Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

ARTICLE

pubs.acs.org/JPCA

Influence of Triplet State Multidimensionality on Excited State


Lifetimes of Bis-tridentate RuII Complexes: A Computational Study

Tomas Osterman, †
Maria Abrahamsson,‡ Hans-Christian Becker,‡ Leif Hammarstr€om,§ and
,†
Petter Persson*

Chemistry Department, Lund University, Box 124, SE-22100 Lund, Sweden

Physical Chemistry, Department of Chemical and Biological Engineering, Chalmers University of Technology, SE-41296 G€oteborg,
Sweden
§
Chemical Physics, Department of Photochemistry and Molecular Science, Uppsala University, Box 532, SE-75120 Uppsala, Sweden

bS Supporting Information
ABSTRACT: Calculated triplet excited state potential energy surfaces are
presented for a set of three bis-tridentate RuII-polypyridyl dyes covering a wide
range of room temperature excited state lifetimes: [RuII(tpy)2]2+, 250 ps;
[RuII(bmp)2]2+, 15 ns; and [RuII(dqp)2]2+, 3 μs (tpy is 2,20 :60 ,200 -terpyridine,
bmp is 6-(2-picolyl)-2,20 -bipyridine, and dqp is 2,6-di(quinolin-8-yl)pyridine).
The computational results provide a multidimensional view of the 3MLCT
3
MC transition for the investigated complexes. Recently reported results of
significantly prolonged 3MLCT excited state lifetimes of bis-tridentate RuII-
complexes, for example [RuII(dqp)2]2+, are found to correlate with substantial
differences in their triplet excited state multidimensional potential energy
surfaces. In addition to identification of low-energy transition paths for 3MLCT
3
MC conversion associated with simultaneous elongation of two or more RuN
bonds for all investigated complexes, the calculations also suggest significant
differences in 3MLCT state volume in the multidimensional reaction coordinate space formed from various combinations of RuN
bond distance variations. This is proposed to be an important aspect for understanding the large differences in experimentally
observed 3MLCT excited state lifetimes. The results demonstrate the advantage of considering multidimensional potential energy
surfaces beyond the FranckCondon region in order to predict photophysical and photochemical properties of bis-tridentate RuII-
polypyridyl dyes and related metal complexes.

’ INTRODUCTION metal-to-ligand charge transfer (1MLCT) state (not shown in


Tris-bidentate RuII-polypyridyl complexes, such as [RuII(bpy)3]2+ Scheme 1) leads to rapid (typically within 1 ps) intersystem
(bpy is 2,20 -bipyridine), have attracted large interest as sensitizers crossing to the lowest triplet MLCT state (3MLCT). The
3
in photochemical and photobiological electron transfer applica- MLCT state can decay directly to the ground state (GS) via
tions, due to their favorable photophysical properties, including radiative or nonradiative pathways, or through thermally
long excited state lifetimes (around 1 μs at ambient temperature activated processes proceeding via either a triplet metal
for [RuII(bpy)3]2+).16 For the formation of molecular arrays, centered state (3MC), or a higher MLCT state (MLCT’)
bis-tridentate RuII-complexes, such as [RuII(tpy)2]2+ (tpy is 2,20 : with a larger degree of mixed singlettriplet spin character. It
60 ,200 -terpyridine), are advantageous compared to tris-bidentate can be noted that Scheme 1 includes a significant stabilization
complexes as the former possess a C2 axis allowing for linear of the relaxed 3MC state relative to the relaxed 3MLCT state
assembly. However, a room temperature excited state lifetime of in accordance with recent computational results for these
250 ps makes [RuII(tpy)2]2+ less useful for many photochemical complexes.16,17
applications.7,8 Interestingly, several new bis-tridentate RuII- Experimentally, the temperature dependence of the excited
complexes with ligand bite angles closer to the ideal 180° have state lifetime can be used to obtain information about the excited
recently been presented,9,10 including a new class of complexes state processes involved in excited state decay. Typically, experi-
based on [RuII(dqp)2]2+ (dqp is 2,6-di(quinolin-8-yl)pyridine), mental data are analyzed using an Arrhenius-type relationship
that rival [RuII(bpy)3]2+ in terms of excited state properties and where the deactivation kinetics is described according to eq 1
display room temperature 3MLCT excited state lifetimes of
several microseconds.1115 Received: July 23, 2011
Scheme 1 depicts key excited state processes for RuII- Revised: November 23, 2011
polypyridyl complexes. Initial photoexcitation to a singlet Published: December 09, 2011

r 2011 American Chemical Society 1041 dx.doi.org/10.1021/jp207044a | J. Phys. Chem. A 2012, 116, 1041–1050
The Journal of Physical Chemistry A ARTICLE

Scheme 1. Schematic Illustration of Deactivation Channels of the 3MLCT State in Typical RuII Complexesa

a
The initial photoexcitation process has been omitted for clarity. Decay from the 3MLCT state can proceed either directly to the ground state through
radiative or nonradiative processes, or through thermal activation to a state of either metal centered (3MC), or metal-to-ligand charge transfer (MLCT0 ),
character.

(for data well above the solvent glass transition temperature): It is desirable to provide a more comprehensive understanding
  of the factors that govern the excited state kinetics of RuII-
ΔE1
kobs ¼ k0 þ A1 exp ð1Þ complexes, and quantum chemical calculations can contribute
RT important insights.22,23 Many calculations of RuII-polypyridyl
The first term, k0, corresponds to the intrinsic decay from the complexes have been published, including a number of investiga-
emissive state to the ground state, thus including radiative and tions of ground state properties and vertical excitations.24 Such
nonradiative components. The Arrhenius term accounts for an studies have included investigations of environmental effects
activated decay in which the complex passes from the 3MLCT including, e.g., effects of solvents25,26 and counterions,27 as well
state to some shorter-lived, upper excited state,18 vide infra. as supramolecular28 and heterogeneous interactions.29 It is
It is striking that although radiative rate constants are fairly valuable to calculate triplet state properties beyond the Franck
similar for most RuII-polypyridyl complexes, the 3MLCT excited Condon region, including characterization of triplet excited state
state lifetimes vary by more than 3 orders of magnitude at properties (3MLCT and 3MC states). This has been done for
ambient temperature. This reflects significant variations in the selected RuII-complexes,16,17,3035 as well as for other related
nonradiative decay kinetics, in particular involving the thermally transition metal complexes such as IrIII-complexes for OLED
activated processes contained in the Arrhenius term. For many applications.36 These calculations have provided support for
bis-tridentate RuII-complexes this occurs through thermal con- common beliefs that 3MC states have energy minima that have
version of the 3MLCT state to a 3MC state, from which rapid significantly longer RuN bonds compared to the 3MLCT states
nonradiative decay to the ground state occurs. The short 3MLCT of the bis-tridentate RuII-polypyridyl complexes studied here,
state lifetime of [RuII(tpy)2]2+ is generally ascribed to a weak something which has been difficult to characterize experimentally
ligand field, causing a small energy gap between the 3MLCT state because of the efficient nonradiative decay of the 3MC states.16,17
and short-lived 3MC states, which results in efficient thermally To gain further insight into the mechanistic details of 3MLCT
3
activated nonradiative decay. Recent successes to prolong MC activated transitions, theoretical information about the
3
MLCT excited state lifetimes for bis-tridentate RuII-polypyridyl reaction profiles would be very useful. However, recent DFT
complexes, mentioned above, were based on design of ligands calculations point to the difficulty of identifying formal transition
that allow for more octahedral coordination spheres, thus states for the comparatively flat excited state potential energy sur-
increasing the t2geg gap and consequently increasing the faces using standard transition state searches.17
3
MLCT3MC splitting without concomitant stabilization of In this paper, we calculate triplet state potential energy sur-
the 3MLCT state. An alternative decay pathway, sometimes faces (PESs), with the aim of providing a more comprehensive
invoked, is thermal activation to a higher triplet MLCT state picture of multidimensional 3MLCT3MC excited state deacti-
often referred to as the “fourth MLCT state” (MLCT0 in vation pathways. In particular, DFT calculations are here used to
Scheme 1), which is more short-lived than the lowest 3MLCT provide a continuous adiabatic description of the lowest triplet
state due to increased singlet character.3,12,19 Furthermore, PES. On this surface, regions corresponding to the 3MLCT and
3
there are examples of experimental results that have been MC states, which are used to describe experimental excited state
interpreted as the establishment of an equilibrium between decay mechanisms from a nonadiabatic perspective as shown in
the 3MLCT and the 3MC states.4,20,21 Scheme 1, can be distinguished on the basis of differences in their
1042 dx.doi.org/10.1021/jp207044a |J. Phys. Chem. A 2012, 116, 1041–1050
The Journal of Physical Chemistry A ARTICLE

Figure 1. Chemical structures of the studied complexes: [RuII(tpy)2]2+ (a), [RuII(bmp)2]2+ (b, bmp is 6-(2-picolyl)-2,20 -bipyridine), and [RuII(dqp)2]
2+
(c).

electronic properties. For this study, we have selected three bis- have been investigated for the [RuII(dqp)2]2+ complex, in order
tridentate RuII-polypyridyl dyes that cover a wide range of room to assess the effect of using different DFT functionals and basis
temperature excited state lifetimes: [RuII(tpy)2]2+, 250 ps; sets. The [RuII(dqp)2]2+ complex was selected for these valida-
[RuII(bmp)2]2+, 15 ns, and [RuII(dqp)2]2+, 3 μs, Figure 1. tions because of its significant potential as chromophore in dyads
and triads for vectorial electron transfer.46 It is also beneficial to
’ COMPUTATIONAL METHODS consider this experimentally interesting complex further because
earlier computational work17 has suggested that the excited state
The ground state and triplet state PESs of three bis-tridentate
energies for the lowest relaxed 3MLCT and 3MC states are close,
RuII-polypyridyl complexes have been investigated using density
making a reliable computational assignment of the excited state
functional theory (DFT) calculations. All quantum chemical cal-
energy ordering particularly challenging.
culations have been performed using the Gaussian03 program.37
Complete gas phase geometry optimizations have been per-
The calculations comprise DFT calculations using different stan-
formed for the ground, 3MLCT, and 3MC states. Both the hybrid
dard basis sets and effective core potentials for the Ru atom
provided in the Gaussian program. In particular, the widely used B3LYP functional and the PBE1PBE functional have been tested.
B3LYP38,39 and PBE1PBE4042 functionals have been used, in All states were optimized with the 6-31G(d,p)-SDD basis set
conjunction with standard Gaussian type orbital (GTO) basis combination, where the 6-31G(d,p) basis set was used for all
sets of at least double-ζ quality. The SDD Stuttgart/Dresden atoms except Ru for which SDD was employed. Selected
effective core potential (ECP) was used to provide an effective structural results are summarized in Table 1 and are presented
core potential for Ru.43 Ground state properties are calculated as three central structural parameters (R-, O-, and S-values), all
using the spin-restricted singlet formalism, while spin-unrestricted with an absolute mean deviation representing the calculated
DFT (UDFT) calculations are performed for the lowest triplet variations in the value. The R-value describes the average RuN
state calculations. This extends previous work using similar bond distance, which is relevant since shorter RuN bond
methodology to locate stationary points for RuII-complexes.16,17 distances are generally expected to result in a larger energy gap
A comparison of results obtained using different combinations of between the 3MLCT and 3MC states. Furthermore, the 3MC
functionals and basis sets is presented for the experimentally most state is expected to display significant elongation of the equatorial
interesting [RuII(dqp)2]2+ complex, in particular providing infor- RuN bond distances compared to the 3MLCT state. The
mation about the method sensitivity for calculated triplet properties. importance of particular RuN bonds for the excited state
Selected excited state electronic properties have also been properties is analyzed further below in the context of relevant
investigated for the [RuII(dqp)2]2+ complex using the B3LYP and reaction coordinates. The O-value (octahedricity-value) is a
PBE1PBE functionals, several basis sets, and a PCM solvent de- measure of the mean absolute deviation of the set of NRuN
scription for ethanol (G03 keywords iefpcm and solvent = EtOH). angles from their “ideal” octahedral values, see also further
PES scans, with a spacing of 0.02 Å or multiples thereof as discussion by Lundqvist.47 Thus, it is a measure of the angular
judged necessary by the curvature of the investigated state, have distortion of the complex away from a perfectly octahedral
been performed for the three investigated complexes with a view geometry as discussed in standard ligand field theory. As the
to investigate 3MLCT3MC rearrangement pathways. Each approach to increase the 3MLCT state lifetime has been based on
point on a given PES is obtained using a full geometry B3LYP/ increasing the ligand field splitting by increased bite angles, the
6-31G(d,p)-SDD relaxation of all structural parameters except O-value should be one of the key parameters for predicting
those fixed as the probed coordinates in the PES scan. The the 3MLCT state lifetime. Finally, the S-parameter (stacking-
resulting PESs are referred to as relaxed PESs. It can be noted that parameter) measures intramolecular ligandligand interactions
the unrestricted gas phase calculations of the PESs neglect effects in terms of the average interligand distance between the centers
from spinorbit coupling,44 and solvent dynamics45 that also play of gravity in the benzene and pyridine rings in one quinoline unit
important roles for the excited state dynamics. The accuracy of the and the benzene ring of the other quinoline unit, as previously
PESs in regions where there are nearly degenerate states may discussed by Johansson et al.48
furthermore be limited by restrictions imposed by the use of a single- The RuN bond lengths for all states are calculated to be ca.
determinant DFT description. 0.03 Å shorter with the PBE1PBE functional compared to the
B3LYP functional. This effect is well-known for ground state
’ RESULTS calculations on similar complexes, where the PBE1PBE func-
tional is typically in somewhat better agreement with experi-
Stationary Points for RuII Complexes. First, structural and mental observations. The 3MLCT state is found to be very
electronic properties of the ground, 3MLCT, and 3MC states similar to the ground state in terms of its R-, O-, and S-values for
1043 dx.doi.org/10.1021/jp207044a |J. Phys. Chem. A 2012, 116, 1041–1050
The Journal of Physical Chemistry A ARTICLE

Table 1. Calculated Structural Properties of [RuII(dqp)2]2+ for the GS (Ground State), 3MLCT, and 3MC Statesa
3 3
GS MLCT MC

method R O S R O S R O S

B3LYP/6-31G(d,p)-SDD 2.11 ( 0.02 0.1 ( 1.6 3.93 ( 0.00 2.10 ( 0.04 0.5 ( 2.3 3.90 ( 0.04 2.25 ( 0.16 1.5 ( 7.9 3.86 ( 0.05
PBE1PBE/6-31G(d,p)-SDD 2.08 ( 0.02 0.1 ( 1.2 3.69 ( 0.00 2.07 ( 0.04 0.4 ( 1.9 3.68 ( 0.02 2.21 ( 0.14 1.4 ( 7.3 3.63 ( 0.05
exptlb 2.03 ( 0.03 0.6 ( 2.0 3.38 ( 0.08
a
Distances in Å and angles in degrees. Deviations are calculated as σn-values. b Calculated from crystallographic data in ref 11.

Table 2. Calculated Electronic Properties of [RuII(dqp)2]2+ for the 3MLCT and 3MC States using Different DFT Methodsa
3 3
method MLCT MC

optimization property E spin density E spin density

B3LYP/6-31G(d,p)-SDD B3LYP/6-31G(d,p)-SDD 1.99 0.66 1.97 1.73


B3LYP/6-311G(2df,p)-SDD 1.83 0.72 1.81 1.79
B3LYP/6-31G(d,p)-SDD[EtOH] 1.94 0.62 1.92 1.76
B3LYP/6-311G(2df,p)-SDD[EtOH] 1.87 0.77 1.89 1.79
PBE1PBE/6-31G(d,p)-SDD 2.01 0.66 2.07 1.80
PBE1PBE/6-31G(d,p)-SDD PBE1PBE/6-31G(d,p)-SDD 1.98 0.73 2.04 1.78
a
Triplet state energies, E, are given relative to the ground state in eV.

both functionals. The calculations give an indication of a short- Table 3. Summary of Stationary Point (GS (Ground State),
3
ening of the average RuN bond distance (by 0.01 Å) and a MLCT, and 3MC States) Data for the Three Investigated
slightly more distorted octahedral metal coordination (through Complexesa
an increase in the O-value by ca. 0.4°) when going from the 3 3
complex quantity GS MLCT MC
ground state to the 3MLCT state. The differences for both the
choice of functional and between the ground and 3MLCT states [RuII(tpy)2]2+ E (eV) 0.00 2.19 1.91
are quite small compared to the typical accuracy of ∼0.02 Å for Q1 2.12 2.11 2.12
DFT optimizations. It can be noted that the similarity of the Q2 2.12 2.10 2.38
calculated 3MLCT state geometries to the ground state geome- R 2.08 ( 0.05 2.08 ( 0.04 2.21 ( 0.13
tries, with a tendency for slightly shorter RuN bonds in the O 2.5 ( 12.0 2.8 ( 12.7 3.4 ( 17.3
3
MLCT states, is consistent with, e.g., X-ray absorption spec-
[RuII(bmp)2]2+ E (eV) 0.00 2.09 1.95
troscopy measurements for the related [Ru(bpy)3]2+ complex.49
Q1 2.13 2.12 2.25
The structural distortions involved in reaching the optimized
3 Q2 2.13 2.12 2.42
MC state are more significant. Both DFT functionals show that
the 3MC state has significantly longer RuN distances (R-value R 2.11 ( 0.03 2.10 ( 0.04 2.26 ( 0.13
increases by more than 0.1 Å) and is more distorted (O-value O 2.2 ( 7.5 2.2 ( 7.7 3.6 ( 15.0
increases from about 0.3 to about 1.5°) compared to the ground [RuII(dqp)2]2+ E (eV) 0.00 1.99 1.96
state and 3MLCT state geometries. This is also accompanied by Q1 2.12 2.10 2.15
significantly increased variations in the R- and O-values, indica- Q2 2.12 2.13 2.47
tive of an overall less regular octahedral structure. R 2.11 ( 0.02 2.10 ( 0.04 2.25 ( 0.16
Notably, both functionals give very similar differences in RuN O 0.1 ( 1.6 0.5 ( 2.3 1.5 ( 7.9
distances between the optimized 3MLCT and 3MC states. The a
Results from B3LYP/6-31G(d,p)-SDD calculations. Q1 and Q2 are
shorter RuN bonds with PBE1PBE are accompanied by composite reaction coordinates for ligands 1 and 2, respectively. They
noticeable reductions in the ligandligand distances for the are calculated as the average of the RuN distances for the two terminal
studied states, as manifested in the S-values. It is noteworthy nitrogen atoms of each ligand, respectively. Q1 = [D(RuN1) +
that the longer RuN bonds for the 3MC state compared to the D(RuN3)]/2; Q2 = [D(RuN4) + D(RuN6)]/2. R- and O-values
ground and the 3MLCT states do not lead to increased ligand as defined above.
ligand distances, something that could be due to decreased
relevance of the Ru metal center on the ligand geometry as the electronic structure calculations have been included as well. For
RuN bonds are weakened. The B3LYP and the PBE1PBE all choices of method, the 3MLCT and 3MC states can be clearly
functionals also give similar results for the respective structural distinguished on the basis of the significant change in spin density
properties of the 3MLCT and 3MC states, relative to the ground on the Ru atom. This is consistent with expectations, since 3MC
state geometry. states formally have two unpaired electrons on the Ru metal
Excited state electronic properties have also been investigated center, while the 3MLCT states formally have one unpaired
for the [RuII(dqp)2]2+ complex using the B3LYP and PBE1PBE electron on the metal center and one on the ligands. Interest-
functionals and several basis sets, Table 2. Results from calcula- ingly, the spin density for the [RuII(dqp)2]2+ complex is, in con-
tions including a PCM solvent description for ethanol for the trast to the other complexes, almost entirely localized on one of
1044 dx.doi.org/10.1021/jp207044a |J. Phys. Chem. A 2012, 116, 1041–1050
The Journal of Physical Chemistry A ARTICLE

Figure 2. Single ligand two-dimensional triplet state PES (i.e., 1L-2D-3PES) scans for [RuII(dqp)2]2+ from a 3D-landscape (panel a) and a 2D-
isosurface (panel b) perspective, respectively. RuN reaction coordinates in Å and energies in eV. The red lines are visual guides that schematically
indicate low-energy transition paths between the triplet energy minima.

the ligands (see Supporting Information). Similar to previous This 2D-3PES comprises the simultaneous scan of two Ru
findings,17 the difference in energy between the 3MLCT and Nquinoline bond distances on a single ligand, i.e., RuN1 and
3
MC states is less than 0.1 eV with all tested methods. This small RuN2 according to the labels in Figure 2. We refer to this type
energy difference is comparable to typical errors of the method of 2D-3PES involving two terminal nitrogen atoms on a single
and thus precludes a definite assignment of which relaxed triplet ligand (L) as a 1L-2D-3PES. These scanning dimensions were
state is lower in energy. It is, however, reassuring that the relative selected to provide a realistic view of a low-energy activated
energies of these two triplet states are found to be reasonably decay, considering the geometries of the 3MLCT and 3MC states
insensitive to the choice of DFT functional and basis set. summarized in Table 3. In particular, explicit consideration of the
Calculated properties of all three complexes under investiga- stretching of two terminal RuN bonds from a single ligand
tion, using the B3LYP/6-31G(d,p)-SDD standard method, are provides a comprehensive view of the development of a 3MLCT
summarized in Table 3 for the stationary states (GS, 3MLCT, state minimum region. It can be noted that, not unexpectedly, the
3
MC). The triplet geometries have been used to identify key calculated triplet energy surface is quite flat over large areas of
structural parameters for 3MLCT3MC conversion pathways coordinate space compared to the corresponding ground state
discussed below. Particularly noticeable are, as expected, elonga- surface. This means that, for this complex, as for the subsequently
tions of the equatorial RuN bond distances of one of the investigated ones, larger areas of coordinate space can be
ligands, due to population of one antibonding eg orbital. These expected to be populated thermally, and it often becomes more
are here characterized through the introduction of two effective meaningful to talk about energy minima regions and transition
reaction coordinates, Q1 and Q2, describing the elongations for regions between energy minima regions, rather than definite
each of the two ligands, respectively. For all three complexes, the minima and transition points. The 3MLCT state region allows for
geometry of the 3MLCT state is similar to the geometry of the significant asymmetric stretching of the two RuN bonds with-
ground state, but the 3MC states have significantly elongated out reaching the 3MLCT3MC crossing region. The asym-
RuN bonds. This is consistent with a stabilization of the 3MC metric 3MLCT state minima shown here are consistent with
states following a decrease in ligand field splitting. It can be noted the previous computational findings for the [RuII(dqp)2]2
+3
that we are limiting our consideration to the most stable of the two MLCT state minimum.17 A smooth low-energy activated
calculated 3MC states discussed previously for the [RuII(dqp)2]2+ 3
MLCT3MC transition path is found when both RuN bonds
and [RuII(bmp)2]2+ complexes.17 of a single ligand are stretched. This includes a low-energy
3
The calculations of triplet minima for the three complexes MLCT3MC transition region located outside the 3MLCT
provide a coherent view of 3MLCT and 3MC states that are quite state region, and ranging between two asymmetric transition
close in relaxed energy, but where the 3MC states have structures points located at about (2.10, 2.30) and (2.30, 2.10). The lowest
that include significantly elongated equatorial RuN corre- transition points are calculated at this particular level of approx-
sponding to a reduced ligand field splitting of a MC state. These imation to be ca. 0.1 eV above the 3MLCT state minimum.
calculated minimum structures provide a starting point for ii. [RuII(tpy)2]2+. The 3MLCT3MC transition for [RuII(tpy)2]2+
identifying relevant coordinates for calculations, presented in was first investigated using a similar 1L-2D-3PES scan as
the following, of multidimensional triplet PESs designed to described above for [RuII(dqp)2]2+. The resulting PES is shown
provide more comprehensive information about 3MLCT3MC in Figure 3a, and indicates rapid conversion of the 3MLCT state
excited state pathways. above the ground state minimum, where all four RuNquinoline
Excited State Potential Energy Surfaces. i. [RuII(dqp)2]2+. bonds are about 2.1 Å, to the 3MC state already through minor
A relaxed two-dimensional triplet potential energy surface (2D- stretching (approximately 0.1 Å) of the RuN bonds. Further
3PES) for the [RuII(dqp)2]2+ complex is presented in Figure 2. analysis of the calculated structures shows a discontinuity for the
1045 dx.doi.org/10.1021/jp207044a |J. Phys. Chem. A 2012, 116, 1041–1050
The Journal of Physical Chemistry A ARTICLE

Figure 3. 2D-3PES for [RuII(tpy)2]2+. Panel a shows a landscape perspective of a 1L-2D-3PES scan. Panels b and c show landscape and isosurface
perspectives, respectively, of a 2 L-2D-3PES scan. RuN reaction coordinates in Å and energies in eV. The red lines are visual guides that schematically
indicate low-energy transition paths between the triplet energy minima.

3
MLCT3MC transition at intermediate RuN distances. This
is the result of higher dimensional effects due to spontaneous
stretching of the RuN bonds of the second ligand once the
RuN bonds of the scanned ligand are stretched by a small
amount. In order to characterize the 3MLCT3MC conversion
better, an alternative 2D-3PES was generated (Figure 3b,c) using
Q1 and Q2 as reaction coordinates. Q1 and Q2 correspond to
symmetric stretching of the two ligands as described above (giving a
2L-2D-3PES), which can be illustrated for the stationary points as
follows. The 3MLCT state is located as a small minimum region
around a point near (Q1, Q2) = (2.12, 2.12), i.e., with all four terminal
RuN bonds close to 2.12 Å. There are, in contrast, two equivalent
3
MC state minima, at (2.15, 2.38) and (2.38, 2.15). These corre-
spond to one tpy ligand having both terminal RuN bonds
stretched to 2.15 Å and the other tpy ligand having both terminal
RuN bonds stretched to 2.38 Å, cf., Table 3.
Considering the 3MLCT3MC conversion path in the
2L-2D-3PES model, there is no higher dimensional collapse of
the PES as seen for the 1L-2D-3PES. Instead, elongation of the
RuN bonds of one of the two ligands leads to a low energy
3
MLCT3MC crossing close to the 3MLCT state minimum, Figure 4. Calculated PES low-energy crossing region between ground
which results in a small calculated 3MLCT state region. As a (blue) and first triplet (red) states for [RuII(tpy)2]2+. RuN reaction
consequence, no significant 3MLCT state region is formed coordinates in Å and energies in eV.
before the crossing region to the 3MC state is reached, and
asymmetric stretching effects for a single ligand, captured by the indicate that there is indeed a low-energy crossing seam where the
1L-2D-3PES scan for [RuII(dqp)2]2+, are therefore less impor- ground state surface rises above that of the 3MC state surface at
tant for the [RuII(tpy)2]2+ complex. stretched Q1 and Q2. The lowest energy crossing point, located around
For the short-lived [RuII(tpy)2]2+ complex, fast regeneration (2.12, 2.70), is found about 0.3 eV above the 3MC minimum. A
of the ground state, in addition to rapid 3MLCT3MC conver- symmetric crossing point is found at around (2.50, 2.50), but is
sion, requires efficient decay of the 3MC state, which has been calculated to be higher in energy by more than an additional 0.2 eV.
suggested to occur nonradiatively via crossing with the ground The existence of a low-energy crossing region suggests that, once
state at longer RuN bonds for other RuII-polypyridyl com- formed, the 3MC state is indeed likely to undergo fast decay leading
plexes.34 To investigate the feasibility of such facile ground state to ground state regeneration.
recovery, the 2L-2D-3PES was extended using a coarser grid iii. [RuII(bmp)2]2+. The [RuII(bmp)2]2+ complex forms mixed
(step size in each direction 0.08 Å) to longer Q1 and Q2 5- and 6-chelate NRuN rings with each ligand and should
coordinates. The relaxed PES scan was then augmented by single therefore give a larger ligand bite angle than [RuII(tpy)2]2+, but
point singlet ground state calculations at the triplet optimized possibly not as large as for [RuII(dqp)2]2+. [RuII(bmp)2]2+ has
geometries in a search for low-energy 3MCground state cross- been found experimentally to have an excited state lifetime
ing regions. The somewhat unorthodox choice to calculate single between those of [RuII(tpy)2]2+ and [RuII(dqp)2]2+.10 A 1L-2D-
point ground state energies at relaxed triplet state points, rather 3PES scan was therefore performed also for this complex,
than the other way around, is motivated by the primary interest in Figure 5. The PES shows a qualitatively intermediate behavior
the deactivation process, i.e., where the excited state dynamics is between [RuII(tpy)2]2+ and [RuII(dqp)2]2+ such that a clearly
expected to follow low energy pathways on the excited state recognizable 3MLCT state region similar to the [RuII(dqp)2]2+
surface until a crossing region with the ground state is encountered. complex has started to form, but there is an abrupt collapse to the
The results for this scan are shown in Figure 4. These calculations lower 3MC state for stretched RuN bonds similar to what was
1046 dx.doi.org/10.1021/jp207044a |J. Phys. Chem. A 2012, 116, 1041–1050
The Journal of Physical Chemistry A ARTICLE

Table 4. Excited State Lifetimes at Room Temperature and


Parameters from Arrhenius-Type Analysis of Temperature
Dependent Emission Lifetime Data
complex τ/ns A1/s‑1 ΔE1/cm‑1

[RuII(tpy)2]2+a 0.25 1.9 x1013 1500


II
[Ru (bmp)2] 2+b
15 1.0  1015 3400
[RuII(dqp)2]2+c 3000 1.5  1010 2600
a
Data from Hecker, C. R.; Gushurst, A. K. I.; McMillin, D. R. Inorg.
Chem. 1991, 30, 538541. b Data from ref 10. c Data from ref 52.

k1 is proportional to ka in both limiting cases where kc . kb and


kb , kc.
Experimentally, the magnitude of the pre-exponential factor
for activated decay, A1, is often used to assign a deactivation
mechanism. Normally, activated decay is thought to proceed
through an irreversible surface crossing from the 3MLCT state to
a 3MC state, with a pre-exponential factor corresponding to a
typical vibrational frequency, i.e., A1 ≈ 10121014 s1. Pre-
Figure 5. Calculated 1L-2D-3PES scan for [RuII(bmp)2]2+. RuN exponential factors in the range 107108 s1 have also been
reaction coordinates in Å and energies in eV. observed, and attributed to decay via a higher MLCT state with
more pronounced singlet character (MLCT0 in Scheme 1).20,21,50
seen for [RuII(tpy)2]2+, owing to the spontaneous relaxation of In a few cases, A1-values around 1091010 s1 have been
the second ligand. As this complex is somewhat less interesting reported.12,19,21,5052 These observations have been ascribed
from an experimental point of view, attempts to fully characterize either to reversible surface crossings between the 3MLCT state
the 3MLCT3MC transition path for this complex were not and a 3MC state21,50 or between the 3MLCT state and higher,
pursued further. more short-lived 3MLCT states.12,19,51 In either case A1 equals
the decay rate constant of the higher state to the ground
’ DISCUSSION state. Experimental data for the herein investigated complexes
The large variations in observed excited state lifetimes for the (Table 4) reveal significantly different pre-exponential factors,
here investigated RuII-complexes deserve particular theoretical and that A1 for [RuII(dqp)2]2+ is significantly less than
attention. In phenomenological modeling of the experimental 10121014 s1.52
kinetics, these variations mainly manifest themselves as differ- The calculated potential energy surfaces provide a theoretical
ences in the thermally activated decay, i.e., in the Arrhenius term basis for improved understanding of excited state processes.
of eq 1. A kinetic scheme (eqs 2 and 3) is commonly used to Ideally one should be able to connect phenomenological kinetic
describe the activated processes.18 models to fundamental kinetic theory, such as transition state
theory (TST).53 In TST, the rate of a process is given by eq 654
ka
3
MLCT h 3 MC ð2Þ   ! !
kb kB T ΔS‡ ΔH ‡
k¼k exp exp  ð6Þ
kc h R RT
3
MC sf GS ð3Þ
According to this scheme, the Arrhenius term in eq 1 where k is the transmission coefficient, kB is the Boltzmann
constant, h is Planck’s constant, and ΔS‡ and ΔH‡ are the
k1 ¼ A1 expð  ΔE1 =RTÞ ð4Þ difference between the reactant state and the transition state in
entropy and enthalpy, respectively. A direct comparison between
is related to the mechanistic model presented in eqs 2 and 3 such eqs 4 and 6 cannot be done simply from the here calculated PESs,
that but various aspects of the calculated PES that are relevant for the
k1 ¼ ka ðkc =ðkb þ kc ÞÞ ð5Þ excited state kinetics are discussed below.
Experimentally, it is much harder to estimate energies of 3MC
if steady-state conditions are assumed for the 3MC state.7 states than of emissive 3MLCT states, but calculations pro-
Two limiting cases can be identified in this model. When kc . kb, vide measures for both, see Tables 2 and 3. Comparison with
the kinetics simplifies to k1 = ka, corresponding to the process previously reported data11,13,16 reveals a slight but consistent
of irreversible surface crossing followed by rapid ground state overestimation of the 3MLCT state energy, but this can, at least
recovery. If kb . kc, the kinetics is governed by the formation of partly, be attributed to the exclusion of solvent effects in the
an equilibrium between the two excited states involved such that calculations. The lack of experimental values to contrast the
k1 = (ka/kb)kc where (ka/kb) is the equilibrium constant K. In this calculated 3MC state energies against makes direct comparisons
case, the activation energy term, ΔE1, is the free energy difference precarious. Therefore, we focus on the relative measures for the
between the two states involved in the equilibrium plus any investigated series, and examine the 3MLCT3MC ΔE trend to
activation energy for surface crossing to the ground state, if such a extract information and avoid any systematic errors. The results
barrier exists.18 Thus, measured activation energies may not presented above show that the energy separation between
always reflect the same properties, but the activated decay rate the 3MC and the 3MLCT state is small for all three complexes.
1047 dx.doi.org/10.1021/jp207044a |J. Phys. Chem. A 2012, 116, 1041–1050
The Journal of Physical Chemistry A ARTICLE

Scheme 2. Schematic Illustration of the Relation between the Ligand Field Splitting at the GS (Ground State) Geometry (Left)
and Thermally Activated 3MLCT3MC Deactivation along an Effective Reaction Coordinate Q (Right)

They also suggest that the 3MC state of [RuII(dqp)2]2+ is The calculated PESs also provide information about the
destabilized compared to the 3MC states of the other two 3
MLCT3MC surface crossings. In particular, an increasingly
complexes. This agrees well with experimentally obtained ground pronounced 3MLCT3MC transition barrier for the tpy-bmp-dqp
state bond lengths and angles where [RuII(dqp)2]2+ is shown series (shown in Figures 2, 3, and 5) qualitatively correlates
to be more octahedral (see Table 3). The prediction of the well with the measured trend of increasing excited state lifetimes,
3
MLCT3MC ΔE trend also agrees well with the observed trend although it can be noted that experimental results show a larger
for excited state lifetimes, see Table 4. barrier for [RuII(bmp)2]2+ than for [RuII(dqp)2]2+. This appar-
The small calculated energy separation between the relaxed ent discrepancy can be explained by the different kinetic limits of
3
MC and 3MLCT states may at first sight appear contrary to deactivation for these two complexes, as judged from their
standard ligand field theory which suggests that the 3MC state respective pre-exponential factors A1 (eq 1). Thus, the experi-
should be significantly destabilized compared to the 3MLCT state mental activation energies do not reflect the same property and
for highly octahedral RuII-complexes such as [RuII(dqp)2]2+. cannot be compared directly. In terms of the schematic repre-
However, the calculated triplet PESs (Figures 2, 3, and 5) also sentation of the 3MLCT3MC decay illustrated in Scheme 2, the
corroborate previous calculations17 that the 3MC states are increased ligand field splitting in the vertical excitation region,
significantly displaced compared to the 3MLCT states. The and the shift in energy and position of the relaxed 3MC state
calculated 3MLCT state minima are characterized by all, or all minima, together carry over to the 3MLCT3MC crossing point.
but one, RuN bond lengths being near those of typical Using this picture, one can speculate how the ligand field splitting
ground state single RuN bond distance, ca. 2.1 Å, while the is coupled to the excited state lifetime. Short-lived complexes,
3
MC states show significant elongation of the RuN bonds such as [RuII(tpy)2]2+, would then have crossing points similar to
directly involved in the change of electronic state. Thus, while point A in Scheme 2 with low kinetic barriers for forward
the standard ligand field theory usually refers to vertical 3
MLCT3MC conversion. This would promote a large forward
excitations from the ground state, the low relaxed 3MC energy rate ka in eq 2, with an associated high rate for activated decay
is seen to arise from the significant stabilization of the 3MC according to eq 4. In contrast, long-lived complexes, such as
state at larger Q values (longer RuN bonds). This is [RuII(dqp)2]2+, would have 3MLCT3MC crossing points that
illustrated in Scheme 2 for two complexes A and B with are shifted away from the 3MLCT state minimum both in energy
similar ground state and 3MLCT state energy surfaces, but and reaction coordinate, Q, similar to crossing point B in
with an increased ligand field splitting in complex B corre- Scheme 2, resulting in slower 3MLCT state activated decay
sponding to a higher vertical 3MC state energy (left) com- (smaller ka and k1 values in eqs 2 and 5, respectively). Thus,
pared to that of complex A. The stabilization of 3MC states at the calculated PESs suggest that the relative positions of the
larger Q values results in 3MLCT3MC crossing points at inter- 3
MLCT and 3MC states in terms of both energy and location
mediate Q values marked A and B for the two complexes, along the effective reaction coordinate have significant influence
respectively. The PES plots also show that the position of the on the activated decay kinetics,16,35,36 and that neither vertical
3
MC state relative to the 3MLCT state along the effective reac- excitations, nor comparisons of calculated equilibrium state
tion coordinate is different for the three investigated complexes. energies, are sufficient by themselves to fully understand the
A large shift, as in [RuII(dqp)2]2+, is necessary in order to form an complex excited state decay dynamics of these complexes.
extended 3MLCT state region and a large activation barrier for The calculations also suggest that the shapes of the multi-
the 3MLCT3MC transition. dimensional triplet energy surfaces can influence the activated
1048 dx.doi.org/10.1021/jp207044a |J. Phys. Chem. A 2012, 116, 1041–1050
The Journal of Physical Chemistry A ARTICLE

decay kinetics through entropic factors that have often been kinetic limitations, but it is not possible to convert the calculated
neglected.18 The PES calculations indicate that the reaction PES properties into accurate quantitative predictions of experi-
coordinate volume that is available for molecular motion inside mental lifetimes. Computational errors for the calculated abso-
the 3MLCT state minimum region increases significantly from lute energy barriers are likely to be significant, e.g., for the
the short-lived [RuII(tpy)2]2+ to the long-lived [RuII(dqp)2]2+ 3
MLCT3MC conversions, considering the small energy bar-
complex. In particular, the calculations indicate the emergence of riers involved (∼0.1 eV). Experimentally, kinetic effects such as
additional low-energy vibrational modes inside the 3MLCT state explicit solvent dynamics can contribute to making direct com-
region (not directly involved in the 3MLCT3MC conversion) parisons with experiments inaccurate. Furthermore, deviations
with increasing 3MLCT state volume, such as, e.g., the asym- from the simple irreversible 3MLCT3MC surface crossing can
metric RuN stretch mode for the [RuII(dqp)2]2+ complex complicate the experimental analysis and weaken the quantitative
shown in Figure 2. This qualitatively indicates that the state’s connection between computational and experimental results.
entropy increases as the conformational space increases, with
entropic effects generally influencing rate constants via the pre-
exponential factor, cf., eq 6. Increased 3MLCT state entropy may ’ CONCLUSIONS
thus contribute to reduce the forward rate constant ka for First principles quantum chemical calculations have been used
3
MLCT3MC conversions that often determine the overall to produce multidimensional triplet excited state potential en-
excited state lifetime, e.g., when k1 = ka in the limit where kc . kb ergy surfaces for a set of three bis-tridentate RuII-polypyridyl
in the kinetic scheme given by eqs 2 and 3. The substantial complexes with very different excited state lifetimes. The calcu-
3
MLCT state volume for the [RuII(dqp)2]2+ complex could in lated PESs indicate clear qualitative differences for the triplet
principle be a contributing factor for reducing the 3MLCT3MC excited state properties of the three investigated bis-tridentate
crossing frequency (ka in eq 2). However, within the current RuII-complexes that survive beyond the FranckCondon region.
analysis of the parameters in eq 1, the experimental data show The presented energy surfaces provide a trend for the 3MLCT
3
that the crossing is reversible. Then the relatively large 3MLCT MC activation process that qualitatively agrees well with experi-
state volume will lower its free energy and therefore shift the mentally obtained excited state lifetimes, constituting a signifi-
equilibrium of eq 2 toward the 3MLCT state. Thus, in both cant addition to the understanding of the excited state decay
kinetic limits (i.e., with either irreversible or reversible 3MCLT processes.
3
MC crossing) the comparatively large 3MLCT state volume for As expected, the 3MLCT states are calculated to have energy
[RuII(dqp)3]2+ may contribute to reduce the rate of excited state minima located close to the ground state geometry. For the 3MC
deactivation. states, the calculated energy surfaces confirm the generally held
Noticeable differences in the shapes of the 3MLCT and 3MC view that these have minima that are significantly displaced to
states in the calculated multidimensional PESs for the three longer RuN bond lengths. The 3MLCT and 3MC states have
investigated complexes are, furthermore, revealed. In contrast to significantly different energy profiles, with a shallower 3MC state
the well-defined 3MLCT states' minima, the 3MC states are energy profile relating to the different nature of this electronic
calculated to be shallow states where two or more RuN bonds state. This makes for significantly larger available low-energy
3
have been significantly elongated, resulting in the presence of MC state regions in the available coordinate space. The location
much larger 3MC state low-energy regions that extend to large Q. of the 3MLCT3MC activated crossing regions also has a strong
This increases the available reaction coordinate volumes for the influence on the effective 3MLCT state hypervolume available.
3
MC states compared to the corresponding 3MLCT states in a Thus, multidimensional energy surfaces help to reconcile experi-
way that is likely to influence the excited state decay kinetics. mental information that wide ranges of excited state lifetimes can
Specifically, the 3MC states are likely to be favored statistically be observed for complexes with quite similar experimentally
over the 3MLCT states for all investigated complexes due to their observed activation energies. This illustrates the strength of using
more shallow nature. Depending on the other kinetic factors calculations to investigate multidimensional energy surfaces for
involved, this effect will either push the 3MLCT3MC equilib- light-harvesting complexes that include transient excited state
rium toward the 3MC product state, or contribute to making the structures that are difficult to fully characterize experimentally,
3
MLCT3MC forward transition an effectively irreversible and it will be interesting in future work to build on the insights
process. It is worth noting that the multidimensional energy gained here to improve further on the accuracy of the calculations
surface shape effects show up only partially in one-dimensional to see if better quantitative agreement with experimental work
schematic representations using a single effective reaction can be achieved. Such developments can include both the use of
coordinate, such as Scheme 2, while they are accentuated in the more accurate quantum chemical methods, and the performance
multidimensional reaction coordinate spaces, since the same of explicitly dynamical simulations.
effect applies to several RuN bonds.
A final observation concerning the shapes of the calculated
PES is that the shallow nature of the 3MC state for large Q-values ’ ASSOCIATED CONTENT
carries over to the prediction of a low-energy 3MC-ground state
crossing seam, shown in Figure 4 for the short-lived [RuII(tpy)2]2+
bS Supporting Information. Complete ref 37 and figure of
calculated spin densities for the 3MLCT and 3MC states for all
complex. This provides a direct path for ground state recovery, three complexes. This material is available free of charge via the
cf., eq 3. The observed singlettriplet crossing for stretched Internet at http://pubs.acs.org.
RuN coordinates is, in fact, consistent with typical bond disso-
ciation curves for a wide range of molecular systems where low- ’ AUTHOR INFORMATION
energy triplet states are expected for stretched bonds.
The calculated PES provides an improved qualitative under- Corresponding Author
standing of the 3MLCT3MC activated decay pathway and its *E-mail: Petter.Persson@teokem.lu.se.
1049 dx.doi.org/10.1021/jp207044a |J. Phys. Chem. A 2012, 116, 1041–1050
The Journal of Physical Chemistry A ARTICLE

’ ACKNOWLEDGMENT (29) Persson, P.; Lundqvist, M. J. J. Phys. Chem. B 2005, 109,


The Swedish Research Council (VR), the Swedish National 11918–11924.
(30) Alary, F.; Heully, J.-L.; Bijeire, L.; Vicendo, P. Inorg. Chem.
Supercomputer Centre (NSC), and the Lund Supercomputer 2007, 46, 3154–3165.
Centre (LUNARC) are gratefully acknowledged for generous (31) Alary, F.; Boggio-Pasqua, M.; Heully, J.-L.; Marsden, C. J.;
support. Vicendo, P. Inorg. Chem. 2008, 47, 5259–5266.
(32) Salassa, L.; Garino, C.; Salassa, G.; Gobetto, R.; Nervi, C. J. Am.
Chem. Soc. 2008, 130, 9590–9597.
’ REFERENCES (33) Salassa, L.; Garino, C.; Salassa, G.; Nervi, C.; Gobetto, R.;
(1) Gray, H. B.; Winkler, J. R. Annu. Rev. Biochem. 1996, 65, 537–561. Lamberti, C.; Gianolio, D.; Bizzarri, R.; Sadler, P. J. Inorg. Chem. 2009,
(2) Hagfeldt, A.; Gr€atzel, M. Acc. Chem. Res. 2000, 33, 269–277. 48, 1469–1481.
(3) Alstrum-Acevedo, J. H.; Brennman, M. K.; Meyer, T. J. Inorg. (34) Heully, J.-L; Alary, F.; Boggio-Pasqua, M. J. Chem. Phys. 2009,
Chem. 2005, 44, 6802–6827. 131, 184308.
(4) Campagna, S.; Puntoriero, F.; Nastasi, F.; Bergamini, G.; Balzani, V. (35) Dixon, I. M.; Alary, F.; Heully, J.-L. Dalton Trans. 2010, 39,
Top. Curr. Chem. 2007, 280, 117–214. 10959–10966.
(5) Magnuson, A.; Anderlund, M.; Johansson, O.; Lindblad, P.; (36) Yang, L.; Okuda, F.; Kobayashi, K.; Nozaki, K.; Tanabe, Y.;
Lomoth, R.; Polivka, T.; Ott, S.; Stensj€o, K.; Styring, S.; Sundstr€ om, V.; Ishii, Y.; Haga, M. Inorg. Chem. 2008, 47, 7154–7165.
Hammarstr€om, L. Acc. Chem. Res. 2009, 42, 1899–1909. (37) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.;
(6) Wagenknecht, P. S.; Ford, P. C. Coord. Chem. Rev. 2011, Robb, M. A.; Cheeseman, J. R.; Montgomery, J. A., Jr.; Vreven, T.;
255, 591–616. Kudin, K. N.; Burant, J. C.; et al. Gaussian 03, revision C.02; Gaussian,
(7) Sauvage, J.-P.; Collin, J.-P.; Chambron, J.-C.; Guillerez, S.; Inc.: Wallingford, CT, 2004.
Coudret, C. Chem. Rev. 1994, 94, 993–1019. (38) Becke, A. D. J. Chem. Phys. 1993, 98, 5648–5652.
(8) Medlycott, E. A.; Hanan, G. S. Coord. Chem. Rev. 2006, 250, (39) Lee, C. T.; Yang, W. T.; Parr, R. G. Phys. Rev. B 1988, 37,
1763–1782. 785–789.
(9) Wolpher, H.; Johansson, O.; Abrahamsson, M.; Kritikos, M.; (40) Perdew, J. P.; Burke, K.; Ernzerhof, M. Phys. Rev. Lett. 1996, 77,
Sun, L. C.; Åkermark, B. Inorg. Chem. Commun. 2004, 7, 337–340. 3865–3868.
(10) Abrahamsson, M.; Wolpher, H.; Johansson, O.; Larsson, J.; (41) Perdew, J. P.; Burke, K.; Ernzerhof, M. Phys. Rev. Lett. 1997, 78,
Kritikos, M.; Eriksson, L.; Norrby, P. O.; Bergquist, J.; Sun, L. C.; 1396.
Åkermark, B.; Hammarstr€om, L. Inorg. Chem. 2005, 44, 3215–3225. (42) Adamo, C.; Barone, V. J. Chem. Phys. 1999, 110, 6158–6169.
(11) Abrahamsson, M.; J€ager, M.; Osterman, € T.; Eriksson, L.; (43) Dolg, M.; Wedig, U.; Stoll, H.; Preuss, H. J. Chem. Phys. 1987,
Persson, P.; Johansson, O.; Becker, H.-C.; Hammarstr€om, L. J. Am. 86, 866–872.
Chem. Soc. 2006, 128, 12616–12617. (44) Nozaki, K. J. Chin. Chem. Soc. 2006, 53, 101–112.
(12) Abrahamsson, M.; Becker, H.-C.; Hammarstr€om, L.; Bonnefous, (45) Moret, M.-E.; Tavernelli, I.; Chergui, M.; Rothlisberger, U.
C.; Chamchoumis, C.; Thummel, R. P. Inorg. Chem. 2007, 46, 10354–10364. Chem.Eur. J. 2010, 16, 5889–5894.

(13) Abrahamsson, M.; J€ager, M.; Kumar, J.; Osterman, T.; Persson, P.; (46) Kumar, R. J.; Karlsson, S.; Streich, D.; Jensen, A. R.; Jager, M.;
Becker, H.-C.; Johansson, O.; Hammarstr€om, L. J. Am. Chem. Soc. 2008, Becker, H.-C.; Bergquist, J.; Johansson, O.; Hammarstr€om, L. Chem.
130, 15533–15542. Eur. J. 2010, 16, 2830–2842.
(14) Schramm, F.; Meded, V.; Fliegl, H.; Fink, K.; Fuhr, O.; Qu, Z.; (47) Lundqvist, M. J. Quantum Chemical Modeling of Dye-Sensitized
Klopper, W.; Finn, S.; Keyes, T. E.; Ruben, M. Inorg. Chem. 2009, 48, Titanium Dioxide: Ruthenium Polypyridyl and Perylene Dyes, TiO2
5677–5684. Nanoparticles, and Their Interfaces. Ph.D. Dissertation, Uppsala University,
(15) Hammarstr€om, L.; Johansson, O. Coord. Chem. Rev. 2010, 254, Sweden, October 13, 2006.
2546–2559. (48) J€ager, M.; Eriksson, L.; Bergquist, J.; Johansson, O. J. Org. Chem.
(16) Abrahamsson, M.; Lundqvist, M. J.; Wolpher, H.; Johansson, O.; 2007, 72, 10227–10230.
Eriksson, L.; Bergquist, J.; Rasmussen, T.; Becker, H.-C.; Hammarstr€om, (49) Gawelda, W.; Johnson, M.; De Groot, F. M.; Abela, R.; Bressler, C.;
L.; Norrby, P.-O.; Åkermark, B.; Persson, P. Inorg. Chem. 2008, 47, Chergui, M. J. Am. Chem. Soc. 2006, 128, 5001–5009.
3540–3548. (50) Barigelletti, F.; Juris, A.; Balzani, V.; Belser, P.; von Zelewsky, A.
(17) Borg, O. A.; Godinho, S. S. M. C.; Lundqvist, M. J.; Lunell, S.; J. Phys. Chem. 1987, 91, 1095–1098.
Persson, P. J. Phys. Chem. A 2008, 112, 4470–4476. (51) Gardner, J. M.; Abrahamsson, M.; Farnum, B. H.; Meyer, G. J.
(18) Juris, A.; Balzani, V.; Barigelletti, F.; Campagna, S.; Belser, P.; J. Am. Chem. Soc. 2009, 131, 16206–16214.
von Zelewsky, A. Coord. Chem. Rev. 1988, 84, 85–277. (52) Abrahamsson, M. Tuning of the Excited State Properties of
(19) Harriman, A.; Izzet, G. Phys. Chem. Chem. Phys. 2007, 9, Ruthenium(II)-Polypyridyl Complexes. Ph.D. Dissertation, Uppsala
944–948. University, Sweden, December 1, 2006.
(20) Barigelletti, F.; Juris, A.; Balzani, V.; Belser, P.; von Zelewsky, A. (53) Eyring, H. J. Chem. Phys. 1935, 3, 107–115.
Inorg. Chem. 1983, 22, 3335–3339. (54) Laider, K. J.; King, M. C. J. Phys. Chem. 1983, 87, 2657–2664.
(21) Allen, G. H.; White, R. P.; Rillema, D. P.; Meyer, T. J. J. Am.
Chem. Soc. 1984, 106, 2613–2620.
(22) Vlcek, A., Jr. Coord. Chem. Rev. 2000, 200202, 933–977.
(23) Cramer, C. J.; Truhlar, D. G. Phys. Chem. Chem. Phys. 2009, 11,
10757–10816.
(24) Vlcek, A., Jr.; Zalis, S. Coord. Chem. Rev. 2007, 251, 258–287.
(25) Guillemoles, J.-F.; Barone, V.; Joubert, L.; Adamo, C. J. Phys.
Chem. A 2002, 106, 11354–11360.
(26) Fantacci, S.; De Angelis, F.; Selloni, A. J. Am. Chem. Soc. 2003,
125, 4381–4387.
(27) Lundqvist, M. J.; Galoppini, E.; Meyer, G. J.; Persson, P. J. Phys.
Chem. A 2007, 111, 1487–1497.
(28) Meylemans, H. A.; Damrauer, N. H. Inorg. Chem. 2009, 48,
11161–11175.

1050 dx.doi.org/10.1021/jp207044a |J. Phys. Chem. A 2012, 116, 1041–1050

You might also like