You are on page 1of 9

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/259350366

Electron transfer reactions of ruthenium(II)-bipyridine complexes carrying


tyrosine moiety with quinones

Article in Luminescence · November 2014


DOI: 10.1002/bio.2617 · Source: PubMed

CITATIONS READS

8 1,311

4 authors, including:

P. Muthu Mareeswaran Eswaran Rajkumar


Anna University, Chennai Madras Christian College
69 PUBLICATIONS 567 CITATIONS 28 PUBLICATIONS 222 CITATIONS

SEE PROFILE SEE PROFILE

Sathish Veerasamy
Bannari Amman Institute of Technology
51 PUBLICATIONS 704 CITATIONS

SEE PROFILE

All content following this page was uploaded by Sathish Veerasamy on 30 October 2017.

The user has requested enhancement of the downloaded file.


Research article
Received: 31 July 2013, Revised: 7 October 2013, Accepted: 27 October 2013 Published online in Wiley Online Library

(wileyonlinelibrary.com) DOI 10.1002/bio.2617

Electron transfer reactions of ruthenium(II)–


bipyridine complexes carrying tyrosine moiety
with quinones
Paulpandian Muthu Mareeswaran,a Eswaran Rajkumar,a,b
Veerasamy Sathisha and Seenivasan Rajagopala*
ABSTRACT: Three ruthenium(II)–bipyridine complexes carrying a tyrosine moiety were synthesized and photophysical and
electron transfer studies with quinones were carried out using absorption and emission spectral techniques. The binding
efficiency of quinones with ruthenium(II)–bipyridine complexes was also studied using these techniques. The binding
efficiency was moderate and similar for all complexes with all quinones. The quenching modes were also similar and efficient
for all complexes with all quinones. The quenching processes were diffusion controlled. The rate of electron transfer was
calculated using semiclassical theory. Copyright © 2013 John Wiley & Sons, Ltd.

Additional supporting information may be found in the online version of this article at the publisher’s web-site.

Keywords: electron transfer; photophysics; ruthenium(II)–bipyridine–tyrosine; quinones; semiclassical theory

Introduction reduced in two ways: (i) back ET from the electron acceptor
(quinones) and (ii) intramolecular electron transfer from the
Electrons shuttle between water and CO2 during photosynthesis tyrosine moiety (27–30).
(1). Water is split and transfers its electron on the water oxida- Quinones are ubiquitous in nature. They are ET mediators
tion catalyst (WOC, manganese cluster) (1,2). The reaction center between PSI and PSII, and act as anticancer drugs in medicine
of photosystem II (PSII) receives the electron from WOC via a (31–38). Quinones appear to be predestined as electron
tyrosine moiety attached to the active center of PSII, which is acceptors in nature for a variety of reasons (39–41). In order to un-
also undergoing a hydrogen bonding interaction with the derstand the electron-accepting properties of quinones in natural
WOC (3–5). In the excited state, PSII transfers its electron to photosynthesis, several model photosensitizers (metal–porphyrin,
nearby electron acceptors, pheophytin and quinones (QA and [Ru(NN)3]2+ complexes) have been designed and inter- and intra-
QB) (6–8). These quinones deliver the electron to photosystem molecular ET reactions with quinones have been studied (42–44).
I (PSI) in the excited state, which is further used in reduction of The ET reactions of several quinones with excited state [Ru(NN)3]2+
CO2 to form carbohydrates (1,9). Thus, the excited states of PSI complexes have been reported previously (42–45). ET from
and PSII attached to tyrosine and nearby quinones are the media- phenolate to [Ru(NN)3]2+ has been reported extensively from this
tors of electron transfer (ET) between the WOC and PSI (6–11). laboratory (46–49). Herein, we report a detailed study on the excited
Therefore, it is essential that the tyrosine–reaction center–quinone state ET reactions of [Ru(NN)3]2+ complexes carrying a tyrosine
system is mimicked exactly in order to understand more precisely moiety with quinones using steady-state and time-resolved
the redox processes occuring in photosynthesis (9–14). measurements. The semiclassical theory of ET has also been applied
Ruthenium(II) polypyridyl complexes ([Ru(NN)3]2+) are widely successfully for the excited state ET reactions of [Ru(NN)3]2+
used to mimic the ET behavior of the PSII reaction center complexes carrying a tyrosine moiety with quinones.
(15–17). The advantage with [Ru(NN)3]2+ complexes is that the
properties of the excited state can be varied systematically by
introducing electron-donating and electron-withdrawing subs-
titutents in the 4,4′-position of the 2,2′-bipyridine (bpy) ligand
Experimental
(17–21). Various amino acid moieties have been introduced into The compounds bpy (98.9%), RuCl3·3H2O (99.98%) were from
ligand bpy of [Ru(bpy)3]2+ complexes via peptide linkage, and Merck and used without further purification. Benzoquinone
the photophysical properties have been studied widely (22–25).
However in PSII, tyrosineZ is the amino acid that mediates ET from
the manganese cluster (3,11). Therefore, [Ru(NN)3]2+ complexes * Correspondence to: Seenivasan Rajagopal, School of Chemistry,
carrying the tyrosine moiety may be an excellent mimic of the PSII Madurai Kamaraj University, Madurai, Tamil Nadu, India. E-mail:
reaction center carrying tyrosine (26). Excitation of a [Ru(NN)3]2+ rajagopalseenivasan@yahoo.com
complex by a photon leads to the formation of a triplet metal- a
School of Chemistry, Madurai Kamaraj University, Madurai, Tamil Nadu, India
to-ligand charge transfer (3MLCT) excited state and transfer of
its electron to the electron acceptor (quinones). Ru3+ is then b
Vel Tech University, Avadi, Chennai, Tamil Nadu, India

Luminescence 2013 Copyright © 2013 John Wiley & Sons, Ltd.


P. M. Mareeswaran et al.

(BQ; > 98.0%); naphthoquinone (NQ; 97.0%); 2,5-dichlo- smoke formed above the solution in the reaction flask, and the
robenzoquinone (DCBQ; 99.0%); tetrachloro-1,4-benzoquinone solution was heated to reflux under nitrogen for 2.5 h and then
(TCBQ; 99.0%); and 7,7,8,8-tetracyanoquinodimethane (TCNQ; cooled to room temperature. White crystals of triethylamine
98.0%) were from Sigma-Aldrich and their structures are shown hydrochloride formed and were filtered off. The filtrate was
in Scheme 1. The compounds 4,4′-dimethyl-2,2′-bipyridine concentrated to ~ 5 mL. The crude product was purified by
(dmbpy; 99%); 2,2′-bipyridine-4,4′-dicarboxylic acid (dcbpy; 98%) repetitive column chromatography on neutral aluminum oxide
and tyrosine ethyl ester (99.0%) were also from Sigma-Aldrich. with gradient eluents, CH2Cl2 and CH2Cl2/CH3OH (90:10 v/v). The
desired fractions were combined and the solvents evaporated
to dryness to give a red solid. The red solid was dissolved in water
Synthesis of ruthenium(II)–bipyridine and to this aqueous solution was added concentrated aqueous
complexes carrying a tyrosine moiety solution of NH4PF6 to produce a red precipitate. Filtration and
The ligands 4-methyl-4′-carboxy-2,2′-bipyridine (cmbpy), 4,4′- washing with water and petroleum ether gave a red solid with
dicarboxyethyl ester-2,2′-bipyridine (dcet) and the complexes [Ru 70% yield. ESI-MS (M-PF6+ Na) 1074.44 m/z.
(bpy)2]Cl2, [Ru(dmbpy)2]Cl2, [Ru(dcet)2]Cl2, [Ru(bpy)2(cmbpy)]·2PF6,
[Ru(dmbpy)2(cmbpy)]·2PF6 and [Ru(dcet)2(cmbpy)]·2PF6 were Preparation of [Ru(dmbpy)3–T)]·2PF6
synthesized using reported procedures (50–53). [Ru(NN)3]2+
complexes carrying a tyrosine moiety, Ru(bpy)3–T, Ru(dmbpy)3– The same synthetic procedure was adopted for the synthesis of
T, Ru(dcet)3–T, were synthesized using the following procedure [Ru(dmbpy)2(c-t)]·2PF6 complex. The product yield is 66%.
and characterized using ESI-MS spectral techniques. The spectral ESI-MS (M-2PF6) 875.03 m/z.
results were similar to existing reports (22,53,54).
Preparation of [Ru(dcet)3–T)]·2PF6
Preparation of [Ru(bpy)3–T)]·2PF6 [Ru(dcet)2(cmbpy)]·2PF6 (470 mg, 0.39 mmol) was refluxed in 15
[Ru(bpy)2(cmbpy)]·2PF6 (400 mg, 0.43 mmol) was dissolved in 30 mL of thionyl chloride for 3 h. Excess of thionyl chloride was
mL of thionyl chloride and the solution was heated to reflux removed under reduced pressure, and the residue dried under
under nitrogen for 2 h. Evaporation of the excess thionyl vacuum at 50°C for 1 h. The resulting solid (acid chloride) was
chloride in vacuum gave a dark red oil, which was immediately redissolved in 5 mL of dry CH2Cl2. The solution was added
used for the next step of the reaction. Tyrosine ethyl ester (160 dropwise to a solution of tyrosine ethyl ester (265 mg, 0.42
mg, 0.65 mmol) was suspended in acetonitrile (15 mL, 99.9%) mmol) and triethylamine (0.5 mL) in dry CH2Cl2 (10 mL). The
and the solid in the suspension was dissolved after addtion of resulting solution was stirred under nitrogen at room temperature
triethylamine (0.4 mL, 2.5 mmol). The clear solution was added overnight. The solvent was removed, and water (50 mL) added
dropwise to the above complex in acetonitrile (5 mL). White and extracted with CH2Cl2 (4 × 50 mL). The crude product was
purified on a silica gel column using a mixture of MeCN/H2O/
KNO3 (saturated) (40/3/1) as the eluent. After purification, the
solvent was evaporated, the solid dissolved in a minimum amount
of water, and a saturated aqueous solution of NH4PF6 was added.
The mixture was extracted with CH2Cl2, and the combined organic
phase was washed with a diluted aqueous Na2CO3 solution and
then dried over Na2SO4. Solvent was removed under reduced
pressure to afford 63% yield of the desired product. ESI-MS
(M-PF6 + Na) 1335.63 m/z.

Absorption and emission spectral


measurements
Sample solutions of the [Ru(NN)3–T]2+ complexes and the quinones
were freshly prepared for each measurement. The [Ru(NN)3–T]2+
concentration was fixed at 2 × 10-5 M and the quinone concentra-
tion was varied between 2 × 10-5 and 2 × 10-4 M. The same sample
solutions were used for both absorption and emission studies. The
absorption spectrum was recorded using an Analtik-Jena Specord
S100 spectrometer and the emission spectrum using a JASCO
spectrofluorometer. All the sample solutions used for emission
measurements were deaerated for ~ 30 min using dry N2 gas
purging by keeping solutions in cold water to ensure that there
was no change in the volume of the solution. All measurements
were carried out at room temperature.

Electrochemical measurements
Electrochemical measurements were carried out using cyclic
Scheme 1. Structure of ruthenium (II) complexes and quinones used in this study. voltammetric technique with a EG&G Princeton Applied

wileyonlinelibrary.com/journal/luminescence Copyright © 2013 John Wiley & Sons, Ltd. Luminescence 2013
A photophysic approach

Research Potentiostat/Galvanostate Model 273A. Measurements straight line with an intercept of unity in all quenching studies. The
were taken in acetonitrile using a platinum electrode as the Stern–Volmer plots for the quenching of [Ru(NN)3–T]2+ with
working electrode with 0.1 M tert-butyl ammonium perchlorate quinones are shown in Figs S8-S10. The slope of this plot, the
as the supporting electrolyte and platinum wire as the auxiliary Stern–Volmer constant, Ksv is related to kq through equation (4).
electrode. The separation between the cathodic and anodic
peaks and the relative intensities of the cathodic and anodic
currents were taken as criteria for reversibility. Results and discussion
[Ru(NN)3]2+ complexes carrying tyrosine moiety ([Ru(bpy)3–T]2+,
Determination of binding constants using [Ru(dmbpy)3–T]2+ and [Ru(dcet)3–T]2+) were synthesized, as
given in the Experimental section and characterized using mass
absorption and emission techniques
spectra (22). The structures of complexes are shown in Scheme 1
a ) for the binding of [Ru(NN)3–T]
The binding constant (Kabs 2+
com- and the mass spectra are shown in Figs S1–S3.
plexes with quinones was evaluated using Benesi–Hildebrand
method (equation (1)) (55).
h 2þ i Steady-state absorption and emission prop-
1=ΔA ¼ 1=K a abs Δε RuðNNÞ3 –T þ 1=Δε½Q (1) erties of [Ru(NN)3–T]2+ complexes
The 1MLCT state of the parent complex [Ru(bpy)3]2+ has absorp-
Here, ΔA is the change in the absorbance of the [Ru(NN)3–T] 2+ tion in the visible region at 450 nm. Introduction of a tyrosine
complex on the addition of quinone. Δε is the difference in the moiety via amide linkage produces a red-shift in the absorption
molar extinction coefficient between the free and quinone-bound maximum to a λmax of 458 nm. The electron-releasing methyl
[Ru(NN)3–T]2+ complex. [[Ru(NN)3–T]2+] is the total concentration group is a good σ donor, but a poor π acceptor and the reverse
of [Ru(NN)3–T]2+and [Q] is the total concentration of quinone. is the case with the electron-withdrawing carboxyethyl ester
We have calculated the binding constant for the system from group. Thus, the operation of σ-donor and π-acceptor properties
the luminescence intensity data using the following modified of the substituents in the ligand bipyridine increases the
Stern–Volmer equation (equation (2)) (18,56). electron density on the metal–ligand bond, thereby stabilizing
the 1MLCT state, leading to a red-shift in the absorption
log ½ðI0 –IÞ=I ¼ nlog ½Q þ log K a em (2) maximum. Therefore, the absorption, maximum corresponding
where, I0, I, [Q], Kemand n are the emission intensity in the ab- to the MLCT transition of [Ru(dmbpy)3–T]2+ and [Ru(dcet)3–T]2+
a
sence of quencher, emission intensity in the presence of is shifted towards the red region with respect to the [Ru(bpy)
3–T]
2+
quencher, concentration of quinone, binding constant and complex. The overlayed spectra of [Ru(bpy)3]2+, [Ru(bpy)
stoichiometric ratio, respectively. 3–T] 2+
, [Ru(dmbpy)3–T]2+ and [Ru(dcet)3–T]2+ are shown in Fig. 1
(a). The respective absorption λmax values are shown in Table S1.
The emission maximum of [Ru(bpy)3]2+ in CH3CN is 612 nm.
Determination of luminescence quenching Introduction of a tyrosine moiety via amide linkage produces a
constant kq red-shift in the emission maximum to a λmax of 644 nm. Because
of the σ electron donation capacity of methyl group, the energy
The observed quenching rate constant kq for the oxidative gap between the ground and 3MLCT states decreases. Therefore,
quenching of [Ru(bpy)3–T]2+ with quinones used in the study was a red-shift is observed when a –CH3 group is introduced into
obtained from the Stern–Volmer plots using equation (3)(56). bpy. However, when the carboxyethyl ester group is introduced,
Io =I ¼ 1 þ K sv ½Q (3) the ligand becomes highly π accepting. Therefore, the energy
gap between 3MLCT and the ground state increased. Thus, for
Ksv ¼ kq τ (4) the complex [Ru(dcet)3–T]2+, a blue-shift to the tune of 22 nm
is observed compared with [Ru(dmbpy)3–T]2+. The normalized
where, Ksv, kq and τ are the Stern–Volmer constant, quenching con- emission spectra of four Ru(II) complexes are shown in Fig. 1
stant and excited sate lifetime, respectively. The plot of Iο/I vs [Q] is a (b). The respective emission λmax values are given in Table S1.

2+ 2+ 2+ 2+
Figure 1. (a) Absorption spectra and (b) normalized emission spectra of [Ru(bpy)3] (----), [Ru(bpy)2C–T)] (—), [Ru(dmbpy)2(C–T)] (- - - ) and [Ru(dcEt)2(C–T)] (-.-.-).

Luminescence 2013 Copyright © 2013 John Wiley & Sons, Ltd. wileyonlinelibrary.com/journal/luminescence
P. M. Mareeswaran et al.

Excited state lifetime and redox properties Study of the binding of *[Ru(NN)3–T]2+ with
The excited state lifetime of parent [Ru(bpy)3]2+ in CH3CN is 850 quinones using absorption spectral technique
ns. Introduction of a tyrosine moiety via an amide group The changes in the absorption spectra of [Ru(NN)3–T]2+ com-
increases the excited state lifetime of the complex (1250 ns). plexes in the presence of increasing concentrations of quinones
However, introduction of an electron-releasing group in the ligand are shown in Figs 3, S4–S6. There is a slight change in the
destabilizes the 3MLCT state by higher σ-donating ability and poor absorption intensity of MLCT of [Ru(NN)3–T]2+, which is
π-accepting ability. Thus lifetime is reduced. Even though it has sufficient for the determination of binding constant for the
higher lifetime than the parent complex [Ru(bpy)3]2+, it has binding of quinone with [Ru(NN)3–T]2+ complexes. Li et al.
lifetime of 985 ns which is 265 ns less than that of the similar (57–59) have already established the importance of the π–π
complex [Ru(bpy)3–T]2+. However, introduction of the carboxyethyl stacking between phenols and [Ru(bpy)3]2+. We have used
ester group into the 2,2′-bipyridine ligand increases the π-accepting Benesi–Hildebrand method to calculate the binding constant.
capacity. It stabilizes the system up to 1350 ns, which is 100 ns The binding constants (Kabs a ) are given in Table 1. The
higher than the [Ru(bpy)3–T]2+ and of 500 ns higher than the Benesi–Hildebrand plots are shown in Figs S7–S9. From the
parent complex. The respective excited state lifetime values are binding constant values it can be inferred that the efficiency
given in Table S1. of binding is moderate and similar to all quinones in the
The parent complex, [Ru(bpy)3]2+, has an oxidation potential of range of 9–33 M-1.
+1.26 V and introduction of the tyrosine moiety via amide bonding
increases the oxidation potential to +1.39 V. However, the introduc-
tion of substituents into the 2,2′-bipyridne ligand produces little shift
Binding and oxidative quenching of
in the oxidation potential of the Ru(II) complex. The complex
[Ru(dmbpy)3–T]2+ has an oxidation potential of +1.37 V and *[Ru(NN)3–T]2+ with quinones using
[Ru(dcEt)3–T] 2+ has an oxidation potential of +1.29 V. The cyclic emission spectral technique
voltammogram of [Ru(dmbpy)3–T]2+ is shown in Fig. 2.
The increase in the concentration of quinones quenched
the emission intensity of [Ru(NN)3–T]2+ complexes. The
quenching of [Ru(NN)3–T]2+ complexes with quinones is
shown in Figs 4, S10–S12. The binding constant (Kem a ) for
the binding of quencher with the sensitizer is calculated using
the modified Stern–Volmer equation (18). The modified Stern–
Volmer plots are shown in Figs S13–S15. The binding constant
values are given in Table 1. In this case, the binding efficiency
for all five quinones is similar with all three [Ru(NN)3–T]2+
complexes. The bimolecular quenching rate constants, kq, of the
*[Ru(NN)3–T]2+ complex by quinones in CH3CN were calculated
using the Stern–Volmer equation (51). Stern–Volmer plots are
shown in Figs S16–S18. From the slope of the Stern–Volmer
plots, the bimolecular quenching rate constants, kq, are
calculated and these values are given in Table 2.
Although the quenching constant kq is similar for both
[Ru(bpy)3–T]2+ and [Ru(dmbpy)3–T]2, the value is one order
higher for [Ru(dcet)3–T]2+. This is because of the favorable
ΔG0 value for [Ru(dcet)3–T]2+. Although there is a possibility
of quenching due to energy transfer, it has already been
2+
Figure 2. Cyclic voltagram of [Ru(dmbpy)2(C–T)] . established that quenching of the emission of [Ru(NN)3–T]2+

-5 -5 -4 -5 -4
Figure 3. Absorption spectrum of Ru(dcet)3–T (2 × 10 M) with incremental addition of quinones (a) BQ (2 × 10 to 2 × 10 M), (b) NQ (2 × 10 to 2 × 10 M).

wileyonlinelibrary.com/journal/luminescence Copyright © 2013 John Wiley & Sons, Ltd. Luminescence 2013
A photophysic approach

Table 1. Binding constants of [Ru(NN)3–T] complexes with quinones using absorption (Kabs -1 em -1
a , M ) and emission (Ka , M )
techniques

Quinones [Ru(bpy)3–T]2+ [Ru(dmbpy)3–T]2+ [Ru(dcet)3–T]2+


Kabs
a n Kem
a Kabs
a n Kem
a Kabs
a n Kem
a

BQ 9.9 ± 0.34 0.71 18.1 ± 0.43 8.3 ± 0.22 0.82 17.6 ± 0.41 10.6 ± 0.57 0.66 25.7 ± 0.91
NQ 10.2 ± 0.12 0.94 20.1 ± 0.61 9.7 ± 0.78 0.78 22.6 ± 0.77 11.1 ± 0.33 0.73 32.8 ± 1.10
TCNQ 10.7 ± 0.55 0.67 22.6 ± 0.21 10.8 ± 0.67 0.69 19.1 ± 0.24 9.8 ± 0.63 0.85 31.4 ± 0.22
DCBQ 9.6 ± 0.10 0.58 20.5 ± 0.32 10.4 ± 0.34 0.72 27.5 ± 0.54 10.1 ± 0.42 0.59 29.1 ± 0.13
TCBQ 10.5 ± 0.78 0.95 25.8 ± 0.89 11.1 ± 0.54 0.55 30.3 ± 0.32 11.7 ± 0.12 0.91 33.7 ± 0.58

-5 -5 -4 -5 -4
Figure 4. Emission spectrum of Ru(dmbpy)3–T (2 × 10 M) with incremental addition of quinones (a) BQ (2 × 10 to 2 × 10 M), (b) NQ (2 × 10 to 2 × 10 M).

Table 2. Rate constants for oxidative quenching, rate of electron transfer, equilibrium constant and free energy

Quencher Equilibrium Rate of oxidative Rate of electron log k23 ΔGο, eV


constant, Keq quenching, kq transfer, k23
[Ru(bpy)3–T]
BQ 6.6 1.2 × 109 2.9 × 109 9.46 0.13
NQ 8.4 6.5 × 109 9.6 × 108 8.98 0.29
TCNQ 7.1 2 × 1010 7.2 × 109 9.86 -0.28
DCBQ 7.6 1.9 × 109 5.9 × 109 9.77 -0.39
TCBQ 10.3 2.6 × 1010 1.2 × 1010 10.08 -0.55
[Ru(dmbpy)3–T]
BQ 5.2 8.0 × 109 8.7 × 108 8.94 0.11
NQ 6.7 4.4 × 109 3.4 × 108 8.53 0.27
TCNQ 6.9 1.0 × 109 2.6 × 109 9.43 -0.57
DCBQ 7.2 9.1 × 109 2.3 × 109 9.38 -0.30
TCBQ 9.7 8.7 × 109 3.8 × 109 9.59 -0.41
[Ru(dcet)3–T]
BQ 7.1 5.2 × 1010 2.0 × 109 9.31 0.23
NQ 8.7 3.4 × 1010 1.3 × 109 9.14 0.32
TCNQ 7.7 2.5 × 1010 6.3 × 109 9.88 -0.65
DCBQ 8.1 2.1 × 1010 7.4 × 109 9.87 -0.42
TCBQ 11.1 1.8 × 1010 4.5 × 109 9.66 -0.67

with quinones is due to ET (22). The probability of reduction of Dynamics of ET reactions of [Ru(NN)3–T]2+
oxidized [Ru(NN)3–T]2+ by intramolecular ET from a tyrosine
with quinones
moiety in acetonitrile medium is also negligible (22).
Therefore, we have assigned the quenching as ET and used The rate of ET from a donor molecule to an acceptor in a solvent
the quenching constant values for calculating rate of is controlled by the change in free energy of the reaction (ΔG0),
electron transfer. the reorganization energy (λ) and the ET distance (d) between

Luminescence 2013 Copyright © 2013 John Wiley & Sons, Ltd. wileyonlinelibrary.com/journal/luminescence
P. M. Mareeswaran et al.

the donor and the acceptor. The ET rate constant (ket) in both constants, respectively. The terms rD and rA are the radii of the
the classical and semiclassical theories can be represented by electron donor and acceptor, respectively and d is the separation
equation (5)(17)(38). distance between the donor and acceptor in the encounter
  complex. This model is most applicable in cases where the donor
ket ¼ кel υn exp ΔG # =ðRTÞ (5) and acceptor are roughly spherical, and their center-to-center
where, кel is the electronic transmission coefficient, υn the nu- distance (rDA) is large compared with the sum of the sphere radii.
clear frequency and ΔG# is the free energy of activation. When The values of rD and rA can be estimated by the semi-empirical
the ET distance, d, is kept constant, the rate of the ET process (PM6 level) molecular model (10.8 Å for [Ru(bpy)3–T]2+; 11.2 Å
is decided by ΔG0 and the reorganization energy, λ through for [Ru(dmbpy)3–T]2+ and 12.9 Å for [Ru(dcet)3–T]2+ and for
the Marcus equation (equation (6)) (17,60). quinones in the range 2.97–5.0 Å). Because the ΔG0 and λ
values are known, the rate constant for ET from the excited
 2
ΔG # ¼ λ þ ΔG0 =ð4λÞ (6) state of [Ru(NN)3–T]2+ to quinone can be calculated. In
equation (8), HDA = 2 × 103 eV, λ = 0.81–0.92 eV, ν = 1000–1500
cm1 and T = 298 K. These are the optimum values for the reaction,
Substitution of the above expression into equation (5) gives chosen by a trial and error method (63).
the basic relation for ket in terms of ΔG0 and λ: Because quenching occurs via ET, the redox quenching
h  2 i process can be discussed in terms of the mechanism shown in
ket ¼ кel υn exp  λ þ ΔG0 =ð4λRTÞ (7) Scheme 2. By applying steady-state treatments to the short-lived
species in Scheme 2, the following expression (equation (11)) for
the observed bimolecular quenching rate constant, kobs(kq) can
According to classical Marcus theory, ET can occur only at the
be derived.
intersection of the two potential energy surfaces. In such case, a
more effective route for the ET rate is derived from the semiclassical
theory, which can be represented by equation (8). k
kq ¼  12  (11)
1=2
∞   1 þ k 12 =k 23 K eq
2 S m
k et ¼ 4π =hjHDA j ð4πλo kT Þ
2
∑ e S =m! (8)
m¼0
h i where Keq is the equilibrium constant for the formation of the
exp ðλo þ ΔG° þ mhv Þ2 =4λo kT encounter complex and k12 is the rate constant for the diffusion
process to form the encounter complex. The value of k12 is calcu-
lated from equation (12) (61-63).
In equation (8) HDA is the electronic coupling coefficient
between the redox centers, the reorganization energy λ is k 12 ¼ 2RT=3000η½2 þ r D =r A þ r A =r D f (12)
composed of solvational λo and vibrational λi contributions with 1 u 2 2
where f = d∫e /kT dr/r with u = ZDZAe /DS[e /1 + Kd] e /r Kd Kr

s = λi/hν, ν is the high-energy vibrational frequency associated where K = (8πe2Nη/1000DSkT)1/2, rD and rA are the radii of the
with the acceptor and m is the density of product vibrational reactants and η is the viscosity of the medium.
levels. The terms h and k are Planck’s and Boltzmann’s The diffusion rate constant, k12, calculated according to
constants, respectively Smoluchowski (64) for non-charged molecules, has a value of
According to Rehm and Weller, the free-energy change of ET 1.9 × 1010 dm3 mol-1 s-1. Keq was estimated using the Fuoss
(ΔG°) can be calculated from equation (9) (61-62). and Eigen equation (equation (13)) (65).
ΔG° ¼ EðD=D þ Þ  EðA=A

 Eoo  e2 =aε (9)  
Þ K eq ¼ 4πNd3 =3000 expðw r =RT Þ (13)
where E(D/D+;) is the oxidation potential of donors, E(A/A), the where w r is the work required to bring the reactants to the
reduction potential of acceptor, Eo-o the lowest excited state separation distance d. Because we use neutral quenchers
energy of Ru(II) complexes, and e2/aε is a columbic term. throughout this study, wr is zero. The value of Keq is in the range
The ΔG° values thus estimated for different donor and acceptor 5.2–11.1 M-1 for the oxidative quenching of [Ru(NN)3–T]2+ with
pairs in CH3CN are given in the Table 2. quinones. Keq values given in Table 2 are close to the binding
The value of λo can be evaluated classically by using dielectric constant values obtained from absorption and emission
continuum model, equation (10). techniques (Kabs and Kem
a a ) in Table 1. Because the values of k12
  and Keq are known, the value for k23, the rate constant for the
λo ¼ e2 =4πεo ð1=2r D þ 1=2r A –1=dÞ 1=Dop –1=Ds (10)
process of ET in the encounter complex, can be calculated from
the observed kq values using equation (11).
Where e is the transferred electronic charge, εo is the permittiv- In order to treat the dynamic quenching process in terms of ther-
ity of free space, Dop and Ds are the optical and static dielectric modynamic function (ΔG0), we correlated the ET rate constant, k23,

2+
Scheme 2. Mechanism for the oxidative quenching of *[Ru(NN)3–T] with quinones.

wileyonlinelibrary.com/journal/luminescence Copyright © 2013 John Wiley & Sons, Ltd. Luminescence 2013
A photophysic approach

2. Kalyanasundaram K. Photochemistry of polypyridine and porphyrin


complexes. London: Academic Press, 1992.
3. Barber J. Photosynthetic generation of oxygen. Phil Trans B.
2008;363:2665–74.
4. Vermaas WJ, Styring S, Schröder W, Andersson B. Photosynthetic water
oxidation: the protein framework. Photosynth Res 1993;38:249–63.
5. Dau H, Andrews JC, Roelofs TA, Latimer MJ, Liang W, Yachandra VK,
et al. Structural consequences of ammonia binding to the manga-
nese center of the photosynthetic oxygen-evolving complex: an X-
ray absorption spectroscopy study of isotropic and oriented photo-
system II particles. Biochemistry 1995;34:5274–87.
6. Sun L, Burkitt M, Tamm M, Raymond MK, Abrahamsson M,
LeGourriérec D, et al. Hydrogen-bond promoted intramolecular ET
to photogenerated Ru(III): a functional mimic of tyrosine Z and
histidine 190 in photosystem II. J Am Chem Soc 1999;121:6834–42.
7. Magnuson A, Berglund H, Korall P, Hammarström L, Åkermark B,
Styring S, et al. Mimicking electron transfer reactions in photosystem
II: synthesis and photochemical characterization of a ruthenium(II)
-1 -1 0 2+
Figure 5. Plot of log k23, M s vs ΔG , eV for the oxidative quenching of [Ru(NN)3–T] tris(bipyridyl) complex with a covalently linked tyrosine. J Am Chem
complexes with quinones. Soc 1997;119:10720–5.
8. Diner BA, Bautista JA, Nixon PJ, Berthomieu C, Hienerwadel R, Britt
RD, et al. Coordination of proton and electron transfer from the
redox-active tyrosine, YZ, of photosystem II and examination of the
values estimated from the kq values with the free energy change +
electrostatic influence of oxidized tyrosine, YD[radical dot](H ). Phys
(ΔG0) of the electron transfer process (equation (7)). The plot of Chem Chem Phys 2004;6:4844–50.
log k23 vs ΔG0 is shown in Fig. 5 and the ET rate constant increases 9. Oja V, Laisk A. Photosystem II antennae are not energetically connected:
with increasing the driving force (ΔG0) of the ET reaction and at- evidence based on flash-induced O2 evolution and chlorophyll fluores-
tains saturation at high ΔGo values (0.4 eV). The values of k23 cence in sunflower leaves. Photosynth Res 2012;114:15–28.
10. Sjödin M, Irebo T, Utas JE, Lind J, Merényi G, Åkermark B, et al. Kinetic
(ket) can also be calculated using semiclassical theory from equa- effects of hydrogen bonds on proton-coupled electron transfer from
tion (8). The experimental k23 values along with the calculated ket phenols. J Am Chem Soc 2006;128:13076–83.
(solid line) values were plotted against ΔG° values (Fig. 5) for all 11. Najafpour M, Moghaddam A, Yang Y, Aro E-M, Carpentier R, Eaton-
[Ru(NN)3–T]2+ complexes. Figure 5 shows that the rate constants Rye J, et al. Biological water-oxidizing complex: a nano-sized manga-
for ET reaction of chosen redox system are in accordance with nese–calcium oxide in a protein environment. Photosynth Res
2012;114:1–13.
Rhem–Weller model. 12. Irebo T, Reece SY, Sjödin M, Nocera DG, Hammarström L. Proton-
coupled electron transfer of tyrosine oxidation: buffer dependence
and parallel mechanisms. J Am Chem Soc 2007;129:15462–4.
13. Sjödin M, Styring S, Wolpher H, Xu Y, Sun L, Hammarström L.
Conclusion Switching the redox mechanism: models for proton-coupled elec-
tron transfer from tyrosine and tryptophan. J Am Chem Soc
The tyrosine–reaction center–quinone system in PSII is success- 2005;127:3855–63.
fully mimicked by studying the ET reaction of [Ru(NN)3–T]2+ 14. Sjödin M, Styring S, Åkermark B, Sun L, Hammarström L. Proton-coupled
complexes with quinones. Modification of periphery of the 2,2′- electron transfer from tyrosine in a tyrosinerutheniumtris-bipyridine
bipyridine ligands affects the photophysical properties and complex: comparison with tyrosine Z oxidation in photosystem II. J Am
Chem Soc 2000;122:3932–6.
hence, the ET properties of [Ru(NN)3–T]2+ complexes. The
15. Nazeeruddin MK, Kalyanasundaram K. Acid-base behavior in the
binding constant values obtained from absorption and emission ground and excited states of ruthenium(II) complexes containing
techniques (Kabs
a and Kem
a ) and equilibrium constant values (Keq) tetraimines or dicarboxybipyridines as protonatable ligands. Inorg
obtained from equation (13) established the interaction of Chem 1989;28:4251–9.
quinones with [Ru(NN)3–T]2+ complexes. The quenching rate 16. Darwent JR, Kalyanasundaram K. Electron-transfer reactions of qui-
nones, hydroquinones and methyl viologen, photosensitized by
constants are varied with respect to the nature of the (elec- tris(2,2[prime or minute]-bipyridine)-ruthenium(II). J Chem Soc Fara-
tron releasing and withdrawing) substituents. The plot of day Trans 1981;77:373–82.
log k23 vs ΔG0 shows that the ET process from [Ru(NN)3]2+ 17. Juris A, Balzani V, Barigelletti F, Campagna S, Belser P, von Zelewsky
carrying a tyrosine moiety with quinones is in accordance A. Ru(II) polypyridine complexes: photophysics, photochemistry,
with the Rhem–Weller model. eletrochemistry, and chemiluminescence. Coord Chem Rev
1988;84:85–277.
18. Bae E, Choi W. Effect of the anchoring group (carboxylate vs
phosphonate) in Ru-complex-sensitized TiO2 on hydrogen produc-
tion under visible light. J Phys Chem B 2006;110:14792–9.
Acknowledgements
19. Csjernyik G, Éll AH, Fadini L, Pugin B, Bäckvall J-E. Efficient
We sincerely thank Prof. P. Ramamurthy, National Center for ruthenium-catalyzed aerobic oxidation of alcohols using a biomi-
Ultrafast Processes, University of Madras, Taramani, Chennai for metic coupled catalytic system. J Org Chem 2002;67:1657–62.
his help in time resolved measurements. We sincerely thank Dr 20. Kuang D, Ito S, Wenger B, Klein C, Moser J-E, Humphry-Baker R, et al.
High molar extinction coefficient heteroleptic ruthenium complexes
M. Vairamani, IICT, Hyderabad for his help in HR-MS. for thin film dye-sensitized solar cells. J Am Chem Soc
2006;128:4146–54.
References 21. Waern JB, Desmarets C, Chamoreau L-M, Amouri H, Barbieri A,
Sabatini C, et al. Luminescent Cyclometalated RhIII, IrIII, and (DIP)
1. Scheuring S. The supramolecular architecture of the bacterial photo- 2RuII complexes with carboxylated bipyridyl ligands: synthesis,
synthetic apparatus studied by atomic force microscopy (AFM). In: X-ray molecular structure, and photophysical properties. Inorg Chem
Aartsma T, Matysik J, editors. Biophysical techniques in photosynthe- 2008;47:3340–8.
sis. Advances in photosynthesis and respiration. 26. Dordtrecht: 22. Rajkumar E, Rajagopal S, Ramamurthy P, Vairamani M. Photophysics
Springer, 2008:1–11. of ruthenium(II) complexes carrying amino acids in the ligand 2,2′-

Luminescence 2013 Copyright © 2013 John Wiley & Sons, Ltd. wileyonlinelibrary.com/journal/luminescence
P. M. Mareeswaran et al.

bipyridine and intramolecular electron transfer from methionine to 46. Rajendran T, Thanasekaran P, Rajagopal S, Gnanaraj GA, Srinivasan C,
photogenerated Ru(III). Inorg Chim Acta 2009;362:1629–36. Ramamurthy P, et al. Steric effects in the photoinduced electron transfer
23. Geißer B, Alsfasser R. A peptide approach to covalently linked reactions of ruthenium(II)-polypyridine complexes with 2,6-disubstituted
2+
[Ru(bipy)3]2+–ferrocene and [Ru(bipy)3] –tyrosine conjugates. phenolate ions. Phys Chem Chem Phys 2001;3:2063–9.
Inorg Chim Acta 2003;344:102–8. 47. Thanasekaran P, Rajagopal S, Srinivasan C. Photoredox reactions of
24. Ozawa H, Haga M-a, Sakai K. A photo-hydrogen-evolving molecular tris(2,2′-bipyrazine)-tris(2,2′-bipyrimidine)- and tris(2,3-bis[2-pyridyl]
device driving visible-light-induced EDTA-reduction of water into pyrazine)ruthenium(II) cations with phenolate ions in aqueous ace-
molecular hydrogen. J Am Chem Soc 2006;128:4926–7. tonitrile. J Chem Soc Faraday Trans 1998;94:3339–44.
25. Striplin DR, Reece SY, McCafferty DG, Wall CG, Friesen DA, Erickson BW, 48. Thanasekaran P, Rajendran T, Rajagopal S, Srinivasan C, Ramaraj R,
et al. Solvent dependence of intramolecular electron transfer in a heli- Ramamurthy P. Marcus inverted region in the photoinduced elec-
cal oligoproline assembly. J Am Chem Soc 2004;126:5282–91. tron transfer reactions of ruthenium(II)polypyridine complexes
26. Irebo T, Zhang M-T, Markle TF, Scott AM, Hammarström L. Spanning with phenolate ions. J Phys Chem A 1997;101:8195–9.
four mechanistic regions of intramolecular proton-coupled electron 49. Rajagopal S, Gnanaraj GA, Mathew A, Srinivasan C. Excited state elec-
2+
transfer in a Ru(bpy)3 –tyrosine complex. J Am Chem Soc tron transfer reactions of tris(4,4′-dialkyl-2,2′-bipyridine)ruthenium(II)
2012;134:16247–54. complexes with phenolate ions: structural and solvent effects. J
27. Kodera Y, Hara H, Astashkin AV, Kawamori A, Ono T-a. EPR study of Photochem Photobiol A 1992;69:83–9.
trapped tyrosine Z+ in Ca-depleted photosystem II. Biochim Biophys 50. Sprintschnik G, Sprintschnik HW, Kirsch PP, Whitten DG. Photochem-
Acta 1995;1232:43–51. ical reactions in organized monolayer assemblies. 6. Preparation and
28. Svensson B, Etchebest C, Tuffery P, van Kan P, Smith J, Styring S. A model photochemical reactivity of surfactant ruthenium(II) complexes in
for the photosystem II reaction center core including the structure of the monolayer assemblies and at water-solid interfaces. J Am Chem
primary donor P680. Biochemistry 1996;35:14486–502. Soc 1977;99:4947–54.
29. Grätzel M. Solar energy conversion by dye-sensitized photovoltaic 51. McCafferty DG, Bishop BM, Wall CG, Hughes SG, Mecklenberg SL,
cells. Inorg Chem 2005;44:6841–51. Meyer TJ, et al. Synthesis of redox derivatives of lysine and their
30. Clifford JN, Palomares E, Nazeeruddin MK, Grätzel M, Nelson J, Li X, use in solid-phase synthesis of a light-harvesting peptide. Tetrahe-
et al. Molecular control of recombination dynamics in dye- dron 1995;51:1093–106.
sensitized nanocrystalline TiO2 films: free energy vs distance depen- 52. Sullivan BP, Salmon DJ, Meyer TJ. Mixed phosphine 2,2’-bipyridine
dence. J Am Chem Soc 2004;126:5225–33. complexes of ruthenium. Inorg Chem 1978;17:3334–41.
31. Saito K, Rutherford AW, Ishikita H. Mechanism of proton-coupled 53. Mecklenburg SL, McCafferty DG, Schoonover JR, Peek BM, Erickson
quinone reduction in photosystem II. Proc Natl Acad Sci U S A BW, Meyer TJ. Spectroscopic study of electron transfer in a
2013;110:954–9. trifunctional lysine with anthraquinone as the electron acceptor.
32. Hasan SS, Yamashita E, Baniulis D, Cramer WA. Quinone-dependent Inorg Chem 1994;33:2974–83.
proton transfer pathways in the photosynthetic cytochrome b6f 54. Sjodin M, Ghanem R, Polivka T, Pan J, Styring S, Sun L, et al. Tuning
complex. Proc Natl Acad Sci 2013;110:4297–302. proton coupled electron transfer from tyrosine: a competition be-
33. Johnsson Wass JRT, Ahlberg E, Panas I, Schiffrin DJ. Quantum chem- tween concerted and step-wise mechanisms. Phys Chem Chem Phys
ical modeling of the reduction of quinones. J Phys Chem A 2004;6:4851–8.
2006;110:2005–20. 55. Connors KA. Binding constants: the measurement of molecular com-
34. Ishikita H, Knapp E-W. Control of quinone redox potentials in photo- plex stabilit. New York: Wiley, 1987.
system II: electron transfer and photoprotection. J Am Chem Soc 56. Lakowicz JR. Quenching of fluorescence. In: Principles of fluorescence
2005;127:14714–20. spectroscopy. New York: Springer, 2006:277–330.
35. Lenaz G. Quinone specificity of complex I. Biochim Biophys Acta 57. Li C, Hoffman MZ, Pizzocaro C, Maihot G, Bolte M. Ground-state
1998;1364:207–21. interactions between ruthenium(II)diimine complexes and
36. Itoh S, Ohshiro Y. The chemistry of heterocyclic o-quinone cofactors. phenol and monochlorophenols in aqueous solution. Inorg
Nat Prod Rep 1995;12:45–53. Chem 1998;37:3078–82.
37. Nohl H, Jordan W, Youngman RJ. Quinones in biology: functions in 58. Li C, Hoffman MZ, Pizzocaro C, Mailhot G, Bolte M. Interactions be-
electron transfer and oxygen activation. Adv Free Radal Biol Med tween the excited states of ruthenium(II)diimine complexes and
1986;2:211–79. phenols in aqueous solution. J Phys Chem A 1998;102:7370–4.
38. Marcus RA, Sutin N. Electron transfers in chemistry and biology. 59. Cang L, Sun H, Hoffman MZ. Ground- and excited-state interactions
Biochim Biophys Acta 1985;811:265–322. between the tris(2,2′-bipyridine)ruthenium(2+) ion and phenol in
39. Kurreck H, Huber M. Model reactions for photosynthesis – photoin- aqueous solution. J Photochem Photobiol A 1997;108:129–33.
duced charge and energy transfer between covalently linked por- 60. Marcus RA. Electron transfer reactions in chemistry: theory and ex-
phyrin and quinone units. Angew Chem Int Ed 1995;34:849–66. periment (Nobel lecture). Angew Chem Int Ed 1993;32:1111–21.
40. Luque NB, Schmickler W. Are the reactions of quinones on graphite 61. Rehm D, Weller A. Kinetics of fluorescence quenching by electron
adiabatic? Electrochim Acta 2013;88:892–4. and H-atom transfer. Israel J Chem 1970;8:259–71.
41. Kasson TD, Barry B. Reactive oxygen and oxidative stress: N-formyl 62. Rehm D, Weller A. Kinetik und Mechanismus der Elektronübertragung
kynurenine in photosystem II and non-photosynthetic proteins. bei der Fluoreszenzlöschung in Acetonitril. Ber Bunsen Phys Chem
Photosynth Res 2012;114:97–110. 1969;73:834–9.
42. Eugster N, Fermín DJ, Girault HH. Photoinduced electron transfer at 63. Kavarnos GJ, Turro NJ. Photosensitization by reversible electron
liquid–liquid interfaces: dynamics of the heterogeneous photore- transfer: theories, experimental evidence, and examples. Chem Rev
duction of quinones by self-assembled porphyrin ion pairs. J Am 1986;86:401–49.
Chem Soc 2003;125:4862–9. 64. Smolunchouski MZ. Versuch einer mathematischen Theorie der
43. Szrebowaty P, Kapturkiewicz A. Free energy dependence on tris(2,2′- Koagulationskinetik kolloider Lösungen. Phys Chem 1917;92:129–68.
bipyridine)ruthenium(II) electrochemiluminescence efficiency. Chem 65. Fuoss RM. Ionic association. III. The equilibrium between ion pairs
Phys Lett 2000;328:160–8. and free ions. J Am Chem Soc 1958;80:5059–61.
44. D’Souza F, Deviprasad GR. Studies on porphyrinquinhydrone
complexes: molecular recognition of quinone and hydroquinone
in solution. J Org Chem 2001;66:4601–9.
45. Okamoto K, Fukuzumi S. Hydrogen bonds not only provide a Supporting Information
structural scaffold to assemble donor and acceptor moieties of zinc
porphyrinquinone dyads but also control the photoinduced
electron transfer to afford the long-lived charge-separated states. Additional supporting information may be found in the online
J Phys Chem B 2005;109:7713–23. version of this article at the publisher’s web-site.

wileyonlinelibrary.com/journal/luminescence Copyright © 2013 John Wiley & Sons, Ltd. Luminescence 2013
View publication stats

You might also like