Zhe
Wang
a,
Rikuo
Akisaka
a,
Sohshi
Yabumoto
b,
Tatsuo
Nakagawa
b,
Sayaka
Hatano
a and
Manabu
Abe
*ac
aDepartment of Chemistry, Graduate School of Science, Hiroshima University, 1-3-1 Kagamiyama, Higashi-Hiroshima, Hiroshima 739-8526, Japan. E-mail: mabe@hiroshima-u.ac.jp
bUnisoku Co., Ltd., 2-4-3 Kasugano, Hirakata, Osaka 573-0131, Japan
cHiroshima University Research Centre for Photo-Drug-Delivery-Systems (HiU-P-DDS), Hiroshima University, 1-3-1 Kagamiyama, Higashi-Hiroshima, Hiroshima 739-8526, Japan
First published on 10th November 2020
Localised singlet diradicals are key intermediates in bond homolysis processes. Generally, these highly reactive species undergo radical–radical coupling reaction immediately after their generation. Therefore, their short-lived character hampers experimental investigations of their nature. In this study, we implemented the new concept of “stretch effect” to access a kinetically stabilised singlet diradicaloid. To this end, a macrocyclic structure was computationally designed to enable the experimental examination of a singlet diradicaloid with π-single bonding character. The kinetically stabilised diradicaloid exhibited a low carbon–carbon coupling reaction rate of 6.4 × 103 s−1 (155.9 μs), approximately 11 and 1000 times slower than those of the first generation of macrocyclic system (7.0 × 104 s−1, 14.2 μs) and the parent system lacking the macrocycle (5 × 106 s−1, 200 ns) at 293 K in benzene, respectively. In addition, a significant dynamic solvent effect was observed for the first time in intramolecular radical–radical coupling reactions in viscous solvents such as glycerin triacetate. This theoretical and experimental study demonstrates that the stretch effect and solvent viscosity play important roles in retarding the σ-bond formation process, thus enabling a thorough examination of the nature of the singlet diradicaloid and paving the way toward a deeper understanding of reactive intermediates.
Regarding localised diradicals, highly reactive species that undergo fast radical–radical coupling reaction, the low-temperature matrix isolation of diradical T-DR1 was first achieved in 1975 (Scheme 1a). The isolation allowed a detailed investigation of the ground-state spin multiplicity and reactivity of this species, resulting in the elucidation of its triplet ground state and its heavy-atom tunnelling reaction.45–47 Furthermore, carbon–carbon singlet diradical S-DR2a (τ293 = 80 ns in n-pentane), in which electron-withdrawing groups (EWGs) lower the energy of the singlet state with regard to that of the triplet state, was first detected in 1998,48 whereas the longer-lived S-DR2b (τ293 = 209 ns in benzene) featuring flexible alkoxy groups has been studied in our laboratory.49–52
Localised singlet diradicaloids can be classified as Type-1 or Type-2 according to their most stable electronic configuration, which in turn depends on the relative HOMO and LUMO energy levels (Scheme 1).53,54 Hence, the π-single bonding character (C–π–C) characterises Type-1 molecules because the bonding orbital ψS is HOMO.55 Recently, long-lived singlet diradicaloids S-DR2c (τ293 = 23.8 μs in benzene) and S-DR2d (τ293 = 9.67 ms in toluene), featuring the bulky-substituent and nitrogen-atom effects, respectively, were observed at 293 K.56,57 Additionally, several heavy-atom analogues S-DR2e–k, including Type-2 singlet diradicaloids, have been isolated (Scheme 1).58–64 Very recently, five- and four-membered cyclic silicon analogues of π-single bonded species were reported.58,65,66
In 2012, macrocyclic structures were designed to kinetically stabilise carbon–carbon singlet diradicaloids (Scheme 2).67 In these scaffolds, structural rigidity precludes the σ bond formation between the radical centres, a phenomenon termed “stretch effect” (Scheme 2a). Recently, this effect was studied using macrocyclic singlet diradicaloid S-DR3a. The moderate increase in the lifetime of S-DR3a to τ293 = 14.2 μs in benzene (Scheme 2b) indicated that the construction of macrocyclic structures is a useful strategy to extend the lifetime of singlet diradicaloids and enable more detailed investigations.68 This finding prompted us to devise a new fine-tuned macrocyclic structure. In this study, singlet diradicaloid S-DR3b featuring a naphthalene-containing macrocyclic system was designed and examined by computational and experimental studies. In addition, its reactivity toward radical–radical coupling, the features of this reaction, and the properties of its products were investigated in detail.
Scheme 2 (a) Stretch effect induced by the macrocycle. (b) Localised diradicaloids investigated in this study. aLifetime values τ293 were determined in benzene. |
Fig. 1 (a) HOMO (ψS) and (b) LUMO (ψA) orbitals and their occupation numbers (n) calculated for S-DR3b at the CASSCF(2,2)/6-31G(d) level of theory. |
The kinetic stabilisation of S-DR3b by the macrocyclic structure was evaluated by comparing its computational results with those of S-DR2b at the same level of theory (Table 1, entries 1–3). Calculations of the ring-closed products CP were performed using the restricted method at the (R)ωB97X-D/6-31G(d) level of theory, whereas the corresponding transition states TS were assessed by computing the vibrational frequency and their intrinsic reaction coordinate (IRC, Fig. S17 in ESI†). According to the IRC calculations, the transition states of the ring-closing reactions towards cis- and trans-TS3b produced the metastable ring-closed conformers par-CP3b with a face-to-face orientation of the benzene rings (par-CP structure, Table 1). The barrierless par-twi isomerisation afforded the more stable conformers cis-twi- and trans-twi-CP3b with a nearly perpendicular orientation of the phenyl residues (twi-CP structure, Table 1). As expected, the energies of the ring-closed products cis- and trans-twi-CP3b were 35.15 and 37.00 kJ mol−1 higher (ΔEDR–CP) than those of cis- and trans-twi-CP2b, respectively (entries 1 and 3). The corresponding energy differences with CP3a were found to be 19.99 and 21.29 kJ mol−1 (entries 2 and 3). Thus, the difference between the energies of the singlet diradical and ring-closed product significantly decreased upon introduction of the macrocyclic structure in 3b. Moreover, the small energy difference between closed-shell cis-twi-CP3b and S-DR3b (ΔEDR–CP = 3.99 kJ mol−1) suggests a significant contribution of the stretch effect to the increase in the molecular strain of CP3b, thus kinetically stabilising S-DR3b. Additionally, the stretch effect is reflected in the longer C–C bonds calculated for CP3b (Table 1, entries 1–3, values in brackets). Furthermore, this effect is significant in the parallel conformation of par-CP. The transition state enthalpies of cis- and trans-TS3b were larger than those of TS2b and TS3a by 5.91 and 8.60, 3.20 and 7.71 kJ mol−1, respectively. According to the computational analyses, the stretch effect in the newly designed macrocyclic structure is expected to provide a long-lived singlet diradical.
Entry | Functions | S-DR | T-DR | cis-TS | cis-par-CP | cis-twi-CP | trans-TS | trans-par-CP | trans-twi-CP | |
---|---|---|---|---|---|---|---|---|---|---|
1 | ωB97X-D | 2b | 0.00 [2.380] | 7.68 [2.399] | 51.37 [2.063] | 1.66 [1.594] | −39.14 [1.585] | 68.92 [2.054] | −26.74 [1.576] | −63.66 [1.566] |
2 | 3a | 0.00 [2.379] | 7.58 [2.399] | 54.08 [2.047] | 7.67 [1.600] | −23.98 [1.584] | 69.81 [2.058] | −24.53 [1.574] | −47.95 [1.565] | |
3 | 3b | 0.00 [2.381] | 7.49 [2.400] | 57.28 [2.041] | 24.67 [1.622] | −3.99 [1.590] | 77.52 [2.112] | 1.86 [1.587] | −26.66 [1.567] | |
4 | B3LYP | 2b | 0.00 [2.389] | 9.43 [2.413] | 44.23 [2.070] | 24.99 [1.612] | −16.66 [1.608] | 65.41 [2.016] | 1.48 [1.589] | −37.88 [1.582] |
5 | 3a | 0.00 [2.389] | 9.10 [2.413] | 50.35 [1.989] | 29.48 [1.630] | −0.30 [1.610] | 66.58 [2.053] | 1.82 [1.592] | −20.06 [1.583] | |
6 | 3b | 0.00 [2.391] | 9.08 [2.414] | 55.29 [1.940] | 53.25 [1.681] | 17.56 [1.623] | 76.21 [2.116] | 36.81 [1.614] | 8.83 [1.585] | |
7 | CAM-B3LYP | 2b | 0.00 [2.382] | 7.22 [2.401] | 56.88 [2.030] | 17.51 [1.592] | −25.96 [1.586] | 76.37 [2.022] | −7.62 [1.572] | −48.20 [1.566] |
8 | 3a | 0.00 [2.383] | 7.00 [2.401] | 61.82 [2.011] | 20.99 [1.601] | −9.47 [1.587] | 77.38 [2.027] | −8.72 [1.574] | −30.93 [1.566] | |
9 | 3b | 0.00 [2.384] | 6.98 [2.402] | 66.45 [2.004] | 46.07 [1.631] | 9.49 [1.596] | 84.55 [2.082] | 26.08 [1.591] | −10.52 [1.570] | |
10 | M06-2x | 2b | 0.00 [2.372] | 7.93 [2.393] | 40.22 [2.132] | −14.71 [1.590] | −59.63 [1.582] | 56.30 [2.131] | −42.02 [1.576] | −81.38 [1.563] |
11 | 3a | 0.00 [2.372] | 7.70 [2.392] | 42.08 [2.122] | −8.75 [1.597] | −43.69 [1.581] | 58.27 [2.135] | −40.17 [1.572] | −65.51 [1.562] | |
12 | 3b | 0.00 [2.373] | 7.72 [2.394] | 45.00 [2.118] | 6.95 [1.620] | −23.53 [1.588] | 68.41 [2.126] | −14.36 [1.583] | −30.43 [1.564] | |
13 | ωB97 | 2b | 0.00 [2.386] | 6.25 [2.402] | 62.27 [2.029] | −6.21 [1.576] | −49.45 [1.569] | 80.40 [2.023] | −33.52 [1.560] | −73.87 [1.555] |
14 | 3a | 0.00 [2.386] | 6.20 [2.402] | 67.11 [2.013] | −1.36 [1.581] | −33.51 [1.570] | 81.94 [2.024] | −33.54 [1.562] | −56.65 [1.554] | |
15 | 3b | 0.00 [2.387] | 6.13 [2.403] | 71.62 [2.007] | 21.30 [1.598] | −13.57 [1.575] | 87.36 [2.075] | −2.49 [1.572] | −36.07 [1.557] | |
16 | APF-D | 2b | 0.00 [2.371] | 8.95 [2.393] | 40.57 [2.070] | 5.00 [1.593] | −37.86 [1.588] | 60.70 [2.060] | −20.17 [1.576] | −59.76 [1.568] |
17 | 3a | 0.00 [2.370] | 8.67 [2.392] | 42.66 [2.054] | 11.92 [1.601] | −23.08 [1.586] | 60.33 [2.067] | −19.09 [1.575] | −45.06 [1.566] | |
18 | 3b | 0.00 [2.371] | 8.70 [2.393] | 45.19 [2.047] | 26.61 [1.627] | −3.09 [1.591] | 69.16 [2.095] | 7.02 [1.587] | −24.57 [1.567] |
Computations were also conducted at the B3LYP,75 CAM-B3LYP,76 M06-2x,77 ωB97,78 and APF-D79 functions with the 6-31G(d) basis set. Although the relative energies computed by distinct methods were different, the general tendencies corroborate the stretch effect. For example, cis- and trans-twi-CP3b were higher in energy than S-DR3b by the B3LYP method (entry 6), whereas the data obtained by other computational methods indicates that the twisted ring-closed compounds are more stable than S-DR3b. The energy differences between S-DR3b and CP3b were much smaller than between S-DR2b,3a and CP2b,3a owing to the macrocyclic structure. Notably, the difference in the enthalpy between S-DR2b and cis-TS2b computed using the M06-2x and APF-D methods (ΔHrel = 40.22 and 40.57 kJ mol−1, entries 10 and 16 respectively) were closest to the experimental activation energy values for the reaction of S-DR2b to cis-CP2b (Ea = 30.5 ± 0.4 kJ mol−1).49
To gain a deeper understanding of the effect of the designed macrocyclic skeleton, the geometry of the triple bonds and naphthyl moiety were analysed (Fig. 2).100 The triple bonds in S-DR3b were slightly bent to 178° and 174°, whereas the naphthyl moiety deviated from planarity by 2.4° (Fig. 2a). Larger values were obtained for cis-twi-CP3b, in which the bending angles of the triple bonds and naphthyl moiety were 162° and 5.2°, respectively (Fig. 2b). A similar bent structure was also confirmed for trans-twi-CP3b (Fig. 2c). Subsequently, the effect of bending on the molecular strain was assessed by computing the strain energies (SEM) of the macrocyclic units in AZ3b, S-DR3b, cis-CP3b, and trans-CP3b at the (R,U)ωB97X-D/6-31G(d) level of theory (Table 2a), since the structural parameters computed by this method were well aligned with the experimental data (X-ray crystallography, Table S2 in ESI†). The corresponding values were compared with those in 3a. The value of SEM was calculated by subtracting the total electronic energy of the non-strained macrocyclic structure (nS), 2,7-bis(3-(phenylethynyl)-phenyl)naphthalene, from the energy of strained macrocycles in AZ-S, DR-S, cis-CP-S, and trans-CP-S. The latter were obtained by replacing the azo, diradical, and ring-closing units in AZ3a,b, S-DR3a,b, cis-CP3a,b, and trans-CP3a,b with two hydrogen atoms (Table 2a, example for 3b). Their energies were obtained by partial optimisation of the C–H bonds without optimising other moieties. The strain energies in AZ3a,b and S-DR3a,b were relatively small (entries 1–4, SEM = 8.01 and 11.49 for 3a; 11.24 and 16.38 kJ mol−1 for 3b, respectively), whereas larger values were obtained for CP3a,b with bent alkynes (entries 5–12, approximately 15–19 for 3a; 29–38 for 3b kJ mol−1). Thus, the strain energy of 3b was found to be larger than that of 3a.
Fig. 2 Bending angles of triple bonds and naphthyl moieties in (a) S-DR3b, (b) cis-twi-CP3b, and (c) trans-twi-CP3b, optimised at the (R,U)ωB97X-D/6-31G(d) level of theory. |
Entry | Compounds | Energies in kJ mol−1 | ||
---|---|---|---|---|
SEM | SE | |||
a Values relative to S-DR3a. b Values relative to S-DR3b. | ||||
1 | AZ | 3a | 8.01 | 9.25 |
2 | 3b | 11.24 | 17.45 | |
3 | S-DR | 3a | 11.49 | 5.11 (0.00)a |
4 | 3b | 16.38 | 8.92 (0.00)b | |
5 | cis-par-CP | 3a | 19.19 | 11.47 (6.36)a |
6 | 3b | 29.13 | 32.18 (23.26)b | |
7 | cis-twi-CP | 3a | 14.62 | 19.63 (14.52)a |
8 | 3b | 36.26 | 43.66 (34.74)b | |
9 | trans-par-CP | 3a | 21.45 | 6.28 (1.17)a |
10 | 3b | 34.85 | 36.18 (27.26)b | |
11 | trans-twi-CP | 3a | 15.31 | 19.97 (14.87)a |
12 | 3b | 38.56 | 45.08 (36.16)b |
The molecular strain energies (SE) of AZ3a,b, S-DR3a,b, cis-CP3a,b, and trans-CP3a,b, which were estimated using the isodesmic reaction (Table 2b, example for 3b) and compared to the standard AZ2b, S-DR2b, cis-CP2b, and trans-CP2b, were larger than the corresponding SEM, with the exception of S-DR3a,b (entries 3 and 4). The strain energies relative to S-DR3a,b, which are indicated in parenthesis in Table 2 (entries 3–12), were very similar to the differences between the corresponding ΔEDR–CP of 2b and 3a,b (ΔΔEDR–CP = 6.01, 15.16, 2.21, and 15.71 kJ mol−1 for 3a; 23.01, 35.15, 28.60, and 37.00 kJ mol−1 for 3b, respectively, Table 1, entries 1–3), indicating that the molecular strain strongly correlates with the macrocyclic structures. The molecular strain of 3b was larger than that of 3a. Furthermore, the computations clearly indicate that the kinetic stabilisation of S-DR3b by the macrocyclic scaffold suppresses bond formation in the singlet diradicaloid.
The molecular structure of AZ3b was confirmed by nuclear magnetic resonance spectroscopy (1H, 13C NMR) and ESI mass spectrometry (MS). The endo configuration was confirmed by the observed correlation between a methoxy group and the bridgehead protons in the two-dimensional nuclear Overhauser effect spectrum (NOESY, Scheme 3). Furthermore, X-ray crystallographic analysis of AZ3b corroborated the molecular structure, although disorder was observed for the octyl chains (Fig. S7 in ESI†). The UV-vis absorbance spectrum of AZ3b exhibits a maximum absorption at 356 nm (ε356 = 517 dm3 mol−1 cm−1), which is analogous to that of AZ2b (ε358 = 112 dm3 mol−1 cm−1). Hence, it stems from the overlap of the n–π* electronic transition of the azo chromophore with the π–π* one of the π-conjugated system, whereas the broad absorption band up to 450 nm corresponds to the π-conjugation in the macrocycle (Fig. 3).
Furthermore, trans-CP3b, which was computed to be energetically more stable than cis-CP3b (Table 1), was formed by photolysis of AZ3b at ∼25 °C (Fig. 4), although the calculated energy barrier for the formation of cis-CP3b was lower than that of trans-CP3b (Table 1). As the isomerisation of cis-CP3b to trans-CP3b is supposedly inhibited by a large activation energy (>70 kJ mol−1 from S-DR3b to trans-CP3b at 199 K, Table 1), low-temperature 1H NMR experiments were conducted to identify the primary product of the reaction at 199 K. To this end, the photolysis of AZ3b was carried out in degassed toluene-d8 (6.49 mM) under irradiation with a Nd:YAG laser (30 mJ per pulse, 355 nm), which was introduced into the NMR tube by a quartz rod.81In situ1H NMR monitoring of the reaction revealed the sole formation of trans-CP3b (vinylic signals c and d, Fig. 5a) alongside unreacted AZ3b (signals a and b). The exclusive formation of trans-CP3b is explained by the existence of the puckered diradical puc-1DR3b (path A, Scheme 5).82 Using the same experimental setup, a degassed toluene-d8 solution of AZ3b (0.60 mM, Abs355 = 0.32) and triplet sensitiser benzophenone (9.32 mM, Abs355 = 1.09) was irradiated (355 nm, 199 K). Interestingly, the NMR spectra acquired in situ contained new signals (e and f, approximately 5.9 ppm, Fig. 5b), which were converted to signals c and d in the dark. Hence, signals e and f correspond to cis-CP3b, which subsequently isomerises to the more stable trans-CP3b. The mechanism of the benzophenone-sensitised cis-CP3b formation involves the planar diradical intermediate pl-1DR3b, which is associated with a smaller activation energy (path B, Scheme 5).
Fig. 6 Low-temperature UV-vis absorption spectra of the photolysis of AZ3b (2.46 mM) in a degassed MTHF matrix at 90 K. |
To confirm the spin multiplicities of the species associated with the absorption bands at 460 and 580 nm, low-temperature electron paramagnetic resonance (EPR) spectroscopy was conducted during the photochemical reaction of AZ3b (4.92 mM) in an MTHF matrix irradiated with a Hg lamp (λexc > 250 nm) at 80 K (Fig. 7). EPR signals typical of triplet species were observed during photolysis at 2331 (z1), 2507 (y1), 3097 (x1), 3509 (x2), 4154 (y2), and 4375 Gauss (G) (z2) corresponding to the allowed transition (|Δms| = 1) at 9.4 GHz resonance frequency (Fig. 7a). In addition, the half-field signal (|Δms| = 2) was detected at 1571 G. The signals at approximately 3400 G correspond to doublet impurities, which were also observed in a control experiment (MTHF irradiation under the same conditions). The obtained zero-field splitting (zfs) parameters of the triplet species were D/hc = 0.096 cm−1 and E/hc = 0.019 cm−1. The EPR spectrum simulated with D/hc = 0.096 cm−1 and E/hc = 0.019 cm−1 (Fig. 7b) reproduced well the experimental spectrum. The obtained zfs parameters were similar to those of the triplet excited state of 2-phenylnaphthalene (D/hc = 0.0963 cm−1 and E/hc = 0.0274 cm−1).84 The D value of triplet 1,3-diphenyl-cyclopentane-1,3-diyl diradicals was reported to be approximately 0.05 cm−1,16,85 which is much smaller than that observed in this study. When the irradiation was ceased at 80 K, the triplet signals disappeared (Fig. 7c), confirming that singlet diradicaloid S-DR3b is associated with the UV-vis absorbance at 580 nm (Fig. 6). Further, the triplet signals were short-lived at 5 K and the lifetimes of the decay signal monitored at 1562 G were nearly the same in the temperature range of 5–80 K (τ5 = 2.20 s and τ80 = 2.25 s, Fig. S11 in ESI†). The temperature-independency of the decay process and the large zfs parameters support the hypothesis that these EPR signals correspond to the triplet excited state of the naphthyl unit in AZ3b. Thus, the UV-vis absorption band at 460 nm was assigned to the T–T absorption of the naphthyl moiety (Fig. 6). Indeed, the T–T absorption band of a model compound 3b′ was predicted to appear at 466 nm using the TD-DFT method at the ωB97X-D/6-31G(d) level of theory (Fig. S18 in ESI†).
To further confirm the assignment of the EPR triplet signals, the zfs parameters, D/hc and E/hc, were computed for the triplet states of DR2b, DR3a, and DR3b at the B3LYP/EPR-II86 level using the ORCA 4.2.1 program package87,88 (Table 3). To evaluate the accuracy of computed values, the experimentally known zfs values of triplet molecules DR4–685 were also simulated at the same level of theory (entries 4–6). As shown in entries 4–6, the calculated (cald) zfs parameters, especially D/hc values, well reproduced the experimental values of triplet diradicals DR4–6. Thus, D/hc value of triplet state DR3b should be around 0.057 (entry 3), which is not consistent with the relatively large D/hc value of 0.096 cm−1 in the photolysis of AZ3b (Fig. 7a).
Entry | zfs parameters (in cm−1) exp (calcd)a | ||
---|---|---|---|
D/hc | E/hc | ||
a Calculated at the B3LYP/EPR-II level of theory using ORCA 4.2.1 program package. | |||
1 | DR2b | nd (0.061) | nd (0.0051) |
2 | DR3a | nd (0.059) | nd (0.0064) |
3 | DR3b | nd (0.057) | nd (0.0060) |
4 | DR4 | 0.084 (0.089) | 0.0020 (0.0019) |
5 | DR5 | 0.112 (0.113) | 0.0050 (0.0023) |
6 | DR6 | 0.045 (0.044) | 0.0010 (0.0020) |
Variable temperature laser flash photolysis (VT-LFP) measurements were conducted at five temperatures in the range of 273–303 K. The activation parameters Ea, logA, ΔH‡, ΔS‡, and ΔG‡293 of the ring-closing reaction of S-DR3b to CP3b in benzene were determined by the Arrhenius and Eyring plots (Table 4, Fig. S15 in ESI†). The activation energy and enthalpy of this process, determined as 58.4 ± 1.1 and 56.0 ± 1.1 kJ mol−1, respectively, are approximately 28 and 6 kJ mol−1 higher than the values of the corresponding reactions of S-DR2b and S-DR3a, respectively (entry 3). Unlike the ring-closing process of S-DR2b, the corresponding reactions of S-DR3a,b are associated with positive activation entropies (entries 2,3), although the decay process (i.e. the intramolecular σ-bond formation event), is the same, suggesting that the transition states of theses reactions should be very similar.68 Indeed, the activation entropy of the ring-closing reaction of S-DR3b computed at the (U)ωB97X-D/6-31G(d) level of theory was −22.04 J mol−1 K−1. This unusual observation prompted us to investigate the effect of the solvent on the lifetime in more detail.
Entry | S-DR | τ 293/μs | E a/kJ mol−1 | logA/s−1 | ΔH‡/kJ mol−1 | ΔS‡/J mol−1 K−1 | ΔG‡293/kJ mol−1 |
---|---|---|---|---|---|---|---|
1 | 2b | 0.21 ± 0.01 | 30.5 ± 0.4 | 12.1 ± 0.1 | 28.0 ± 0.4 | −21.5 ± 0.8 | 34.2 ± 0.8 |
2 | 3a | 14.2 ± 0.8 | 52.3 ± 0.4 | 14.1 ± 0.1 | 49.7 ± 0.4 | 17.1 ± 1.2 | 44.7 ± 0.4 |
3 | 3b | 155.9 ± 3.3 | 58.4 ± 1.1 | 14.2 ± 0.2 | 56.0 ± 1.1 | 18.1 ± 2.3 | 50.7 ± 1.1 |
Entry | Solvent | π*/kcal mol−1 | η (20 °C)/cP | τ 293 of S-DR2b/ns | τ 293 of S-DR3b/μs |
---|---|---|---|---|---|
1 | n-Hexane | −0.11 (1) | 0.31 (2) | 90.1 (1) | nd |
2 | Tetrachloride carbon | 0.21 (2) | 0.97 (10) | 187.1 (5) | 17.2 (1) |
3 | Diethyl ether | 0.24 (3) | 0.24 (1) | 136.3 (2) | 46.8 (5) |
4 | Ethyl acetate | 0.45 (4) | 0.46 (5) | 182.6 (4) | 73.1 (8) |
5 | Toluene | 0.49 (5) | 0.59 (7) | 170.4 (3) | 116.5 (10) |
6 | 1,4-Dioxane | 0.49 (5) | 1.18 (11) | 250.6 (8) | 61.4 (6) |
7 | Benzene | 0.55 (7) | 0.65 (8) | 210.0 (6) | 155.9 (11) |
8 | Acetone | 0.62 (8) | 0.32 (3) | 231.4 (7) | 27.9 (3) |
9 | Glycerin triacetate | 0.63 (9) | 23.00 (13) | 517.1 (13) | 400.2 (12) |
10 | Chloroform | 0.69 (10) | 0.58 (6) | 404.8 (12) | 65.4 (7) |
11 | Dichloromethane | 0.73 (11) | 0.44 (4) | 294.0 (9) | 46.6 (4) |
12 | 1,2-Dichloroethane | 0.73 (11) | 0.79 (9) | 307.6 (10) | 22.8 (2) |
13 | Dimethyl sulfoxide | 1.00 (13) | 2.24 (12) | 393.5 (11) | 95.1 (9) |
As the radical–radical coupling reaction strongly depends on the solvent viscosity, the dynamic solvent effect should play an important role in the isomerisation of S-DR to CP.95–98,101 This effect can be expressed by eqn (1), where R, I, and P are the reactant, intermediate, and product, respectively. In a low-viscosity solvent, the conversion of I to P is the rate-limiting step (kf ≫ k1) according to the transition state theory (TST). Thus, the observed rate constant (kobs) is nearly equal to k1.
(1) |
In high-viscosity solvents, the reaction rate kobs is limited by solvent thermal fluctuations, rendering the TST no longer valid. Thus, the observed rate constant kobs can be expressed by eqn (2):
1/kobs = 1/kTST + 1/kf, | (2) |
Acetone (Ac, π* = 0.62 kcal mol−1, η = 0.32 cP) and GTA (π* = 0.63 kcal mol−1, η = 23.0 cP) are equally polar but very differently viscous. We assumed that the solvent thermal fluctuations in acetone are sufficiently fast to render the solvent dynamic effect due to the low viscosity negligible, such that kTST ≈ kAc and kobs ≈ kGTA. Thus, the rate constant for the solvent thermal fluctuation kf can be estimated by kf = (1/kGTA − 1/kAc)−1. The strong linear correlation between kf and the viscosity of GTA99 proved the validity of the solvent dynamic effect for the singlet diradical system (Fig. 9).
Fig. 9 Dependence of the solvent thermal fluctuation rate constant logkf calculated by eqn (2) on the viscosity of GTA. kAc and kGTA were calculated from the Eyring plot. |
Furthermore, the correlations between logkCP and the polarity and viscosity are shown in Fig. 10, where kCP (= kd = 1/τ293 for S-DR3b) is the rate constant of the radical–radical coupling process. Regarding S-DR2b, a good correlation was observed between the solvent polarity and the lifetime, although the high viscosity of GTA led to a large deviation from the linear correlation (Fig. 10a). However, in the case of S-DR3b, obvious correlations between logkCP and the polarity or viscosity of the solvents were not observed (Fig. 10c and d). Nevertheless, the effect of the viscosity on the lifetime suggests that the dynamic solvent effect should be considered to understand the phenomena.
Fig. 10 Correlations between the rate constant logkCP and the solvent polarity (π*) and viscosity (η) for (a and b) S-DR2b and (c and d) S-DR3b. |
The effect of solvent polarity (π*) and viscosity (η) on the lifetime of the singlet diradicals was further examined by performing a regression analysis according to eqn (3), in which A and B are the polarity and viscosity coefficients, respectively, and C is a constant term. All terms are compound-dependent.
τ = Aπ* + Bη + C | (3) |
Table 6 lists the calculated coefficients for S-DR2b and S-DR3b. The polarity of the solvent is the dominant factor determining the lifetime of S-DR2b, as the corresponding coefficient is much larger than the viscosity one (278.43 and 11.06, respectively). In contrast, the coefficients A and B are similar for S-DR3b, suggesting that both polarity and viscosity strongly influence its lifetime (Table 6). The regression analyses were validated by plotting the experimental lifetime values τ293 was plotted against the ones predicted by eqn (3) (Fig. 11). A good linear correlation is observed in both cases (R2 = 0.86), despite the slight deviations of the data points corresponding to benzene and toluene (S-DR3b) and chloroform (S-DR2b). To gain insight into the effect of the macrocyclic structure on the relationship between viscosity and lifetime, the molecular volumes of S-DR2b and S-DR3b were computed at the (U)ωB97X-D/6-31G(d) level of theory. The obtained values of 497.04 and 303.44 cm3 mol−1 for S-DR3b and S-DR2b, respectively, indicate that the solvent viscosity effect is more pronounced in the ring-closing of S-DR3b to cis-CP3b than in the corresponding reaction of S-DR2b.
Coefficient | S-DR2b | S-DR3b |
---|---|---|
A | 278.43 | 15985.07 |
B | 11.06 | 14940.91 |
C | 88.69 | 45794.25 |
Fig. 11 Correlation between the experimental and predicted lifetime values τ293 for (a) S-DR2b and (b) S-DR3b. |
Footnote |
† Electronic supplementary information (ESI) available: Full experimental details including synthetic procedures, characterisation data (NMR, MS, X-ray structure) and additional spectroscopic data; in situ NMR analyses; time profile of low temperature EPR measurements; details for LFP measurements including Arrhenius and Eyring plots; and computational details. See DOI: 10.1039/d0sc05311b |
This journal is © The Royal Society of Chemistry 2021 |