sci., Part B: Polym. Lett. 1967, 5, 813−815. (20) Hasegawa, M.; Hashimoto, Y. Kaleidoscopic Photochemical Behavior of Unsymmetric Diolefin Crystals. Mol. Cryst. Liq. Cryst. Sci. Technol., Sect. A 1992, 219, 1−15. (21) Hasegawa, M. Photodimerization and Photopolymerization of Diolefin Crystals. Adv. Phys. Org. Chem. 1995, 30, 117−171. (22) Garai, M.; Santra, R.; Biradha, K. Tunable Plastic Films of a Crystalline Polymer by Single-Crystal-to-Single-Crystal Photopolymerization of a Diene: Self-Templating and Shock-Absorbing TwoDimensional Hydrogen-Bonding Layers. Angew. Chem., Int. Ed. 2013, 52, 5548−5551. (23) Mandal, R.; Garai, M.; Biradha, K. Hydrogen-bonded Two-fold Interpenetrated Diamondoid Networks for Solid-State [2 + 2] Polymerizations of Criss-crossed Olefins: Molecular Connections vs Supramolecular Connections. Cryst. Growth Des. 2017, 17, 5061− 5064. (24) Garai, M.; Biradha, K. Water-Resistant and Transparent Plastic Films from Functionalizable Organic Polymers: Coordination Polymers as Templates for Solid-State [2 + 2]-Photopolymerization. Chem. - Eur. J. 2017, 23, 273−277. (25) Tieke, B.; Wegner, G. Solid-State Polymerization of 1,4Disubstituted trans,trans-Butadienes in Perovskite-Type Layer Structures. Makromol. Chem., Rapid Commun. 1981, 2, 543−549.

(26) Tieke, B. Solid State Polymerization of Butadienes. Polymerization of Long Chain Derivatives of Sorbic and Muconic Acid. Colloid Polym. Sci. 1985, 263, 965−972. (27) Tieke, B.; Chapuis, G. Solid State Polymerization of Butadienes in Layer Structures. Mol. Cryst. Liq. Cryst. 1986, 137, 101−116. (28) Matsumoto, A.; Matsumura, T.; Aoki, S. Stereospecific Polymerisation of Diethyl (Z,Z)-Hexa-2,4-dienedioate in the Crystalline State. J. Chem. Soc., Chem. Commun. 1994, 1389−1390. (29) Matsumoto, A.; Tanaka, T.; Tsubouchi, T.; Tashiro, K.; Saragai, S.; Nakamoto, S. Crystal Engineering for Topochemical Polymerization of Muconic Esters Using Halogen-Halogen and CH/π Interactions as Weak Intermolecular Interactions. J. Am. Chem. Soc. 2002, 124, 8891−8902. (30) Matsumoto, A. Reactions of 1,3-Diene Compounds in the Crystalline State. Top. Curr. Chem. 2005, 254, 263−305. (31) Hoang, T.; Lauher, J. W.; Fowler, F. W. The Topochemical 1,6Polymerization of a Triene. J. Am. Chem. Soc. 2002, 124, 10656− 10657. (32) Xiao, J.; Yang, M.; Lauher, J. W.; Fowler, F. W. A Supramolecular Solution to a Long-Standing Problem: The 1,6Polymerization of a Triacetylene. Angew. Chem., Int. Ed. 2000, 39, 2132−2135. (33) Itoh, T.; Nomura, S.; Uno, T.; Kubo, M.; Sada, K.; Miyata, M. Topochemical Polymerization of 7,7,8,8-Tetrakis(methoxycarbonyl)quinodimethane. Angew. Chem., Int. Ed. 2002, 41, 4306−4309. (34) Itoh, T. Solution and Solid-State Polymerizations of Substituted p-Quinodimethanes and p-Quinone Methides. Polymer 2005, 46, 6998−7017. (35) Itoh, T.; Suzuki, T.; Uno, T.; Kubo, M.; Tohnai, N.; Miyata, M. cis-Specific Topochemical Polymerization: Alternating Copolymerization of 7,7,8,8-Tetrakis(methoxycarbonyl)quinodimethane with 7,7,8,8-Tetracyanoquinodimethane in the Solid State. Angew. Chem., Int. Ed. 2011, 50, 2253−2256. (36) Baughman, R. H.; Yee, K. C. Solid-State Polymerization of Linear and Cyclic Acetylenes. Macromol. Rev. 1978, 13, 219−239. (37) Enkelmann, V. Structural Aspects of the Topochemical Polymerization of Diacetylenes. In Polydiacetylenes: Adv. Polym. Sci.; Cantow, H.-J, Ed.; Springer-Verlag: Berlin, 1984; Vol. 63, pp 91−136. (38) Nakanishi, H.; Matsuda, H.; Okada, S.; Kato, M. Preparation and Nonlinear Optical Properties of Novel Polydiacetylenes. In Frontiers of Macromolecular Science; Saegusa, T., Higashimura, T., Abe, A., Eds.; Blackwell Scientific Publication: Oxford, 1989; pp 469−474. (39) Sarkar, A.; Okada, S.; Matsuzawa, H.; Matsuda, H.; Nakanishi, H. Novel Polydiacetylenes for Optical Materials: Beyond the Conventional Polydiacetylenes. J. Mater. Chem. 2000, 10, 819−828. (40) Patel, G. N.; Khanna, Y. P.; Ivory, D. M.; Sowa, J. M.; Chance, R. R. Polymer Conversions in the γ-Ray Polymerization of UrethaneSubstituted Diacetylenes. J. Polym. Sci., Polym. Phys. Ed. 1979, 17, 899−903. (41) Tachibana, H.; Kumai, R.; Hosaka, N.; Tokura, Y. Crystal Structures, Polymerization, and Thermochromic Phase Changes in Urethane-Substituted Diacetylenes Crystals with Varying Alkyl Chain Lengths. Chem. Mater. 2001, 13, 155−158. (42) Kim, W. H.; Kodali, N. B.; Kumar, J.; Tripathy, S. K. A Novel, Soluble Poly(diacetylene) Containing an Aromatic Substituent. Macromolecules 1994, 27, 1819−1824. (43) Sarkar, A.; Okada, S.; Nakanishi, H.; Matsuda, H. Polydiacetylenes from Asymmetrically Substituted Diacetylenes Containing Heteroaryl Side Groups for Third-Order Nonlinear Optical Properties. Macromolecules 1998, 31, 9174−9180. (44) Shibuya, T.; Tatewaki, Y.; Okada, S. Synthesis of Polydiacetylenes with Pyridyl Groups Directly Bound to the Main Chain. Mol. Cryst. Liq. Cryst. 2012, 568, 93−97. (45) Ikeshima, M.; Mamada, M.; Katagiri, H.; Minami, T.; Okada, S.; Tokito, S. Synthesis and Solid-State Polymerization of Diacetylene Derivatives with an N-Carbazolylphenyl Group. Bull. Chem. Soc. Jpn. 2015, 88, 843−849. (46) Ikeshima, M.; Mamada, M.; Minami, T.; Tokito, S.; Okada, S. Synthesis and Solid-State Polymerization of Diacetylene Derivatives I

DOI: 10.1021/acs.cgd.8b00815 Cryst. Growth Des. XXXX, XXX, XXX−XXX

Crystal Growth & Design

Article

Directly Substituted by a Phenylcarbazole Moiety. Polym. J. 2016, 48, 1013−1018. (47) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G. A.; Nakatsuji, H.; Caricato, M.; Li, X.; Hratchian, H. P.; Izmaylov, A. F.; Bloino, J.; Zheng, G.; Sonnenberg, J. L.; Hada, M.; Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Vreven, T.; Montgomery, J. A., Jr.; Peralta, J. E.; Ogliaro, F.; Bearpark, M.; Heyd, J. J.; Brothers, E.; Kudin, K. N.; Staroverov, V. N.; Keith, T.; Kobayashi, R.; Normand, J.; Raghavachari, K.; Rendell, A.; Burant, J. C.; Iyengar, S. S.; Tomasi, J.; Cossi, M.; Rega, J.; Millam, M.; Klene, M.; Knox, J. E.; Cross, J. B.; Bakken, V.; Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R. E.; Yazyev, O.; Austin, A. J.; Cammi, R.; Pomelli, C.; Ochterski, J. W.; Martin, R. L.; Morokuma, K.; Zakrzewski, V. G.; Voth, G. A.; Salvador, P.; Dannenberg, J. J.; Dapprich, S.; Daniels, A. D.; Farkas, O.; Foresman, J. B.; Ortiz, J. V.; Cioslowski, J.; Fox, D. J. Gaussian 09, revision D.01; Gaussian, Wallingford, CT, 2013. (48) Okada, S.; Matsuda, H.; Otsuka, M.; Nakanishi, H.; Kato, M. Synthesis and Solid-State Polymerization of ω-(4-Aryl-1,3-butadiynyl) Substituted 1-Alkanol and Alkanoic Acid. Bull. Chem. Soc. Jpn. 1994, 67, 483−489. (49) Ishii, S.; Kaneko, S.; Tatewaki, Y.; Okada, S. Synthesis and Solid-State Polymerization of 4-(Dimethylamino)phenylbutadiyne Derivatives and Their Charge-Transfer Complexes. Mol. Cryst. Liq. Cryst. 2013, 580, 64−68. (50) Okada, S.; Matsuda, H.; Masaki, A.; Nakanishi, H.; Hayamizu, K. Novel Linear and Ladder Polymers from Tetraynes for Nonlinear Optics. Proc. SPIE 1991, 25−34. (51) CrystalClear, Rigaku Corporation, Tokyo, Japan, 2008. (52) Sheldrick, G. M. Program for the Solution of Crystal Structures. Acta Crystallogr., Sect. A: Found. Adv. 2015, 71, 3−8. (53) Sheldrick, G. M. Program for the Refinement of Crystal Structures. Acta Crystallogr., Sect. C: Struct. Chem. 2015, 71, 3−8. (54) Wakita, K. Yadokari-XG, Software for Crystal Structure Analyses, 2001. (55) Kabuto, C.; Akine, S.; Nemoto, T.; Kwon, E. Release of Software (Yadokari-XG 2009) for Crystal Structure Analyses. Nippon Kessho Gakkaishi 2009, 51, 218−224. (56) (a) Wegner, G. Topochemical Polymerization of Monomers with Conjugated Triple Bonds. Makromol. Chem. 1972, 154, 35−48. (b) Nakanishi, H.; Matsuda, H.; Kato, M.; Theocharis, C. R.; Jones, W. Single-Crystal Study of the Solid-State Polymerisation of Butadiynylenebis-(m-acetamidobenzene). J. Chem. Soc., Perkin Trans. 2 1986, 1965−1967. (57) (a) Tokura, Y.; Koda, T.; Itsubo, A.; Miyabayashi, M.; Okuhara, K.; Ueda, A. Optical Spectra in Polydiacetylene Crystals Substituted with Fluorobenzenes. J. Chem. Phys. 1986, 85, 99−104. (b) Nakanishi, H.; Matsuda, H.; Okada, S.; Kato, M. Nonlinear Optical Properties of Polydiacetylenes with Conjugation between the Main Chain and the Substituents. In Nonlinear Optics of Organics and Semiconductors: Springer Proc. Phys.; Kobayashi, T., Ed.; Springer-Verlag: Berlin, 1989; Vol. 36, pp 155−162. (58) Okada, S.; Ohsugi, M.; Masaki, A.; Matsuda, H.; Takaragi, S.; Nakanishi, H. Prearation and Nonlinear Optical Property of Polydiacetylenes from Unsymmetrical Diphenylbutadiynes with Trifluoromethyl Substituents. Mol. Cryst. Liq. Cryst. 1990, 183, 81− 90. (59) Sato, H.; Okada, S.; Kimura, T.; Shimada, S.; Matsuda, H.; Nakanishi, H. Synthesis of Pyridylbutadiyne Derivatives and Improvement of Their Solid-State Polymerizability by Complexation with Alkanoic Acid. Chem. Lett. 2006, 35, 412−413. (60) Bloor, D.; Koski, L.; Stevens, G. C.; Preston, F. H.; Ando, D. J. Solid State Polymerization of Bis-(p-toluene sulphonate) of 2,4Hexadiyne-1,6-diol. Part 1 X-ray Diffraction and Spectroscopic Observations. J. Mater. Sci. 1975, 10, 1678−1688. (61) (a) Hood, R. J.; Müller, H.; Eckhardt, C. J.; Chance, R. R.; Yee, K. C. Optical Properties of a Polydiacetylene Crystal: Poly-[1,6-di(N-

carbazolyl)-2,4-hexadiyne]. Chem. Phys. Lett. 1978, 54, 295−299. (b) Apgar, P. A.; Yee, K. C. Poly[1,6-di(N-carbazolyl)-2,4-hexadiyne]. Acta Crystallogr., Sect. B: Struct. Crystallogr. Cryst. Chem. 1978, 34, 957−959. (62) (a) Chance, R. R.; Patel, G. N.; Witt, J. D. Thermal effects on the optical properties of single crystals and solution-cast films of urethane substituted polydiacetylenes. J. Chem. Phys. 1979, 71, 206− 211. (b) Brouty, C.; Spinat, P.; Whuler, A. Détermination Structurale du Poly{bis[(butoxycarbonylméthylcarbamoyloxy)-4 butyl]-1,2 butène-1 yne-3 ylène} [poly(4BCMU)] par Minimisation des Energies d’Interactions Moléculaires. Acta Crystallogr., Sect. C: Cryst. Struct. Commun. 1988, 44, 1070−1076. (63) (a) Enkelmann, V.; Schleier, G. Poly[1,2-bis(diphenylaminomethyl)-l-buten-3-ynylene]. Acta Crystallogr., Sect. B: Struct. Crystallogr. Cryst. Chem. 1980, 36, 1954−1956. (b) Morrow, M. E.; White, K. M.; Eckhardt, C. J.; Sandman, D. J. Single-Crystal Reflection and Raman Spectra of the Polydiacetylene THD. Chem. Phys. Lett. 1987, 140, 263−269. (64) (a) Wenz, G.; Wegner, G. Molecular Weight Distribution and Solution Properties of a Poly(diacetylene). Makromol. Chem., Rapid Commun. 1982, 3, 231−237. (b) Wenz, G.; Müller, M. A.; Schmidt, M.; Wegner, G. Structure of Poly(diacetylenes) in Solution. Macromolecules 1984, 17, 837−850. (65) Okada, S.; Hayamizu, K.; Matsuda, H.; Masaki, A.; Nakanishi, H. Structures of the Polymers Obtained by the Solid-State Polymerization of Diyne, Triyne, and Tetrayne with Long Alkyl Substituents. Bull. Chem. Soc. Jpn. 1991, 64, 857−863. (66) (a) Sandstedt, C. A.; Eckhardt, C. J.; Downey, M. J.; Sandman, D. J. Optical, Calorimetric, and Mass Spectroscopic Study of Nonthermochromic Crystalline Forms of the Polydiacetylene Bis(ethylurethane)-5,7-Dodecadiyne-1,12-diol (ETCD). Chem. Mater. 1994, 6, 1346−1350. (b) Wang, X.; Sandman, D. J.; Enkelmann, V.; Chen, C.-H.; Foxman, B. M. Crystal and Molecular Structure of the bis-Ethyl Urethane of 5,7-Dodecadiyn-1,12,diol (ETCD). J. Macromol. Sci., Part A: Pure Appl.Chem. 2008, 45, 914−916. (67) Pigos, J. M.; Zhu, Z.; Musfeldt, J. L. Optical Properties of a Supramolecular Assembly Containing Polydiacetylene. Chem. Mater. 1999, 11, 3275−3278. (68) Galindo, C.; Jacques, P.; Kalt, A. Photodegradation of the Aminoazobenzene Acid Orange 52 by Three Advanced Oxidation Processes: UV/H2O2, UV/TiO2 and VIS/TiO2: Comparative Mechanistic and Kinetic Investigations. J. Photochem. Photobiol., A 2000, 130, 35−47. (69) Al-Nu’airat, J.; Altarawneh, M.; Gao, X.; Westmoreland, P. R.; Dlugogorski, B. Z. Reaction of Aniline with Singlet Oxygen (O2 1Δg). J. Phys. Chem. A 2017, 121, 3199−3206.

J

DOI: 10.1021/acs.cgd.8b00815 Cryst. Growth Des. XXXX, XXX, XXX−XXX