Chen
Yang
ab,
Faisal
Mehmood
a,
Tsz Lung
Lam
cd,
Sharon Lai-Fung
Chan
*cd,
Yuan
Wu
a,
Chi-Shun
Yeung
a,
Xiangguo
Guan
a,
Kai
Li
ab,
Clive Yik-Sham
Chung
a,
Cong-Ying
Zhou
ab,
Taotao
Zou
a and
Chi-Ming
Che
*ab
aState Key Laboratory of Synthetic Chemistry, Institute of Molecular Functional Materials, HKU-CAS Joint Laboratory on New Materials and Department of Chemistry, The University of Hong Kong, Pokfulam Road, Hong Kong, China. E-mail: cmche@hku.hk
bHKU Shenzhen Institute of Research and Innovation, Shenzhen, China
cThe Hong Kong Polytechnic University Shenzhen Research Institute, Shenzhen, PR China
dDepartment of Applied Biology and Chemical Technology, The Hong Kong Polytechnic University, Hung Hom, Hong Kong, China. E-mail: sharonlf.chan@polyu.edu.hk
First published on 20th January 2016
A new class of cyclometalated Ir(III) complexes supported by various bidentate C-deprotonated (C^N) and cis-chelating bis(N-heterocyclic carbene) (bis-NHC) ligands has been synthesized. These complexes display strong emission in deaerated solutions at room temperature with photoluminescence quantum yields up to 89% and emission lifetimes up to 96 μs. A photo-stable complex containing C-deprotonated fluorenyl-substituted C^N shows no significant decomposition even upon irradiation for over 120 h by blue LEDs (12 W). These, together with the strong absorption in the visible region and rich photo-redox properties, allow the bis-NHC Ir(III) complexes to act as good photo-catalysts for reductive C–C bond formation from C(sp3/sp2)–Br bonds cleavage using visible-light irradiation (λ > 440 nm). A water-soluble complex with a glucose-functionalized bis-NHC ligand catalysed a visible-light-driven radical cyclization for the synthesis of pyrrolidine in aqueous media. Also, the bis-NHC Ir(III) complex in combination with a cobalt catalyst can catalyse the visible-light-driven CO2 reduction with excellent turnover numbers (>2400) and selectivity (CO over H2 in gas phase: >95%). Additionally, this series of bis-NHC Ir(III) complexes are found to localize in and stain endoplasmic reticulum (ER) of various cell lines with high selectivity, and exhibit high cytotoxicity towards cancer cells, revealing their potential uses as bioimaging and/or anti-cancer agents.
In 2008, Yoon and co-workers40 reported [2 + 2] enone cycloadditions, and MacMillan and co-workers38 published the alkylation of aldehydes, both of which were catalysed by triplet metal-to-ligand-charge-transfer (3MLCT) excited state of [Ru(bpy)3]2+ generated upon visible-light irradiation. Subsequently, Stephenson and co-workers achieved the reductive dehalogenation of activated alkyl halides catalysed by [Ru(bpy)3]2+,44 and reductive dehalogenation of alkyl, alkenyl and aryl iodides by using fac-Ir(ppy)3 as photo-redox catalyst.45 Compared to [Ru(bpy)3]2+ and fac-Ir(ppy)3, the application of luminescent platinum(II) photo-catalysts is still nascent. Recently, our group and Wu's group demonstrated that pincer Pt(II) complexes are capable of catalysing light induced C–C bond formation.46,47
The important features that allow luminescent transition metal complexes to act as useful photo-redox catalysts or photo-sensitizers for light induced reactions include: (1) the long lifetime of their electronic excited states, thus allowing bimolecular reaction to proceed in solution; (2) their electronic excited states as both strong reducing and oxidizing reagents with reduction potentials systematically varied by the auxiliary ligands.41 For the photo-catalysis to have practical interest, the design of highly stable photo-redox catalysts with long-lived electronic excited states in solution is desirable.
Our endeavour to develop transition metal photo-catalysis is to use visible light for activation of small molecules such as CO2, described in this work, and for C–X bond functionalization. A major consideration is to utilize visible light, which falls within the solar spectrum and avoids deleterious high-energy UV-initialized photochemical side reactions.48,49 In the literature, platinum group metal complexes and semi-conductors are usually used as photo-redox catalysts for the photochemical reduction of CO2.50 Earlier examples of transition metal photo-catalysts used for the photochemical reduction of CO2 include cobalt porphyrins,51 Re(bpy)(CO)3Cl52,53 and Ir(terpy)(ppy)Cl.54 More recently, systems comprising fac-Ir(ppy)3 in conjunction with [Ni(Prbimiq1)]2+,55 Fe(porphyrin),56 [Co(TPA)Cl]Cl,57 (TPA = tris(2-pyridylmethyl)amine) and [Co(N5)]2+ (N5 = 2,13-dimethyl-3,6,9,12,18-pentaazabicyclo-[12.3.1]octadeca-1(18),2,12,14,16-pentaene)58 were reported for photochemical reduction of CO2.
The stability of photo-catalysts is an important issue for the practical application of transition metal photochemistry. Numerous studies revealed that the photochemically active excited states of [Ru(bpy)3]2+,59–61 [Ir(ppy)2(bpy)]+,60,62 and fac-Ir(ppy)355,57,63 are not stable under light irradiation for a long period of time as a result of dissociation of coordinated ligand(s) presumably via low lying d–d excited state(s). To address the photo-stability issue, we considered the use of N-heterocyclic carbene (NHC) ligands which have been receiving burgeoning attention in coordination chemistry due to their strong σ-donor strength to develop robust metal photo-sensitizers and photo-catalysts.64–70 Also, the N-substituent of NHC ligands can be used to tune both the physical and chemical properties of the resultant photo-active transition metal complexes such as their solubility in various solvents including water. NHC ligands functionalized with carboxylate, sulfonate, amine/ammonium, and alcohol motifs have been reported for the development of water-soluble transition metal catalysts for Suzuki coupling, hydrosilylation, hydrogenation, olefin metathesis and CO2 reduction.65,71
Compared with Ir(III) complexes supported by bidentate acetylacetonate28 and/or 2,2′-bypyridine1 ligands, the photophysical and application studies of the related bis-NHC Ir(III) complexes are relatively scarce. In 2010, cationic bis-NHC Ir(III) complexes were reported by De Cola and co-workers72 to have application in blue-light emitting electrochemical cells; subsequently, a number of bis-NHC Ir(III) complexes were reported for photophysical and biological studies.65,67,72–74 In this work, a series of strongly luminescent Ir(III) complexes (Chart 1) containing bis-NHC ligands and visible light absorbing C-deprotonated (C^N) ligands was synthesized and their photophysical and electrochemical properties were examined. These complexes display high photo-stability and are strongly emissive with long lifetimes of up to 96 μs in solution at room temperature. The water-soluble luminescent Ir(III) complexes, containing the glucose-functionalized NHC ligand, were found to be active photo-catalysts for radical cyclization leading to the formation of 5-membered pyrrole rings in aqueous media with high substrate conversions and yields. One of the photo-stable Ir(III) complexes was utilized as a photo-sensitizer and in conjunction with a recently reported catalyst [Co(TPA)Cl]Cl to convert CO2 into CO with a turnover number (TON) > 2400, selectivity in gas phase > 95% and yield of 5.6% (1 mL out of 18 mL of CO2 was converted into CO at 5 μM concentration of Co complex). Some of the complexes were also demonstrated as potential bioimaging and/or anti-cancer agents.
Chart 1 Iridium(III) complexes in this work. Complexes 1a,73,741b,731d28 and 2d31 have been reported in the literature. |
Complexes 1a–4Mea, 1b–4Meb, 4Hexb–9b with N-methyl or N-butyl substituent on bis-NHC ligands are soluble in most common aprotic solvents, but not in protic solvents e.g. methanol (MeOH) or water. Complexes 1c, 4Mec, 6Hc, 7c and 9c with glucose functionalized bis-NHC ligand are soluble in MeOH, ethanol (EtOH) and water.
The photo-stability of these Ir(III) complexes with bis-NHC ligands was examined by using 4Meb and 6Hb as representative examples. 4Meb and 6Hb in degassed deuterated MeCN were irradiated using blue light (12 W blue LEDs) for 5 days. The photolysis was monitored by 1H NMR spectroscopy. As depicted in Fig. 1a, in the case of 4Meb, less than 5% of the complex was observed to undergo photochemical decomposition after irradiation of 120 h, revealing its outstanding photo-stability. Under the same conditions, Ru(bpy)3Cl2, fac-Ir(ppy)3 and [(dFCF3ppy)2Ir(dtbbpy)]PF6 (dFCF3ppy = 3,5-difluoro-2-[5-(trifluoromethyl)-2-pyridinyl]phenyl; dtbbpy = 4,4′-bis(1,1-dimethylethyl)-2,2′-bipyridine) were found to decompose after 10 h irradiation as revealed by the changes of their 1H NMR spectra (Fig. 1c).
Fig. 1 1H NMR spectra for (a) 4Meb; (b) 6Hb; (c) Ru(bpy)3Cl2, fac-Ir(ppy)3 and [(dFCF3ppy)2Ir(dtbbpy)]PF6 in deuterated MeCN solution for irradiating by blue light (12 W, λmax = 462 nm)46 (all solutions are degased by nitrogen gas for 10 min); (d) crystal structure diagrams showing the photoinduced transformation of coordination for Ir metals in 6Hb and cis-6Hb (butyl group on bis-NHC ligand and hydrogen atoms are omitted for clarity); (e) UV/Vis absorption (dotted line), excitation (dashed line) and emission (solid line) spectra of solutions of 6Hb and cis-6Hb (concentration of 2.0 × 10−5 M) in degassed DCM at 298 K. |
Interestingly, irradiation of 6Hb using blue LEDs for 10 h led to a clean and quantitative conversion to a new species which did not show further changes upon subsequent irradiation for another 90 h (Fig. 1b, this species is noted as cis-6Hb). The yield of scale synthesis of cis-6Hb from the irradiation of degassed MeCN solution of 6Hb (100 mg in 4 mL MeCN) was 94%. cis-6Hb was found to be stable upon standing in solution in the dark for another 40 h, or exposed to air for another 20 h (Fig. S2, ESI†). To verify that the transformation of 6Hb to cis-6Hb was caused by visible-light irradiation, a negative control experiment was conducted by keeping 6Hb in deuterated MeCN in the dark (Fig. S2, ESI†), and no structural changes were detected by 1H NMR spectroscopy.
The UV-Vis absorption, emission and excitation spectra of cis-6Hb measured in DCM solution at 298 K are different from those of 6Hb with hypsochromic shifts in peak maxima (Fig. 1e). The emission band for cis-6Hb displays vibronic spacings of 1245 cm−1 while that of 6Hb shows spacings of 908 cm−1. Taking into consideration the ESI-MS and spectroscopic data of cis-6Hb, cis-6Hb is likely a structural isomer for 6Hb. The exact structure of cis-6Hb has been determined by X-ray crystallography.
Complex | Mediuma | Absorption (λmax/nm) (10−3ε/M−1 cm−1) | Emission λem/nm (τ/μs) | Φ /% |
---|---|---|---|---|
a Measured in degassed CH2Cl2 and water (2.0 × 10−5 M) at 298 K. b 4Mea, 4Meb and 4Hexb were measured at the concentration of 5 × 10−6 M. c 5b was measured at the concentration of 1 × 10−6 M. d Phosphorescence quantum yields were measured by using [Ru(bpy)3](PF6)2 (Φ = 0.062 in MeCN) as standard. | ||||
1a | CH2Cl2 | 255 (26.8), 267 (24.9), 311 (10.6), 342 (6.1), 380 (3.7), 416 (1.2) | 470 (2.1), 499, 534 | 89 |
1b | CH2Cl2 | 254 (36.8), 266 (35.5), 311 (15.5), 342 (9.1), 381 (5.82), 416 (2.16) | 470 (2.1), 500, 534 | 89 |
1c | H2O | 254 (36.8), 266 (35.5), 311 (15.5), 342 (9.1), 381 (5.82), 416 (2.16) | 469 (2.0), 498, 534 | 89 |
2a | CH2Cl2 | 252 (13.5), 290 (21.7), 332 (12.6), 369 (6.07), 405 (4.92) | 530, 547, 570 (3.03) | 11.1 |
3a | CH2Cl2 | 252 (15.5), 282 (15.1), 313 (20.5), 335 (23.1), 375 (21.9), 389 (20.9), 440 (20.9) | 472, 674 (4.8), 824 | 0.2 |
4Mea | CH2Cl2b | 260 (31.4), 297 (30.0), 316 (33.5), 334 (32.3), 360 (22.4), 378 (16.8), 421 (10.5) | 524, 564 (28.6), 614 | 65 |
4Meb | CH2Cl2b | 267 (52.2), 297 (46.8), 316 (50.9), 335 (49.0), 358 (36.3), 376 (27.3), 421 (16.2) | 525, 566 (28.7), 614 | 75 |
4Hexb | CH2Cl2b | 268 (42.1), 299 (42.3), 320 (50.4), 338 (49.2), 364 (32.6), 379 (25.6), 421 (14.6) | 527, 568 (32.7), 616 | 82 |
5b | CH2Cl2c | 260 (54.1), 273 (42.5), 310 (32.9), 323 (34.8), 377 (67.2), 398 (97.0), 436 (71.6) | 576 (96.1), 625, 683 | 22.6 |
6Hb | CH2Cl2 | 271 (23.0), 321 (25.0), 377 (9.57), 407 (6.81), 436 (3.38) | 531, 560 (6.4) | 78 |
6Hc | H2O | 268 (21.4), 321 (22.1), 377 (8.06), 407 (5.49), 436 (2.42) | 536, 560 (5.0) | 66 |
6Fb | CH2Cl2 | 270 (23.0), 323 (32.1), 383 (10.7), 415 (7.88), 447 (4.08) | 530, 558 (3.2) | 68 |
7b | CH2Cl2 | 296 (26.8), 333 (18.5), 388 (9.31), 437 (7.76), 467 (4.60) | 582, 623 (6.2) | 36 |
7c | H2O | 293 (25.9), 330 (17.4), 386 (9.63), 432 (6.92), 464 (3.71) | 583, 622 (5.9) | 32 |
8b | CH2Cl2 | 260 (80.1), 290 (48.5), 315 (35.4), 376 (34.5), 429 (7.52), 485 (3.47), 494 (2.88) | 618, 659 (5.2) | 3 |
9b | CH2Cl2 | 313 (24.2), 334 (23.8), 384 (12.7), 428 (10.7), 457 (8.20), 485 (3.47) | 617, 667 (7.4) | 13 |
9c | H2O | 315 (18.4), 380 (9.69), 420 (7.6), 454 (4.95), 475 (2.15) | 620 (4.5), 668 | 9 |
Fig. 2 UV-Vis absorption (top) and emission (bottom) spectra of solutions of 1a–3a and 1d–2d in degassed DCM (concentration of 2.0 × 10−5 M) at 298 K. |
On the other hand, the N-alkyl substituents on the bis-NHC ligand are found to have only minor effects, if any, on the UV/Vis absorption spectra of the Ir(III) complexes, as revealed by the overlaid spectra of the group of 1a, 1b, 1c (Fig. S4a, ESI†). Complex 1a absorbs weakly at the wavelength from 430 nm to 500 nm with molar absorptivity less than 500 M−1 cm−1 in contrast to the high values of 2500 M−1 cm−1 for 1d with the same C^N luminophore. The calculated wavelength for ground state HOMO → LUMO (S0 → S1 transition) is 369 nm and 414 nm for 1a and 1d, respectively (Fig. S4e and f, ESI†). These calculated values are in reasonable agreement with the corresponding experimental absorption λmax values 410 and 460 nm respectively.
Complexes 1–9 display strong phosphorescence in deaerated solution at room temperature (Fig. 2, 3 and S4 (ESI†); Table 1). All the complexes show vibronic-structured emissions, and their emission lifetimes are found to be in the microsecond regime. For example, 4Meb exhibits structured emission bands with vibrational spacings of ∼1380 cm−1 and a long emission lifetime of 28.2 μs. Only small negative solvatochromic effects on the emission at 524 nm are found (±5 nm; Fig. S4d, ESI†). These findings, together with TD-DFT calculations, suggest that the photoluminescence of the complexes is derived from triplet metal-perturbed ligand-centred (3LC) π–π* excited states. On the other hand, the structureless emission of 1d (the acetylacetonate (acac) analogous of 1a) at 516 nm should be ascribed to the 3MLCT/LLCT emission.2 Interestingly, complex 3a (Φ = 0.2%; τ = 4.8 μs) displays dual emission (Fig. 2) in contrast to the single emission of the acac analogue 3d,75 suggesting the possibility of modulation of photophysical properties of Ir(III) complexes by the NHC ligands.
Changes in chemical structure of the cyclometalated ligands also result in a profound effect on the photophysical properties of the Ir(III) complexes. For example, 5b displays a significant red-shift in emission maximum (576 nm) when compared to 4Meb and 4Hexb (525 and 527 nm respectively). This can be rationalized by the extended π-conjugation of the C^N ligands, leading to the decrease in energy of metal-perturbed 3LC emission. Moreover, for 4Mea, 4Meb and 4Hexb, their photoluminescence quantum yields in solution are affected by the substituents on the fluorenyl moiety as well as the N-alkyl groups of the bis-NHC ligands, e.g. complex 4Hexb exhibits a higher photoluminescence quantum yield than 4Mea and 4Meb. This is probably attributed to the fact that the long hexyl and N-butyl chains disfavour intermolecular stacking interactions among the planar C^N ligands, leading to reduced triplet–triplet annihilation and a higher photoluminescence quantum yield.
In addition, 5b shows a significantly longer emission lifetime (96.1 μs) than 4Meb (28.7 μs). This might be due to the reduced metal character in the electronic excited state of 5b, and hence slower triplet radiative decay. Similarly, with a smaller parentage of metal character in the frontier molecular orbitals, 6Hb (6.4 μs) shows a longer emission lifetime than 6Fb (3.2 μs). These long-lived triplet excited states allow the Ir(III) NHC complexes to undergo a variety of photochemical reactions, notably for visible-light-driven photo-catalytic reductive C–Br bond cleavage and CO2 reduction which will be illustrated later.
Complex | E pa /V | E pc /V |
---|---|---|
a Supporting electrolyte: 0.1 M nBu4NPF6 in MeCN and values are recorded vs. Ag/AgNO3 (0.1 M) in MeCN; Cp2Fe+/0 occurs at the range of 0.05–0.08 (V) vs. Ag/AgNO3. b Values refer to oxidation peak potential (Epa) at 25 °C for irreversible couples at a scan rate of 100 mV s−1. c Values refer to reduction peak potential (Epc) for the irreversible reduction waves. | ||
1b | 0.87 | −2.52 |
2a | 0.64 | −2.24 |
4Mea | 0.74 | −2.41 |
4Meb | 0.79 | −2.40 |
4Hexb | 0.74 | −2.44 |
5b | 0.72 | −2.25 |
6Hb | 0.98 | −2.14 |
6Fb | 1.04 | −1.94 |
7b | 0.69 | −2.09 |
8b | 0.62 | −1.94 |
9b | 0.62 | −2.16 |
Compared with 1b, [(dfppy)2Ir(bis-NHCBu)]PF6 (ref. 72) displays a more anodic oxidation potential of Eox = 1.04 V and a similar reduction potential of Ere = −2.37 V (vs. Cp2Fe+/0), which is attributed to stabilization of HOMOs as a result of the presence of highly electron-withdrawing F substitution on the C^N ligand (dfppy). For the irreversible reduction wave, a less negative reduction potential is found for 6Hb (Epc = −2.14 V vs. Ag/AgNO3, Table 2, Fig. S4†), as compared to that of 1b (Epc = −2.52 V vs. Ag/AgNO3, Table 2, Fig. 4) which has a less extended π-conjugation of the C^N ligand. The reductive wave is further anodically shifted in the case of 6Fb (Epc = −1.94 V vs. Ag/AgNO3, Table 2, Fig. S5a, ESI†). Therefore, the reduction process should be localized on the C^N ligands.
Fig. 4 Cyclic voltammograms of 1b, 2a, 4Meb, 5b and 6Hb in MeCN with nBu4NPF6 (0.1 M) as supporting electrolyte. Conditions: glass-carbon, working electrode, scan rate: 100 mV s−1. |
The time-resolved absorption and emission spectra of 4Meb recorded at various time intervals after excitation at 355 nm are depicted in Fig. 5. The kinetic decay analysis of bleaching of ground-state of 4Meb (τ350 nm = 15.4 μs, Fig. 5a, left inset) matches well with the growth of the absorption of triplet excited state (τ480 nm = 15.2 μs, Fig. 5a, right inset) as well as the emission lifetime (τ524 nm = 16.5 μs, Fig. 5b, right inset) in MeCN at 298 K.
The transient absorption spectra of 4Mec in aqueous solution recorded at different energies of laser beams (355 nm) reveal different spectral changes. As depicted in Fig. 6a, in addition to the growth of the absorption of triplet excited state of 4Mec from 380 to 700 nm, the emergence of an absorption band from 650 to 730 nm was observed at laser pulse energy ≥ 50 mJ (beam area: 0.5 cm2). This absorption band could be quenched upon addition of acetone (Fig. 6b and S7a and b, ESI†), but no quenching of transient absorption at 495 nm and emission at 524 nm were observed in the presence of acetone (Fig. 6c and S7, ESI†). The decay rate constant monitored at 720 nm of 4Mec (Fig. 6d, absence of acetone) was much faster than that measured at 495 nm in Fig. 6c/S7c† and decay of emission at 524 nm (Fig. S7d, ESI†). In view of these different kinetic behaviours, the transient absorption from 650 to 730 nm depicted in Fig. 6b might originate from hydrated electrons eaq−,76,77 which were formed by the photo-induced ionization of 4Mec in aqueous solution upon excitation with high energy laser beams. This is in line with a reported photoionization of [Pt2(POP)4]4−,78 as well as the findings in the studies of solvated electron with acetone.79,80 Similarly, complex 6Hc in aqueous solution was also observed to undergo photo-ionization as revealed by the increase in transient absorption in the region of 600 to 700 nm (Fig. S8b, ESI†). For 4Meb, its transient absorption spectra monitored at high laser pulse energy in MeCN exhibit similar profiles as for lower energy (Fig. S8b and c, ESI†). However, newly generated long-lived species have been observed at high laser pulse energies after 100 μs (Fig. S8b–e, ESI†), revealing that photo-ionization of 4Meb with the generation of [4Meb]+ likely occurs. The accompanying solvated electrons are not observed in this case due to ready quenching by MeCN.81
Based on the electrochemical data and the determination of E0–0 from the spectroscopic measurements, the excited state redox properties of the bis-NHC Ir(III) complexes can be estimated (Table S3, ESI†). The triplet excited states of the complexes are found to be powerful oxidants and reductants, and some of them are even more reactive towards photoredox reactions than [Ru(bpy)3]2+ and fac-Ir(ppy)3. A representative example is 4Meb (E(IrIV/III*) = −1.51 V vs. SCE) (here IrIV is a simple notation to denote the oxidized IrIII species, the site of the oxidation can be the metal and/or C^N ligand), which is a stronger reductant than [Ru(bpy)3]2+(E(RuIII/II*) of −0.81 vs. SCE).37,82 As a result, it is anticipated that the bis-NHC Ir(III) complexes described herein, upon photoexcitation in the visible-light region, can catalyse a number of reactions which are not feasible by the widely used [Ru(bpy)3]2+.
Chart 2 General photo-catalytic reactions by cyclometalated complexes. EWG = electron-withdrawing group. |
We considered a recent work by Lee and co-workers83 on visible-light-induced reductive cyclization of aromatic iodides and bromides to form indoline using [(ppy)2Ir(dtbbpy)]PF6 as photo-catalyst (PC). Yet, aryl bromides are found to be less reactive than iodides.84 Barriault and co-workers addressed this issue by using dinuclear gold(I) complexes ([Au2(dppm)2]OTf2) as photo-catalysts.85 However, this gold complex only shows absorption in the high-energy UV region (λ < 300 nm), which may result in destructive effects on the products and/or lead to undesired side reactions.
As the present bis-NHC Ir(III) complexes show strong absorption at λ > 400 nm, they were used to photo-catalyse the reductive cyclization of aryl bromides using blue LEDs. Among the bis-NHC Ir(III) complexes examined, 1b, 4Meb and 6Hb displayed good photo-catalytic activity for reductive cyclization of aryl bromide (A1) in terms of both substrate conversion and product yield (Table 3).
Entryb | PC | Conversionc/% | Yieldc/% |
---|---|---|---|
a Complex 3a was not tested because of the low quantum yield (0.2%, see Table 1). b Procedure: substrate 50 μmol, PC (1 mol%), DIPEA (5 equiv.), HCOOH (2.5 equiv.) in 4 mL MeCN solution was degassed by nitrogen, and irradiated by blue light (12 W, λmax = 462 nm) at ambient temperature for 4 h. c Determined by 1H NMR spectroscopy by adding internal standard of 5,5′-dimethyl-2,2′-bipyridine. | |||
1 | 1b | 93 | 59 |
2 | 2b | 1 | 0 |
3 | 4Meb | 90 | 59 |
4 | 5b | 90 | 52 |
5 | 6Hb | 93 | 59 |
6 | 6Fb | 80 | 44 |
7 | 7b | 49 | 24 |
8 | 8b | 11 | 7.3 |
9 | 9b | 0.1 | 0 |
In the case of another aryl bromide substrate A2, 4Meb showed both good substrate conversion and product yield (Table 4, entries 4–6). It was chosen as photo-sensitizer for optimization of the reaction conditions. A number of control experiments were performed, and no reaction was observed in the absence of amine or light (entries 13 and 14). Lowering the loading of PC (Table 3, entry 4), the absence of HCOOH or exposure to air (entries 11 and 12) were observed to decrease the substrate conversion of this reaction. Interestingly, using 1,8-diazabicyclo[5.4.0]undec-7-ene (DBU, entry 10, Eonset+/0 = 0.60 V vs. Cp2Fe+/0), also led to excellent substrate conversion and good product yield comparable to what was obtained with tetramethylethylenediamine (TMEDA, entry 9, Eonset+/0 = 0.11 V vs. Cp2Fe+/0, Fig. S9, ESI†). In contrast, the widely used photo-catalysts [Ru(bpy)3]Cl2 and fac-Ir(ppy)3 (Table 4, entries 15 and 16) showed little or no conversion under similar conditions.
Entrya | PC (substrate) | Amines | Conversionb/% | Yieldb/% |
---|---|---|---|---|
a Procedure: substrate 50 μmol, PC (2 mol%), amine (5 equiv.), HCOOH (2.5 equiv.) in 4 mL aqueous solution was degassed by nitrogen, and irradiated by blue light (12 W, λmax = 462 nm) at ambient temperature for 10 h. b Determined by 1H NMR spectroscopy by adding internal standard of 5,5′-dimethyl-2,2′-bipyridine. c Irradiated for 4 h. d Absence of HCOOH. e Presence of air (no degassing). f Absence of amine. g Absence of light. | ||||
1 | 1b (A1) | DIPEA | 97 | 67 |
2 | 4Meb (A1) | DIPEA | 97 | 64 |
3 | 6Hb (A1) | DIPEA | 96 | 64 |
4 | 4Meb (A2) | DIPEA | 98 | 54 |
5 | 1b (A2) | DIPEA | 99 | <15 |
6 | 6Hb (A2) | DIPEA | 84 | 51 |
7c | 4Meb (A3) | TEA | 97 | 79 |
8 | 4Meb (A1) | TEA | 91 | 64 |
9 | 4Meb (A1) | TMEDA | 52 | 32 |
10 | 4Meb (A1) | DBU | 96 | 72 |
11d | 4Meb (A1) | DIPEA | 85 | 54 |
12e | 4Meb (A1) | DIPEA | 62 | 28 |
13f | 4Meb (A1) | — | 0 | 0 |
14g | 4Meb (A1) | DIPEA | 0 | 0 |
15 | Ru(bpy)3Cl2 (A1) | DIPEA | 0 | 0 |
16 | fac-Ir(ppy)3 (A1) | DIPEA | 27 | 16 |
The radical cyclization of the alkyl bromides, B1 and B2, catalysed by 1b, 4Meb or 6Hb proceeded smoothly, with reasonable to excellent substrate conversion and product yields. The yield of cyclization of B1 was improved to up to 90% (entry 5, Table 3) when formic acid was added and 6Hb was used as a photo-catalyst (PC). When 2,2,6,6-tetramethyl-1-piperidinyloxy (TEMPO) was added, the reaction was totally inhibited, indicating the involvement of a radical intermediate in the reaction (entry 12).
The photo-catalytic reaction could be initiated from the oxidative quenching of 4Meb* with aryl/alkyl halides. This is because the excited state reduction potential of 4Meb* (E(IrIV/III*)pc = −1.51 V (vs. SCE), Table S3, Fig. S9†) can allow a direct one electron reduction of aryl/alkyl halides by 4Meb*, leading to carbon–halogen (σ*(C–X)) bond cleavage to give alkyl radical in the case of sp3 carbon or radical anion intermediate for sp2 carbon.86 Subsequent reactions of the alkyl radical or radical anion intermediate with C(sp2)–H bond lead to C–C bond formation. An aminium radical cation generated from the oxidation of amine by [4Meb]+83,85 could serve as an electron donor to complete the reductive process. In the case of [(ppy)2Ir(dtbbpy)]PF6,83 its triplet excited state ([(ppy)2Ir(dtbbpy)]+*) reacts with DIPEA via a reductive quenching mechanism to generate Ir(II) species (E(IrIII/II)pc = −1.51 V (vs. SCE)1, Fig. S9, ESI†), which initiates the subsequent reducing catalytic reaction (Table 5).
Entrya | PC | Amines | Conversionb/% | Yieldb/% |
---|---|---|---|---|
a Entry 1–12: R = H (substrate B1); procedure: substrate 50 μmol, PC (2 mol%), amine (5 equiv.) and HCOOH (2.5 equiv.) in 4 mL MeCN solution was degassed by nitrogen, and irradiated by blue light (12 W, λmax = 462 nm) at 25 °C. b Determined by 1H NMR spectroscopy by adding an internal standard of 5,5′-dimethyl-2,2′-bipyridine. c Absence of HCOOH (20 equiv.). d Absence of light. e Presence of TEMPO (radical trapping reagent, 2 equiv.). f Entry 13: R = Me (substrate B2). | ||||
1 | Ru(bpy)3Cl2 | DIPEA | 19 | 5 |
2 | fac-Ir(ppy)3 | DIPEA | 90 | 72 |
3 | 1b | DIPEA | 99 | 85 |
4 | 4Meb | DIPEA | 99 | 76 |
5 | 6Hb | DIPEA | 99 | 90 |
6 | 4Meb | TEA | 99 | 72 |
7 | 4Meb | TMEDA | 84 | 55 |
8 | 4Meb | DBU | 99 | 64 |
9 | 4Meb | — | 0 | 0 |
10c | 4Meb | DIPEA | 27 | 15 |
11d | 4Meb | DIPEA | 0 | 0 |
12e | 4Meb | DIPEA | 0 | 0 |
13f | 4Meb | DIPEA | 99 | 59 |
Interestingly, modifying the N-substituent of bis-NHC ligand from an alkyl group to a glucose moiety renders photo-catalyst 6Hc soluble in aqueous media. At the outset, we examined the 6Hc-catalysed reductive cyclization of B1 in a mixture of H2O/MeOH (3:1) with ascorbic acid as reductant, but both the substrate conversion and product yield were low. The use of diisopropylethylamine (DIPEA) as reductant instead of ascorbic acid and addition of tetrabutylammonium chloride improved the conversion and yield to 98% and 49%, respectively. Increasing the vol% of methanol in aqueous solution to 75% leads to 99% conversion and 87% product yield. To the best of our knowledge, this is the first example of visible-light-driven radical cyclization for synthesis of pyrrolidine in aqueous media (Table 6).
Entrya | Solvents (H2O/MeOH)b | Reductant | Conversionc/% | Yieldc/% |
---|---|---|---|---|
a Procedure: substrate 50 μmol, 6Hc (2 mol%), reductant (5 equiv.), nBu4NCl (5 equiv.) in 4 mL aqueous solution was degassed by nitrogen, and irradiated by blue light (12 W, λmax = 462 nm) at 25 °C. b Solvent system used is water/methanol in volume ratio (v/v). c Determined by 1H NMR spectroscopy by adding internal standard of 5,5′-dimethyl-2,2′-bipyridine. d Absence of HCOOH. e Absence of nBu4NCl. | ||||
1 | 3/1 | Ascorbic acid | 23 | 10 |
2d | 3/1 | Ascorbic acid | 20 | 14 |
3 | 3/1 | DIPEA | 98 | 49 |
4d,e | 3/1 | DIPEA | 90 | 21 |
5d | 3/1 | DIPEA | 79 | 31 |
6 | 1/1 | DIPEA | 99 | 64 |
7 | 1/3 | DIPEA | 99 | 87 |
8 | 0/1 | DIPEA | 99 | 66 |
A CO2-saturated MeCN/triethylamine solution (4:1, v/v; 4 mL) containing catalytic amounts of 4Meb and [Co(TPA)Cl]Cl was irradiated by blue LEDs (12 W) for a specified time period, and the evolved gases were separated and identified by GC-TCD equipped with a molecular sieve column. The volume of H2 and CO gases were calculated by using CH4 as the internal standard.
As shown in Fig. 7a and b, the amount of CO and H2 generated from the reaction mixture is found to show strong dependence on the concentrations of 4Meb and Co(II) catalysts. Particularly, the highest TON value of over 5000 can be accomplished at 0.5 μM of Co(II), while no generation of gases is found at low concentration of Co(II) (50 nM; Fig. S11, ESI†). Similarly, only negligible amount of product gases can be detected after irradiation for 24 h when [Ir] (0.005 mM) is lower than [Co] (0.05 mM). With the representative system containing 4Meb (0.5 mM) and Co(II) (0.005 mM), (Fig. 7c) the visible-light-driven CO2 reduction in three parallel reaction runs gives TON (CO) > 2400 (conversion of about 18 mL of CO2 into 1 mL of CO) with excellent selectivity in generating CO over H2 in the gaseous phase (>95%) after reaction for 72 h. This result is better than for the system utilizing fac-Ir(ppy)3 as PC, which reveals only TON (CO) > 900 and selectivity (CO) of 85% under similar reaction conditions.57
In order to confirm the roles of catalysts in photo-driven CO2 reduction, several control experiments were performed (Table S5, ESI†). Firstly, in the absence of Co(II) complex, no CO gas was observed in a CO2-saturated MeCN/TEA (4:1, v/v; 4 mL) solution after irradiation for 24 h. On the other hand, in the absence of light, sacrificial amine or PC, the reaction mixture only gives negligible amounts of CO. To ascertain the catalytic role of the Co(II) complex in the reaction, mercury was added to the reaction mixture in order to exclude the possibility of CO generation from heterogeneous Co nanoparticles. To specify, 29.2 μmol of CO (0.714 mL, TON 146) could be generated from the solution with [Ir] (0.5 mM), [Co] (0.05 mM) and elementary Hg (1 mL) after irradiation for 18 h, and this result is comparable with that using the solution mixture without Hg (28.4 μmol of CO formed, TON 142).
Since complexes 1b, 4Meb, 6Ha, 6Hb and 6Hc demonstrate outstanding photophysical properties i.e. high quantum yield with long-lived electronic excited states, cellular imaging of these complexes in HeLa cells were performed. After treatment of human cervical cancer cells (HeLa) with the Ir(III) complexes for 15 min, strong green/yellow luminescence was observed in the cytoplasm (but not nucleus) of cancer cells, as revealed by fluorescence microscopy (Fig. 8). Co-localization analyses indicate that the emission of these complexes are mainly localized in the endoplasmic reticulum (ER), which is stained by red-emissive ER-specific ER-Tracker™; a high Pearson's correlation coefficient (R) between the emissions of complexes and ER-Tracker™ is found (for example, 6Hb shows a high R value of 0.80). Consistently, these complexes do not accumulate in other organelles such as lysosome or mitochondria, as shown by the poor co-localization of the emissions of the complexes with the emission of Lyso-Tracker® and Mito-Tracker® respectively (Fig. S13, ESI†). Noticeably, 1b+ (both the counter anions of triflate and chloride) was found in our laboratory to be specifically localized in the ER of cancer cells (Fig. S14†), but not, as reported elsewhere, in mitochondria.73 With the specific accumulation of the complexes in ER, the cytotoxic properties of the complexes may originate from the induction of ER stress88 and immunogenic cell death.91
The different amounts of metal character in the frontier molecular orbitals of 1a, 1d and Ir(ppy)3, as deduced by TD-DFT calculations, can account for the photophysical properties and long emission lifetimes of the Ir(III) NHC complexes. As shown in Fig. S10 (ESI†), fac-Ir(ppy)3 and 1d have similar energy levels in HOMO and LUMO (around −4.5 eV and −2.0 eV). By simple substitution of the negatively charged ancillary ligands (C^N and acac) in Ir(ppy)3 and 1d by neutral NHC ligand, 1a is found to show lower energy levels of HOMO and LUMO (about −5.2 eV and −2.4 eV). This can be explained by the less electron-donating effect of the neutral NHC ligand compared to the negatively charged auxiliary ligands, as well as a certain contribution of the stabilization of dπ(Ir) by the π-acceptor orbitals of NHC ligand of 1a.68 As a result, the energy level of the HOMO of 1a is 700 mV lower than for 1d, which coincides with the experimental observation that the oxidation potential of 1a is 410 mV more positive than that of 1d. On the other hand, the transition energy of HOMO → LUMO in 1a is estimated to be 0.3 eV larger than in Ir(ppy)3 and 1d, which may account for the blue shift of low-energy absorption of 1a (369 nm vs. 410 nm), as compared to that of 1d and Ir(ppy)3, in UV-Vis absorption spectra. Notably, TD-DFT calculation reveals that frontier molecular orbitals of 1a show a smaller metal character than those of 1d (Ir character in the HOMO of 1a and 1d are found to be ∼30 and ∼40% respectively), and hence the triplet excited states of 1a are likely to have smaller metal character. This would probably slow down the spin–orbit coupling, resulting in slower radiative and non-radiative T1 → S0 decay for 1a. As a result, 1a shows a longer emission lifetime than 1d.
Considering the fact that the excited state of [Ru(bpy)3]2+ (E(RuII*/I) = 0.77 V vs. SCE) has a more powerful oxidative potential than for 4Meb (E(IrIII*/II) = 0.51 V vs. SCE, Table S3†) and no radical cyclization products are obtained when using [Ru(bpy)3]2+, the photo-catalytic route via reductive quenching cycle is not feasible. The photo-catalytic reaction would possibly be initialized via an oxidative quenching cycle of excited states of the photo-catalyst. By carefully examining the transient-absorption spectra of triplet state of 4Meb in MeCN, newly generated long-lived species were observed by using higher energy laser beams (Fig. S8b–g†). These long-lived species were found to be increased in the presence of substrate A1, and could be quenched by DIPEA. This long-lived species might be Ir(IV), which is generated by single electron transfer from [4Meb]* to A1. The calculated excited-state reduction potential of 4Meb (E(IrIV/III*) = −1.51 V vs. SCE, Table S3†) reveals that 4Meb is not a stronger photo-reductant than fac-Ir(ppy)3 (E(IrIV/III*) = −1.73 V vs. SCE).37,93 However, the performance of 4Meb in visible-light-driven radical cyclization and CO2 reduction prove to be better than for fac-Ir(ppy)3. For example, the conversion of substrates A1/B1 to indoline/pyrrolidine by 4Meb (97%/99%) are higher than those by fac-Ir(ppy)3 (27%/90%). Therefore, there should be other reasons for the good performance of 4Meb in photo-catalysis. Plausible reasons could be: (i) the generated Ir(IV) [4Meb]+ (E(IrIV/III) = 0.96 V vs. SCE, Table S3†) are more easily reduced by amines than [fac-Ir(ppy)3]+ (E(IrIV/III) = 0.77 V vs. SCE); and (ii) the higher photo-stability of 4Meb compared with that of fac-Ir(ppy)3.
Consequently, the excellent photo-stability, strong absorptivity, long lifetimes and the photo-ionization behaviours of our bis-NHC Ir(III) complexes possibly enable the photolysis of long-lived excited states of Ir(III)* to Ir(IV) more easily and thus promote the radical cyclization in a catalytic cycle.
Footnote |
† Electronic supplementary information (ESI) available: Additional experimental details, figures and tables. CCDC 1428476–1428479. For ESI and crystallographic data in CIF or other electronic format see DOI: 10.1039/c5sc04458h |
This journal is © The Royal Society of Chemistry 2016 |