A Diprotonated Porphyrin as an Electron Mediator in Photoinduced

11 hours ago - We have successfully constructed hydrogen-bonded supramolecular assemblies based on a diprotonated saddle-distorted porphyrin ...
0 downloads 0 Views 3MB Size
Subscriber access provided by UNIV OF LOUISIANA

C: Energy Conversion and Storage; Energy and Charge Transport

A Diprotonated Porphyrin as an Electron Mediator in Photoinduced Electron Transfer in Hydrogen-Bonded Supramolecular Assemblies Wataru Suzuki, Hiroaki Kotani, Tomoya Ishizuka, Masaki Kawano, Hayato Sakai, Taku Hasobe, Kei Ohkubo, Shunichi Fukuzumi, and Takahiko Kojima J. Phys. Chem. C, Just Accepted Manuscript • DOI: 10.1021/acs.jpcc.9b02449 • Publication Date (Web): 19 Apr 2019 Downloaded from http://pubs.acs.org on April 20, 2019

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 10 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

conformation due to the steric repulsion of peripheral phenyl groups. Therefore, H2DPP can be easily protonated by stoichiometric amount of weak acids such as carboxylic acids to form diprotonated one (H4DPP2+) derived from high basicity of inner imine N atoms due to the out-of-plane lone pairs, resulting the formation of hydrogen-bonded supramolecular assemblies without introducing hydrogen-bonding sites at the periphery.41-47 So far, hydrogen-bonded supramolecular assemblies based on H2DPP with carboxylic acids such as redox-active ferrocene carboxylic acid (FcCOOH) to form a photofunctional homo-triad (H4DPP2+(FcCOO–)2) have been reported.41 Due to the high reduction potential of diprotonated H2DPP (H4DPP2+), H4DPP2+ can act as an electron acceptor in photoinduced ET to form one-electron reduced H4DPP2+ (H4DPP•+) with the lifetime of pico-seconds. (Figure 1b).41,42,44 While porphyrins act as both electron donors and acceptors to form radical species in hydrogen-bonded supramolecular assemblies, no example of utilization of porphyrins as an electron mediator from an electron donor to an electron acceptor in hydrogen-bonded supramolecular assemblies: In other words, no report on utilization of these photogenerated radical species into subsequent ET reactions (Figure 1c). Given that porphyrins act as an electron mediator in supramolecular assemblies, generation of long-lived charge-separated or ET states as well as application to photocatalysis are expected through stepwise ET in supramolecular hetero-triads. Herein, we report the construction of hydrogen-bonded supramolecular assemblies between H4DPP2+ and an electron accepting molecular unit to utilize one electron reduced H4DPP2+ (H4DPP•+), which can be generated in the course of photoinduced ET reactions, as an electron donor. As an electron donor, a RuII-polypyridyl complex bearing a carboxyl group at the periphery (RuIICOOH) has been selected to form a hydrogen-bonded supramolecular assembly with H4DPP2+. In addition to the electron-donating ability of RuIICOOH, the complex would also be the precursor of an oxidation catalyst in supramolecular hetero-triads enabling photocatalytic oxidation reactions.48,49 On the other hand, benzyl viologen (BV2+) derivatives have been employed as an electron acceptor because of the high reduction potentials.50 Note that the one-electron reduced species (BV•+) is relatively stable,51 enabling to form a long-lived ET state in supramolecules. In addition, viologen derivatives have been often used as electron mediators in photocatalytic reduction reactions.52-55 Thus, we have employed a BV2+ derivative having a carboxyl group (BV2+COOH) as an electron acceptor in a hydrogen-bonded supramolecular assembly with H2DPP, H4DPP2+(BV2+COO–)2. We have examined the photodynamics of H4DPP2+(BV2+COO–)2 to confirm that H4DPP•+ acts as an electron mediator in the hydrogen-bonded supramolecule. RESULTS AND DISCUSSION Synthesis and Characterization of a Ru(II)-polypyridyl Complex and a Benzyl Viologen Derivative Bearing a Carboxylic Group. A RuII-polypyridyl complex, [RuIICl(TPA)(pyCOOH)](ClO4–) (RuIICOOH(ClO4–), TPA = tris(2-pyridylmethyl)amine, pyCOOH = 4-pyridinecarboxylic acid) was synthesized from the reaction of [RuII(µ-Cl)(TPA)]2(ClO4–)256

Scheme 1. Synthesis of RuIICOOH(ClO4–) (ClO4Ð)

2+ (ClO4Ð)2 N N N

RuII

N

RuII N

Cl N

N

N Cl

N

pyCOOH MeOH, reflux

N 2

N

RuII N

N Cl

COOH [RuII(µ-Cl)(TPA)]2(ClO4)2

RuIICOOH(ClO4)

Page 2 of 10

Figure 2. (a) An ORTEP drawing of RuIICOOH(ClO4–). Hydrogen atoms except for that attached to the carboxyl oxygen atom (O2) were omitted for clarity. (b) Intermolecular hydrogen bonding (blue dotted lines) between O2-H and Cl of RuIICOOH(ClO4–) to form a dimeric structure.

with pyCOOH in MeOH with 38% yield (Scheme 1).57 RuIICOOH(ClO4–) was characterized by 1H NMR and UV-Vis spectroscopy, ESI-TOF-MS, and elemental analysis. In the 1H NMR spectrum of RuIICOOH in acetone-d6 (Figure S1 in Supporting Information), two doublets derived from the pyCOOH ligand were observed at 7.69 and 8.47 ppm. The coordination geometry of RuIICOOH was confirmed by measurements of a nuclear Overhauser effect (NOE); the NOE was clearly observed between the 1H NMR signals assigned to the protons at the 6-positions of the equatorial pyridine moieties of TPA (9.17 ppm) and those at the 2-position of pyCOOH (8.47 ppm, Figure S2). This result suggests that pyCOOH coordinates to the RuII center at the trans position of the axial pyridine ring of the TPA ligand as shown in Scheme 1. ESI-TOF-MS of RuIICOOH in MeOH showed a peak cluster at m/z = 572.04, whose isotropic pattern was consistent with the computer-simulated one for [RuIICl(TPA)(pyCOONa)]+ (calcd: m/z = 572.04, Figure S3a). In the UV-Vis spectrum of RuIICOOH in acetone, an absorption band was observed at 436 nm, assignable to a ligand-to-metal charge-transfer (LMCT) band due to CT from the chloro ligand to the RuII center on the basis of literature (Figure S3b).58 X-ray crystallographic analysis of RuIICOOH(ClO4–) was conducted using a single crystal of RuIICOOH(ClO4–), which was obtained from a vapor diffusion of diethyl ether into an acetone solution of RuIICOOH(ClO4–). As shown in Figure 2a, the crystal structure of RuIICOOH(ClO4–) revealed that pyCOOH coordinated to the RuII center at the cis position of the tertiary amino nitrogen atom of the TPA ligand, that is, the trans position of the axial pyridine moiety. The coordination geometry in the crystal structure is consistent with the argument on the 1H NMR spectrum of the complex. The position of hydrogen atom attached to the carboxyl oxygen atom was determined in light of the packing structure of RuIICOOH(ClO4–): As shown in Figure 2b, the complex forms a hydrogen-bonding dimer in the crystal with the distance of 3.031(7) Å between O2 atom of one RuIICOOH(ClO4–) molecule and the chloro ligand of another RuIICOOH(ClO4–). To determine the pKa value of the appended carboxyl group of RuIICOOH(ClO4–), UV-vis spectroscopic titration of the complex was performed with an NaOH aqueous solution (NaOHaq) in Briton-Robinson buffer at 298 K (Figure S4). Upon addition of NaOHaq, the LMCT band of RuIICOOH(ClO4–) (414 nm at pH 1.52) slightly decreased together with a blue shift to 409 nm at pH 4.21 (Figure S4a). Then, the pKa value of the carboxyl group of RuIICOOH(ClO4–) was determined to be 2.18 ± 0.03 from the titration curve based on the absorbance change at 414 nm vs pH values (Figure S4b). The pKa value is low enough to diprotonate H2DPP to form H4DPP2+ in acetone quantitatively, judging from the result of quantitative diprotonation of H2DPP by meta-nitrobenzoic acid (pKa = 3.4 in H2O).47,59 In electrochemical measurements on RuIICOOH(ClO4–) in an acetone solution containing 0.1 M TBAPF6

ACS Paragon Plus Environment

Page 3 of 10 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Paragon Plus Environment

Page 4 of 10

Page 5 of 10 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

ACS Paragon Plus Environment

The Journal of Physical Chemistry

(2)

Ð!GBET = e(Eox Ð Ered)

(3)

Intrasupramolecular photoinduced ET occurs from the ground state of RuII centers to 1[H4DPP2+]* in H4DPP2+(RuIICOO–)2 to form the ET state, H4DPP•+(RuIIICOO–)(RuIICOO–), and BET proceeds to afford the ground state of H4DPP2+ rather than the triplet excited state (3[H4DPP2+]*); because the energy level of the ET state (1.13 eV) is lower than that of 3[H4DPP2+]* (1.46 eV). The lifetime ($) of the ET state (150 ps) is comparable with that of H4DPP2+(FcPhCOO–)2 ($ = 150 ps in PhCN, FcPhCOO– = 4-ferrocenebenzoate), which has almost the same donor-acceptor distance (10.1 Å from DFT calculations)41 with that of H4DPP2+(RuIICOO–)2 (9.501 Å from crystal structure of [H4DPP2+(Cl–)(RuIICOO–)](ClO4–) in Figure 5). Therefore, photoinduced ET properties of H4DPP2+(RuIICOO–)2 is concluded to be similar to that of hydrogen-bonded donor-acceptor supramolecular assemblies based on H4DPP2+ with ferrocene derivatives.41 Photophysical and ET properties of H4DPP2+(BV2+COO–)2. To examine the redox properties of H4DPP2+(BV2+COO–)2, CV and DPV measurements on H2DPP were performed in the presence of 2 equivalents of BV2+COOH in deaerated acetone in the presence of 0.1 M TBAPF6 as an electrolyte. As shown in Figure S14, a reversible redox wave derived from one-electron reduction of the BV2+COO– moiety in H4DPP2+(BV2+COO–)2 was observed at –0.75 V vs Fc/Fc+, which was the same with that of free BV2+COOH (Ered = –0.75 V vs Fc/Fc+ in Figure 3). On the other hand, an irreversible redox wave derived from one-electron reduction of H4DPP2+ was observed at –1.01 V vs Fc/Fc+ in contrast to the case of H4DPP2+(RuIICOO–)2, in which a quasi-reversible two-electron redox wave of H4DPP2+ was observed at –0.99 V vs Fc/Fc+ (Figure 7). This irreversible one-electron reduction of H4DPP2+ in H4DPP2+(BV2+COO–)2 suggests that one-electron reduced H4DPP2+ (H4DPP•+) would reduce free BV2+COO– through intermolecular ET in the CV time scale. Thus, judging from the higher reduction potential of BV2+COO– moieties than that of H4DPP2+, intrasupramolecular ET from H4DPP•+ to BV2+COO– should occur in H4DPP2+(BV2+COO–)2. Fluorescence spectrum of H4DPP2+(BV2+COO–)2 showed emission maxima at 779 nm (Figure S15a) in acetone, which was slightly red-shifted compared with H4DPP2+(CF3COO–)2 (Figure S12). From the absorption and emission maxima of H4DPP2+(BV2+COO–)2 (733 and 779 nm, respectively), 1[H4DPP2+]* in H4DPP2+(BV2+COO–)2 was determined to be 1.64 eV, comparable with that in H4DPP2+(RuIICOO–)2 (1.65 eV). Phosphorescence spectrum of H4DPP2+(BV2+COO–)2 in 2-MeTHF showed a shoulder around 850 nm (Figure S15b). Thus, the energy level of 3[H4DPP2+]* in H4DPP2+(BV2+COO–)2 was determined to be 1.46 eV, which was identical to that of H4DPP2+(RuIICOO–)2.

(a)

(b) 0.05

0.05 0.04

ÐÐ 0.8 ns ÐÐ 6.4 ns

560

!Abs at 560 nm

Ð!GET = Ðe(Eox Ð Ered) + 1E*

Next, we examined intrasupramolecular ET from H4DPP•+ that was formed by intermolecular ET from an external electron donor to BV2+COO– in H4DPP2+(BV2+COO–)2 in acetone at 298 K. In the absence of any external electron donors, the formation and the decay of 3 [H4DPP2+]* in H4DPP2+(BV2+COO–)2 were only observed by time-resolved ns-transient absorption spectra of H4DPP2+(BV2+COO–)2 in deaerated acetone, indicating that no photoinduced electron transfer from 3[H4DPP2+]* to BV2+COO– occurred and vice vasa (Figure S16). From the decay at 560 nm derived from 3[H4DPP2+]*, the decay rate constant (kT) of 3[H4DPP2+]* in H4DPP2+(BV2+COO–)2 was determined to be (1.73 ± 0.01) # 104 s–1 (Figure S16b), which was comparable with that of 3[H4DPP2+]* in H4DPP2+(CF3COO–)2 (kT = (1.60 ± 0.05) # 104 s–1) in MeCN (Figure S17). On the other hand, in the presence of decamethylferrocene (Me10Fc) as an external electron donor (Eox = –0.49 V vs Fc/Fc+ in acetone), the decay at 560 nm derived from 3[H4DPP2+]* in H4DPP2+(BV2+COO–)2 was accelerated with increasing the concentration of Me10Fc as observed in time-resolved ps-transient absorption spectra (Figure 9a,b). By changing the concentration of Me10Fc, the rate constant (ket) of ET from Me10Fc to 3[H4DPP2+]* in H4DPP2+(BV2+COO–)2 was determined to be (5.60 ± 0.01) # 109 M–1 s–1 (Figure 9c, blue circle), which was comparable to that ((6.86 ± 0.01) # 109 M–1 s–1) from Me10Fc to 3[H4DPP2+]* in H4DPP2+(CF3COO–)2 (Figure 9c red circle). Finally, the transient absorption spectrum showing the absorption maximum around 610 nm was observed at 6.4 ns after laser excitation at 532 nm (Figure 9a), clearly different from those in the absence of Me10Fc (Figure S16). This spectrum showed a good agreement with that of one-electron reduced BV2+COO–(BV•+COO–), which was generated by one-electron reduction of BV2+COOH with 1 equivalent of Na2S2O4 aqueous solution in acetone (Figure S18). Therefore, these spectral changes suggested intermolecular ET from Me10Fc to 3[H4DPP2+]* to form H4DPP•+, following by faster intrasupramolecular ET from H4DPP•+ to BV2+COO– to afford H4DPP2+ and BV•+COO–. In addition, the lifetime of BV•+COO– was determined to be 1.1 ms from the decay of 610 nm in transition absorption spectra of H4DPP2+(BV2+COO–)2 in the presence of Me10Fc (Figure S19). The energy diagram of photoinduced ET system of hydrogen-bonded

0.03 0.02 0.01

610

0 Ð0.01 500

600

0.03 0.02 0.01 0

700

¥ 0.2 mM ¥ 0.4 mM ¥ 0.6 mM ¥ 0.8 mM ¥ 1.0 mM

0.04

0

1

Wavelength, nm

2 3 Time, µs

4

5

(c) 2.0

kobs, 107 sÐ1

and the lifetime of electron-transfer state ($) was determined to be (1.60 ± 0.06) # 1011 s–1, (6.7 ± 0.9) # 109 s–1, and 150 ps, respectively, based on the time profile of the absorbance at 1040 nm (Figure 8b). The energy diagram of intrasupramolecular photoinduced ET and BET in hydrogen-bonded H4DPP2+(RuIICOO–)2 in acetone is depicted in Scheme 3. The driving forces of photoinduced ET (–!GET, eV) and BET (–!GBET, eV) were determined from the energy level of the ET state of H4DPP2+(RuIICOO–)2 (1.13 eV) and the energy level of 1[H4DPP2+]* using eq 2 and eq 3, where e is the elementary charge, Eox is the one-electron oxidation potential of an electron donor that is RuIICOO– in this case, Ered is the reduction potential of an electron acceptor that is H4DPP2+ in this case, and 1E * is the excitation energy of the singlet excited state of H4DPP2+.

!Abs

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 10

¥ CF3COOH ¥ BV2+COOH

1.5 1.0 0.5 0

0

0.5

1.0

1.5

2.0

[Me10Fc], mM

Figure 9. (a) Transient absorption spectra (%&x = 532 nm) of H4DPP2+(BV2+COO–)2 (0.20 mM) with Me10Fc (0.60 mM) in deaerated acetone at 0.8 ns (red), and 6.4 ns (blue) after picosecond laser excitation at 532 nm. (b) Decays time profiles of 3[H4DPP2+]* monitored at 560 nm with single exponential decay curve fitting: [H2DPP] = 2.0 # 10–4 M, [BV2+COOH] = 8.0 # 10–4 M. (c) Plots of kobs vs [Me10Fc] for H4DPP2+(BV2+COO–)2 (blue) and H4DPP2+(CF3COO–)2 (red).

ACS Paragon Plus Environment

Page 7 of 10

Scheme 4. Energy Diagram of H4DPP2+(BV2+COO–)2 in Acetone. 1[H DPP2+(BV2+COOÐ) ]* 4 2

+ Me10Fc

1.65 eV ISC

The Supporting Information is available free of charge via the Internet at http://pubs.acs.org. Experimental details and analytical data, including Figures S1-S20 and Table S1 (PDF), X-ray crystallographic data for RuIICOOH(ClO4–) and [H4DPP2+(Cl–)(RuIICOO–)](ClO4–) (CIF).

3[H DPP2+(BV2+COOÐ) ]* 4 2

+ Me10Fc

Energy

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

1.46 eV

AUTHOR INFORMATION

Intermolecular ET

H4DPP¥+(BV2+COOÐ)2 + Me10Fc+ h!

0.52 eV Intrasupramolecular ET (fast)

H4DPP2+(BV2+COOÐ)(BV¥+COOÐ) Intrasupramolecular BET " = 1.1 ms

+

[email protected]

Me10Fc+ 0.34 eV

H4DPP2+(BV2+COOÐ)2 + Me10Fc

H4DPP2+(BV2+COO–)2 with Me10Fc in acetone is depicted in Scheme 4. In Scheme 4, the energy levels of each state were calculated on the basis of the emission and UV-Vis absorption spectral data and redox potentials of the components by using eq 2 and eq 3. To confirm the requirement of association of H4DPP2+ with BV2+COO– for generation of the benzyl viologen radical cation (BV•+) through photoinduced electron transfer from H4DPP•+ to the benzyl violgen derivative, ns-transient spectra of H4DPP2+ with Me10Fc were measured in the presence of CF3COOH as a proton source and BV2+COOMe as a terminal electron acceptor, respectively. As shown in Figure S20, no absorption band around 610 nm due to BV•+COOMe was observed, while the decay of 3[H4DPP2+]* at 560 nm was accelerated upon addition of Me10Fc. This spectral change clearly indicates that no intermolecular ET from H4DPP•+ to BV2+COOMe occurred under the reaction conditions. Therefore, the hydrogen bonding between H4DPP2+ and BV2+COO– to form H4DPP2+(BV2+COO–)2 should be essential to generate H4DPP2+(BV2+COO–)(BV•+COO–) in the presence of external electron donors. CONCLUSION In conclusion, we have described successful construction of hydrogen-bonded supramolecular assemblies based on H4DPP2+ with redox-active molecules such as a RuII-polypyridyl complex (RuIICOOH) as an electron donor and a benzyl viologen derivative (BV2+COOH) as an electron acceptor. Formation of supramolecular homo-triads, H4DPP2+(RuIICOO–)2, H4DPP2+(BV2+COO–)2, were confirmed by spectroscopic measurements in acetone and the crystallographic analysis. Photoinduced ET reactions in the supramolecular assemblies have been also scrutinized by using laser flash photolysis. In H4DPP2+(RuIICOO–)2, intrasupramolecular ET from the RuII center of RuIICOO– to the singlet excited state of H4DPP2+ to afford the ET state involving one-electron reduced H4DPP2+ (H4DPP•+) with lifetime of 150 ps. On the other hand, in the presence of an external electron donor, intermolecular ET from the electron donor to the triplet excited state of H4DPP2+ in H4DPP2+(BV2+COO–)2 to form H4DPP•+(BV2+COO–)2. In addition, H4DPP•+ was able to reduce the hydrogen-bonded BV2+COO– through intrasupramolecular ET to afford one-electron reduced BV2+COO– (BV•+COO–), which exhibited relatively long lifetime. This is the first example of utilization of porphyrins as an electron mediator in hydrogen-bonded supramolecular assemblies. The role of an electron mediator of H4DPP2+ in supramolecules are expected to be applied to charge-separation systems to achieve the formation of long-lived electron-transfer states in supramolecular hetero-triads and to perform photocatalytic reductions such as photocatalytic hydrogen evolution. SUPPORTING IMFORMATION

Corresponding Author

ACKNOWLEDGMENTS This work was supported by Grant-in-Aid (17H03027 and 18J12184) from the Japan Society of Promotion of Science (JSPS, MEXT) of Japan. Financial supports through CREST (JST) are also appreciated (JPMJCR16P1).

REFERENCES (1)!Zouni, A.; Witt, H.-T.; Kern, J.; Fromme, P.; Krau', N.; Saenger, W.; Orth, P. Crystal Structure of Photosystem II from Synechococcus elongatus at 3.8 Å Resolution Nature 2001, 408, 739-743. (2)!Umena, Y.; Kawakami, K.; Shen, J.-R. Kamiya, N.Ì Crystal Structure of Oxygen-Evolving Photosystem II at a Resolution of 1.9!Å. Nature 2011, 473, 55-61 (3)!Dau, H.; Zaharieva, I.Ì Principles, Efficiency, and Blueprint Character of Solar-Energy Conversion in Photosynthetic Water Oxidation. Acc. Chem. Res. 2009, 42, 1861-1870. (4)!Fukuzumi, S.; Lee, Y.-M.; Nam, W. Mimicry and Functions of Photosynthetic Reaction Centers. Biochem. Soc. Trans. 2018, 46, 1279-1288. (5)!El-Khouly, M. E.; El-Mohsnawy, E.; Fukuzumi, S. Solar Energy Conversion: From Natural to Artificial Photosynthesis. J. Photochem. Photobiol., C: Photochem. Rev. 2017, 31, 36-83. (6)!Labat, F.; Bahers, T. L.; Ciofini, I.; Adamo, First-Principles Modeling of Dye-Sensitized Solar Cells: Challenges and Perspectives. C. Acc. Chem. Res. 2012, 45, 1268-1277. (7)!Kumara, N.; Lim, A.; Lim, C. M.; Petra, M. I.; Ekanayake, P. Recent Progress and Utilization of Natural Pigments in Dye Sensitized Solar Cells: A Review. Renew. Sust. Energ. Rev. 2017, 78, 301-317. (8)!Gust, D.; Moor, T. A.; Moore, A. L. Solar Fuels via Artificial Photosynthesis. Acc. Chem. Res. 2009, 42, 1890-1898. (9)!Fukuzumi, S. Development of Bioinspired Artificial Photosynthetic Systems. Phys. Chem. Chem. Phys. 2008, 10, 2283-2297. (10)! Fukuzumi, S.; Kojima, T. Photofunctional Nanomaterials Composed of Multiporphyrins and Carbon-Based (-Electron Acceptors. J. Mater. Chem. 2008, 18, 1427-1439. (11)! Fukuzumi, S.; Ohkubo, K.; Suenobu, T. Long-Lived Charge Separation and Applications in Artificial Photosynthesis. Acc. Chem. Res. 2014, 47, 1455-1464. (12)! Fukuzumi, S. Production of Liquid Solar Fuels and Their Use in Fuel Cells. Joule 2017, 1, 1-50. (13)! Windle, C. D.; George, M. W.; Perutz, R. N.; Summers, P. A.; Sun, X. Z.; Whitwood, A. C. Comparison of Rhenium– Porphyrin Dyads for CO2 Photoreduction: Photocatalytic Studies and Charge Separation Dynamics Studied by Time-Resolved IR Spectroscopy. Chem. Sci. 2015, 6, 6847-6864. (14)! Matlachowski, C.; Braum, B.; Tschierlei, S.; Schwalbe, M. Photochemical CO2 Reduction Catalyzed by Phenanthroline Extended Tetramesityl Porphyrin Complexes Linked with a

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(15)!

(16)!

(17)!

(18)!

(19)!

(20)!

(21)! (22)!

(23)!

(24)!

(25)!

(26)!

(27)!

(28)!

(29)!

(30)!

Rhenium(I) Tricarbonyl Unit. Inorg. Chem. 2015, 54, 10351-10360. Lazarides, T.; Sazanovich, I.; Simaan, A. J.; Kafentzi, M. C.; Delor, M.; Mekmouche, Y.; Faure, B.; Réglier, M.; Weinstein, J. A.; Coutsolelos, A. G.; Tron, T. Visible Light-Driven O2 Reduction by a Porphyrin-Laccase System. J. Am. Chem. Soc. 2013, 135, 3095-3103. Szulbinski, W.; Strojek, J. W. Photoinduced Reduction of Water by Tin(IV) and Ruthenium(II) Porphyrins. Inorg. Chim. Acta 1986, 118, 91-97. Wang, S.; Tabata, I.; Hisada, K.; Hori, T. Hydrogen Evolution Sensitized by Tin-Porphyrin in Microheterogeneous Systems. Dye Pigments 2002, 55, 27-33. Fateeva, A.; Chater, P. A.; Ireland, C. P.; Tahir, A. A.; Khimyak, Y. Z.; Wiper, P. V.; Darwent, J. R.; Rosseinsky, M. J. A Water‐ Stable Porphyrin‐Based Metal–Organic Framework Active for Visible‐Light Photocatalysis. Angew. Chem., Int. Ed. 2012, 51, 7440-7444. Amao, Y.; Watanabe, T. Photochemical and Enzymatic Methanol Synthesis from HCO3− by Dehydrogenases Using Water-Soluble Zinc Porphyrin in Aqueous Media. App. Catal. B 2009, 86, 109-113. Natali, M.; Deponti, E.; Vilona, D.; Sartorel, A.; Bonchio, M.; Scandola, F. A Bioinspired System for Light‐Driven Water Oxidation with a Porphyrin Sensitizer and a Tetrametallic Molecular Catalyst. Eur. J. Inorg. Chem. 2015, 3467-3477. Gust, D.; Moore, T. A. Mimicking Photosynthesis. Science 1989, 244, 35-41. Luo, C.; Guldi, D. M.; Imahori, H.; Tamaki, K.; Sakata, Y. Sequential Energy and Electron Transfer in an Artificial Reaction Center:! Formation of a Long-Lived Charge-Separated State. J. Am. Chem. Soc. 2000, 122, 6535-6551. H. Imahori, K. Tamaki, D. M. Guldi, C. Luo, M. Fujitsuka, O. Ito, Y. Sakata, S. Fukuzumi, Modulating Charge Separation and Charge Recombination Dynamics in Porphyrin−Fullerene Linked Dyads and Triads:! Marcus-Normal versus Inverted Region. J. Am. Chem. Soc. 2001, 123, 2607-2617. Imahori, H.; Guldi, D. M.; Tamaki, K.; Yoshida, Y.; Luo, C.; Sakata, Y.; Fukuzumi, S. Charge Separation in a Novel Artificial Photosynthetic Reaction Center Lives 380 ms. J. Am. Chem. Soc. 2001, 123, 6617-6628. Iengo, E.; Pantoş, G. D.; Sanders, J. K. M.; Orlandi, M.; Chiorboli, C.; Fracasso, S.; Scandola, F. A Fully Self-Assembled Non-Symmetric Triad for Photoinduced Charge Separation. Chem. Sci. 2011, 2, 676-685. Poddutoori, P. K.; Sandanayaka, A. S. D.; Hasobe, T. Ito, O.; van der Est, A. Photoinduced Charge Separation in a Ferrocene– Aluminum(III) Porphyrin–Fullerene Supramolecular Triad. J. Phys. Chem. B 2010, 114, 14348-14357. Poddutoori, P. K.; Sandanayaka, A. S. D.; Zarrabi, N.; Hasobe, T. Ito, O.; van der Est, A. Sequential Charge Separation in Two Axially Linked Phenothiazine–Aluminum(III) Porphyrin– Fullerene Triads. J. Phys. Chem. A 2011, 115, 709-717. Poddutoori, P. K.; Bregles, L. P.; Lim. G. N.; Boland, P.; Kerr, R. G.; D’Souza, F. Modulation of Energy Transfer into Sequential Electron Transfer upon Axial Coordination of Tetrathiafulvalene in an Aluminum(III) Porphyrin–Free-Base Porphyrin Dyad. Inorg. Chem. 2015, 54, 8482-8494. Amati, A.; Cavigli, P.; Kahnt, A.; Indelli, M. T.; Iengo, E. Self-Assembled Ruthenium(II) Porphyrin-Aluminum(III) Porphyrin–Fullerene Triad for Long-Lived Photoinduced Charge Separation. J. Phys. Chem. A 2017, 121, 4242-4252. Zarrabi, N.; Agatemor, C.; Lim, G. N.; Matula, A. J.; Bayard, B. j.; Batista, V. S.; D’Souza, F.; Poddutoori, P. K. High-Energy Charge-Separated States by Reductive Electron Transfer

(31)!

(32)!

(33)!

(34)!

(35)!

(36)!

(37)!

(38)!

(39)!

(40)!

(41)!

(42)!

(43)!

Page 8 of 10

Followed by Electron Shift in the Tetraphenylethylene– Aluminum(III) Porphyrin–Fullerene Triad. J. Phys. Chem. C 2019, 123, 131-143. Cavigli, P.; Da Ros, T.; Kahnt, A.; Gamberoni, M.; Indelli, M. T.; Iengo, E. Zinc Porphyrin–Re(I) Bipyridyl–Fullerene Triad: Synthesis, Characterization, and Kinetics of the Electron-Transfer Processes by Visible Excitation. Inorg. Chem. 2015, 54, 280-292. Obondi, C. O.; Lim. G. N.; Jang, Y.; Patel, P.; Wilson, A. K.; Poddutoori, P. K. D’Souza, F. Charge Stabilization in High-Potential Zinc Porphyrin–Fullerene via Axial Ligation of Tetrathiafulvalene. J. Phys. Chem. C 2018, 122, 13636-13647. Ward, M. D. Photo-Induced Electron and Energy Transfer in Non-Covalently Bonded Supramolecular Assemblies. Chem. Soc. Rev. 1997, 26, 365-375. Sessler, J. L.; Jayawickramarajah, J.; Gouloumis, A.; Torres, T.; Guldi, D. M.; Maldonado, S.; Stevenson, K. J. Synthesis and Photophysics of a Porphyrin–Fullerene Dyad Assembled Through Watson–Crick Hydrogen Bonding. Chem. Commun. 2005, 1892-1894. Sánchez, L.; Sierra, M.; Martin, N.; Myles, A. J.; Dale, T. J.; Rebek, J. Jr.; Seitz, W.; Guldi, D. M. Exceptionally Strong Electronic Communication through Hydrogen Bonds in Porphyrin–C60 Pairs. Angew. Chem. Int. Ed. 2006, 45, 4637-4641. D’Souza, F.; Venukadasula, G. M.; Yamanaka, K.; Subbaiyan, N. K.; Zandler, M. E.; Ito, O. Through-Bond Photoinduced Electron Transfer in a Porphyrin-Fullerene Conjugate Held by a Hamilton Type Hydrogen Bonding Motif. Org. Biomol. Chem. 2009, 7, 1076-1080. Yu, M.-L.; Wang, S.-M.; Fang, K.; Khoury, T.; Crossley, M. J.; Fan, Y.; Zhang, J.-P.; Tung, C-H.; Wu, L-Z. Photoinduced Electron Transfer and Charge-Recombination in 2-Ureido-4[1H]-Pyrimidinone Quadruple Hydrogen-Bonded Porphyrin–Fullerene Assemblies. J. Phys. Chem. C 2011, 115, 23634-23641. García-Iglesias, M.; Peuntinger, K.; Kahnt, A.; Krausmann, J.; Vázquez, P.; González-Rodríguez, D.; Guldi, D. M.; Torres, T. Supramolecular Assembly of Multicomponent Photoactive Systems via Cooperatively Coupled Equilibria. J. Am. Chem. Soc. 2013, 135, 19311-19318. de Rege, P. J. F.; Williams, S. A.; Therien, M. J. Direct Evaluation of Electronic Coupling Mediated by Hydrogen Bonds: Implications for Biological Electron Transfer. Science 1995, 269, 1409-1413. Kielmann, M.; Senge, M. O. Molecular engineering of free-base porphyrins as ligands–The N-H•••X binding motif in tetrapyrroles. Angew. Chem., Int. Ed. 2019, 58, 418-441. Honda, T.; Nakanishi, T.; Ohkubo, K.; Kojima, T.; Fukuzumi, S. Structure and Photoinduced Electron Transfer Dynamics of a Series of Hydrogen-Bonded Supramolecular Complexes Composed of Electron Donors and a Saddle-Distorted Diprotonated Porphyrin. J. Am. Chem. Soc. 2010, 132, 10155-10163. Kojima, T.; Honda, T.; Ohkubo, K.; Shiro, M.; Kusukawa, T.; Fukuda, T.; Kobayashi, N.; Fukuzumi, S. A Discrete Supramolecular Conglomerate Composed of Two Saddle‐ Distorted Zinc(II)‐Phthalocyanine Complexes and a Doubly Protonated Porphyrin with Saddle Distortion Undergoing Efficient Photoinduced Electron Transfer. Angew. Chem., Int. Ed. 2008, 47, 6712-6716. Honda, T.; Kojima, T.; Fukuzumi, S. Crystal Structures and Properties of a Monoprotonated Porphyrin. Chem. Commun. 2009, 4994-4996.

ACS Paragon Plus Environment

Ì

Ì

Page 9 of 10 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

(44)! Fukuzumi, S.; Honda, T.; Kojima, T. Structures and Photoinduced Electron Transfer of Protonated Complexes of Porphyrins and Metallophthalocyanines. Coord. Chem. Rev. 2012, 256, 2488-2502. (45)! Kojima, T.; Nakanishi, T.; Honda, T.; Fukuzumi, S. Photoinduced Electron Transfer in Supramolecular Assemblies Involving Saddle-Distorted Porphyrins and Phthalocyanines. J. Porphyrins Phthalocyanines 2009, 13, 14-21. (46)! Suzuki, W.; Kotani, H.; Ishizuka, T.; Ohkubo, K.; Shiota. Y.; Yoshizawa, K.; Fukuzumi, S.; Kojima, T. Thermodynamics and Photodynamics of a Monoprotonated Porphyrin Directly Stabilized by Hydrogen Bonding with Polar Protic Solvents. Chem.-Eur. J. 2017, 23, 4669-4679. (47)! Suzuki, W.; Kotani, H.; Ishizuka, T.; Shiota, Y. Yoshizawa, K.; Kojima, T. Formation of Supramolecular Hetero-Triads by Controlling the Hydrogen Bonding of Conjugate Bases with a Diprotonated Porphyrin Based on Electrostatic Interaction. Chem. Commun. 2017, 53, 6359-6362. (48)! Ishizuka, T.; Kotani, H.; Kojima, T. Characteristics and Reactivity of Ruthenium–Oxo Complexes. Dalton Trans. 2016, 45, 16727-16750. (49)! Ishizuka, T.; Ohzu, S.; Kojima, T. Oxidation of Organic Substrates with RuIV=O Complexes Formed via Proton-Coupled Electron Transfer. Synlett 2014, 25, 1667-1679. (50)! Wang, D.; Crowe, W. E.; Strongin, R. M.; Sibrian-Vazques, M. Exploring the pH Dependence of Viologen Reduction by α-Carbon Radicals Derived from Hcy and Cys. Chem. Commun. 2009, 1876-1878. (51)! Berville, M.; Richard, J.; Stolar, M.; Choua, S.; Breton, N. L.; Gourlaouen, C.; Boudon, C.; Ruhlmann, L.; Baumgartner, T.; Wytko, J. A.; Weiss, J. A Highly Stable Organic Radical Cation. Org. Lett. 2018, 20, 8004-8008. (52)! Tanaka, S.; Nakazono, T.; Yamauchi, K. Sakai, K. Photochemical H2 Evolution Catalyzed by Porphyrin-based Cubic Cages Singly and Doubly Encapsulating PtCl2(4,4′-dimethyl-2,2′-bipyridine). Chem. Lett. 2017, 46, 1573-1575. (53)! Jiang, D.-L.; Choi, C.-K.; Honda, K. Li, W.-S.; Yuzawa, T.; Aida, T. Photosensitized Hydrogen Evolution from Water Using Conjugated Polymers Wrapped in Dendrimeric Electrolytes. J. Am. Chem. Soc. 2004, 126, 12084-12089. (54)! Yang. L.; Frith, J. T.; Garcia-Araez, N.; Owen, J. R. A New Method to Prevent Degradation of Lithium–Oxygen Batteries: Reduction of Superoxide by Viologen. Chem. Commun. 2015, 51, 1705-1708. (55)! Amao, Y.; Yamane, K.; Okura, I. Effect of the Chain Length in Water Soluble Viologen Linked Zinc Porphyrins on Hydrogen Evolved Activity with Hydrogenase. Chem. Lett. 1997, 871-872. (56)! Kojima, T.; Amano, T. Ishii, Y.; Ohba, M.; Okaue, Y.; Matsuda, Y. Synthesis and Characterization of Mononuclear and Dinuclear Ruthenium Complexes with Tris(2-pyridylmethyl)amine and Tris(5-methyl-2-pyridylmethyl)amine. Inorg. Chem. 1998, 37, 4076-4085. (57)! Brewster, T. P.; Konezny, S. J.; Sheehan, S. W.; Martini, L. A.; Schmuttenmaer, C. A.; Batista, V. S.; Crabtree, R. H. Hydroxamate Anchors for Improved Photoconversion in Dye-Sensitized Solar Cells. Inorg. Chem. 2013, 52, 6752-6764. (58)! Kojima, T.; Hayashi, K.; Matsuda, Y. Structures and Properties of Ruthenium(II) Complexes of Pyridylamine Ligands with Oxygen-Bound Amide Moieties:! Regulation of Structures and Proton-Coupled Electron Transfer. Inorg. Chem. 2004, 43, 6793-6804. (59)! Fasman, G. D. Handbook of Biochemistry and Molecular Biology, 3rd Ed, CRC Press, 1976, p 305-351.

(60)! Park, Y. S.; Lee, E. J.; Chun, Y. S.; Yoon, Y. D.; Yoon, K. B. Long-Lived Charge-Separation by Retarding Reverse Flow of Charge-Balancing Cation and Zeolite-Encapsulated Ru(bpy)32+ as Photosensitized Electron Pump from Zeolite Framework to Externally Placed Viologen. J. Am. Chem. Soc. 2002, 124, 7123-7135. (61)! Chang, M.; Chen, M.-Z.; Zhou, C.-Q.; Lin, W.-E.; Chen, J.-X.; Chen, W.-H.; Jiang, Z.-H. Towards Polynuclear Metal Complexes with Enhanced Bioactivities: Synthesis, Crystal Structures and DNA Cleaving Activities of CuII, NiII, ZnII, CoII and MnII Complexes Derived from 4-Carboxy-1-(4-carboxybenzyl) Pyridinium Bromide. Inorg. Chim. Acta 2013, 405, 461-469. (62)! Webb, M. J.; Bampos, N. Noncovalent Interactions in Acid– Porphyrin Complexes Chem. Sci. 2012, 3, 2351-2366. (63)! Hirose, K. A Practical Guide for the Determination of Binding Constants. Journal of Inclusion Phenomena and Macrocyclic Chemistry, 39, 193-209. Inclusion Phenom. Macrocyclic Chem. 2001, 39, 193-209. (64)! Jentzen, W.; Turowska-Tyrk, I.; Scheidt, W. R.; Shelnutt, J. A. Planar Solid-State and Solution Structures of (Porphinato)nickel(II) As Determined by X-ray Diffraction and Resonance Raman Spectroscopy. Inorg. Chem. 1996, 35, 3559-3567. (65)! Nakanishi, T.; Ohkubo, K.; Kojima, T.; Fukuzumi, S. Reorganization Energies of Diprotonated and Saddle-Distorted Porphyrins in Photoinduced Electron-Transfer Reduction Controlled by Conformational Distortion. J. Am. Chem. Soc. 2009, 131, 577-584.

Ì

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Paragon Plus Environment

Page 10 of 10