ArticlePDF Available

Charge transfer dynamics in DNA revealed by time-resolved spectroscopy

Royal Society of Chemistry
Chemical Science
Authors:

Abstract and Figures

In the past few decades, charge transfer in DNA has attracted considerable attention from researchers in a wide variety of fields, including bioscience, physical chemistry, and nanotechnology. Charge transfer in DNA has been investigated using various techniques. Among them, time-resolved spectroscopic methods have yielded valuable information on charge transfer dynamics in DNA, providing an important basis for numerical practical applications such as development of new therapy applications and nanomaterials. In DNA, holes and excess electrons act as positive and negative charge carriers, respectively. Although hole transfer dynamics have been investigated in detail, the dynamics of excess electron transfer have only become clearer relatively recently. In the present paper, we summarize studies on the dynamics of hole and excess electron transfer conducted by several groups including our own.
This content is subject to copyright. Terms and conditions apply.
Charge transfer dynamics in DNA revealed by time-
resolved spectroscopy
Mamoru Fujitsuka*and Tetsuro Majima*
In the past few decades, charge transfer in DNA has attracted considerable attention from researchers in
a wide variety of elds, including bioscience, physical chemistry, and nanotechnology. Charge transfer in
DNA has been investigated using various techniques. Among them, time-resolved spectroscopic
methods have yielded valuable information on charge transfer dynamics in DNA, providing an important
basis for numerical practical applications such as development of new therapy applications and
nanomaterials. In DNA, holes and excess electrons act as positive and negative charge carriers,
respectively. Although hole transfer dynamics have been investigated in detail, the dynamics of excess
electron transfer have only become clearer relatively recently. In the present paper, we summarize
studies on the dynamics of hole and excess electron transfer conducted by several groups including
our own.
1. Introduction
In the past few decades, charge transfer in DNA has attracted
considerable attention from researchers in diverse elds. Under
biological conditions, DNA is continuously attacked by envi-
ronmentally generated oxidants or reductants. Oxidation of
DNA promotes DNA damage,
1,2
and reduction of DNA is part of
the repair mechanism of DNA lesions by photolyase.
35
Charge
transfer in DNA is responsible for the remote oxidation or
reduction process, i.e., the oxidation or reduction of nucleo-
bases apart from the initially oxidized or reduced nucleobase. In
addition, the electrical conductivity of DNA has long been
a research focus, since DNA exhibits a highly stacked structure
of nucleobases in duplex, which is advantageous for electrical
conduction.
6
Thus, applications of DNA to nanowires have also
started to be explored. Charge transfer in DNA is an interesting
subject for physical chemists with respect to understanding
electron transfer processes in polymeric systems and other
related topics.
Charge transfer in DNA has been investigated using various
techniques. For example, the electrical conductivity measure-
ments have shown conductivities ranging from those of an
insulator to those of a superconductor.
710
Furthermore,
product analysis of the oxidative or reductive reactions of DNA
has provided valuable information on charge transfer mecha-
nisms. Such analyses have revealed that holes and excess
Mamoru Fujitsuka received B.S.,
M.S., and Doctor degrees from
Kyoto University. Aer two years
work as a postdoc, he became
a research associate of Institute
for Chemical Reaction Science,
Tohoku University. In 2003, he
moved to the Institute of Scien-
tic and Industrial Research
(SANKEN), Osaka University, as
an associate professor. His
research interest is photochem-
istry of various supramolecules.
Tetsuro Majima received B.S.,
M.S., and Doctor degrees from
Osaka University. Aer working
at the University of Texas at
Dallas (19801982) and at the
Institute of Physical and Chem-
ical Research (RIKEN, Japan)
(19821994), he became an
associate professor of the Insti-
tute of Scientic and Industrial
Research (SANKEN), Osaka
University, and a professor in
1997. His research interests
are radiation and photo
chemistries.
The Institute of Scientic and Industrial Research (SANKEN), Osaka University,
Mihogaoka 8-1, Ibaraki, Osaka 567-0047, Japan. E-mail: fuji@sanken.osaka-u.ac.jp;
majima@sanken.osaka-u.ac.jp
Cite this: Chem. Sci.,2017,8,1752
Received 2nd August 2016
Accepted 8th December 2016
DOI: 10.1039/c6sc03428d
www.rsc.org/chemicalscience
1752 |Chem. Sci.,2017,8,17521762 This journal is © The Royal Society of Chemistry 2017
Chemical
Science
PERSPECTIVE
electrons, which are positive and negative charge carriers in
DNA, respectively, can migrate rather long distances by means
of a multistep hopping process.
11,12
Furthermore, investigations
using time-resolved spectroscopic methods have provided
information on the charge transfer dynamics in DNA, including
the rate constants for single-step tunnelling and hopping
processes (Fig. 1a and b). Using these experimental techniques,
hole transfer (HT) dynamics have been investigated in detail for
decades, whereas the dynamics of excess electron transfer (EET)
have only started to become clear more recently. Our research
group has also carried out time-resolved spectroscopic studies
on charge transfer in DNA,
13
because the dynamics parameters
can provide quantitative and useful information for practical
applications. In the present paper, we summarize the time-
resolved spectroscopic studies on HT and EET in DNA con-
ducted by several research groups including our own.
2. Hole transfer in DNA
Using transient absorption spectroscopy during laser ash
photolysis of DNA conjugated with a photosensitizing electron
acceptor, Lewis et al. reported the distance and driving force
dependencies of the hole injection processes in DNA.
1416
Based
on the charge separation rate between the singlet-excited stil-
bene dicarboxaminde and G through A : T base pairs, they
estimated the damping factor (b)ink
ET
fexp(br)tobe
0.7 ˚
A
1
, where k
ET
and rare the electron transfer rate and
distance required for single-step electron transfer, respectively.
This value is similar to those obtained by product analysis and
other time-resolved spectroscopic methods, including time-
resolved uorescence measurement and pulse radiolysis.
12,17
It
was shown that polyA or polyT between the photosensitizing
electron acceptor and G did not have a signicant eect on the
bvalue. However, the bvalue of the charge recombination
process was found to be larger, 0.9 ˚
A
1
, which was attributed
to its larger driving force. In addition, from the driving force
dependence of the charge separation and recombination
processes, Lewis et al. estimated a reorganization energy of 1.22
eV and electronic coupling of 347 cm
1
when the electron
acceptor and donor nucleobase were placed in close vicinity.
16
When the acceptor and donor were separated by two A : T base
pairs, these values changed to 1.30 eV and 25 cm
1
, respectively.
These values are similar to those reported for various electron
transfer systems in non-adiabatic conditions.
Our research group measured the transient absorption
spectra during the laser ash photolysis of hairpin DNA con-
sisting of A : T base pairs, which were conjugated with naph-
thaldiimide (NDI) and phenothiazine (PTZ) as a photosensitizing
electron acceptor and donor, respectively.
18
Upon selective exci-
tation of NDI using a nanosecond laser pulse, generation of PTZ
radical cation was conrmed within the laser pulse duration.
Although the formation dynamics of PTZ radical cation were not
observed because of the fast hole-hopping among A's, the
distance dependence of the generation yield (Fig. 2) indicated
that PTZ radical cation was generated by multistep hopping of
the hole injected from the singlet excited NDI to DNA. This is in
accordance with the conclusions derived from product analysis.
12
In addition, the A-to-A hopping rate was estimated to be 2 10
10
s
1
. Recently, Lewis and co-workers estimated the A-to-A and G-
to-G hopping rates to be 1.2 10
9
and 4.3 10
9
s
1
,respectively,
based on the direct observation of HT dynamics in a photo-
sensitizing acceptorDNAdonor system by transient absorption
spectroscopy.
19
Thus, it can be concluded that the single step
hole hopping time in A's or G's is on the order of several tens to
hundreds of picoseconds.
Insertion of other nucleobase(s) between A's or G's slows
down the hopping rate among them, because the inserted
nucleobase(s) act as a barrier for the hopping (Fig. 1). We esti-
mated the hopping rates between G's or G and C, separated by
various nucleobase(s), as summarized in Table 1.
20,21
The
hopping rate was found to depend on the type and number of
nucleobase(s) present. These ndings agree with the electron
transfer theory;
22,23
that is, the electron transfer rate depends to
a large extent on the barrier height and length between the
donor and acceptor. These studies revealed the detailed
mechanisms and rate constants of HT in DNA. It should be
emphasized that the rate constant of the charge transfer process
is also essential in certain applications such as the detection of
DNA sequences. For example, we showed that a single molec-
ular uorescence detection method can detect DNA sequences
on the basis of the HT rate in DNA.
24,25
Furthermore, we found
that titanium dioxide can be used as the photosensitizing
electron acceptor for the detection of a mismatch sequence.
26
Fig. 1 Mechanisms of HT and EET in DNA. (a) G-to-G consecutive
hole hopping, (b) GXG(X¼A, C, or T) hole tunnelling, (c) T-to-T
consecutive excess electron hopping, and (d) TYT excess electron
tunnelling. E
ox
and E
red
are oxidation and reduction potentials,
respectively. Note: E
ox
(G) < E
ox
(X), while E
red
(T) > E
red
(Y).
Fig. 2 (Left) Schematic diagram for the HT in NDI- and PTZ-conju-
gated DNA and (Right) distance (N: step number) dependence of
generation yield of PTZ radical cation. Reprinted with permission from
literature.
18
Copyright (2004) American Chemical Society.
This journal is © The Royal Society of Chemistry 2017 Chem. Sci.,2017,8,17521762 | 1753
Perspective Chemical Science
3. Excess electron transfer in DNA
Studies on EET in DNA have been carried out by means of
several experimental methods such as product analysis and
laser ash photolysis. Compared to studies on the HT in DNA,
information on the kinetics such as EET rate is limited. The
small number of studies on EET in DNA is probably due to the
rather low reduction potentials of nucleobases,
27
which limit
the availability of the photosensitizing electron donor for such
analysis. In following section, the knowledge on EET obtained
to date is summarized.
3.1 Excess electron tunnelling in DNA in low temperature
glass
Sevilla and co-workers reported the results of an electron spin
resonance (ESR) study on the EET in DNA.
28
In their system,
DNA and a non-specically intercalated electron acceptor were
irradiated by g-rays in a glassy matrix at 77 K. The excess elec-
tron statistically generated on DNA nucleobases transfers to an
electron acceptor such as mitoxantrone (MX), ethidium
bromide, 1,10-phenanthroline, and 5-nitro-1,10-phenanthro-
line within a the time-scale of minutes to weeks. This EET
process can be tracked by monitoring the signal produced from
the radical anions of nucleobases using ESR. The distance for
the electron tunnelling (Fig. 1d) was estimated to be 410 base
pairs aer 1 min at 77 K. The estimated rate was on the order of
10
0
to 10
1
s
1
, from which they estimated the bvalues of each
intercalator to be 0.851.2 ˚
A
1
which is a similar range to the
bvalues estimated for the HT process.
Sevilla and colleagues further examined the eect of temper-
ature on EET using DNA with MX as the intercalator and D
2
Oas
the solvent, which was employed to distinguish T radical anion
and protonated reduced C.
29
They found that the excess electron
tunnelling from the nucleobase to MX was the dominant process
at a temperature of <77 K. The estimated bvalues of DNA at 4 and
77 K were conrmed to be the same. This lack of temperature
dependence can be attributed to the absence of the contribution
of higher vibronic states in this temperature range. Furthermore,
the deuteration of T radical anion at the C6position was found to
take place irreversibly above 130 K, which competes with the
excess electron tunnelling and acts as an irreversible sink for the
excess electrons. Furthermore, hybridization of G and C causes
the protonation of C radical anion at the N3 position in
a reversible manner to form a stable structure.
Furthermore, in 2002, Sevilla et al. estimated the bvalue and
ET distance for some base sequences according to ESR
measurements on duplexes of polydAdT and polydIdC, which
included MX as an intercalator.
30
The bvalue and ET distance
were estimated to be 0.75 ˚
A
1
and 9.4 base pairs, respectively,
for the polydAdT duplex, and were calculated to be 0.92 ˚
A
1
and
9.5 base pairs for the polydIdC duplex. They attributed these
results to the slow deuteron transfer from I to C radical anion
forming CD radical.
Therefore, from a series of ESR studies on g-ray irradiated
DNA with an intercalator, Sevilla et al. successfully obtained some
parameters of the excess electron tunnelling in DNA, which were
comparable to those of HT in DNA. In particular, the contribu-
tion of the protonated radical anion in EET is an important
nding, which can be clearly observed with ESR measurements
but is dicult to assess using other methods. In addition, it
should be noted that the radiation chemical method is advan-
tageous in the study of EET in DNA, because this method can
easily generate the reduced forms of various organic molecules.
31
Several groups have investigated charge transport in DNA using
another important radiation chemical technique, pulse radiol-
ysis. For example, Kobayashi et al. reported the transient
absorption spectra of reduced nucleobases, which revealed the
delocalization of the excess electron over several nucleobases.
32
However, oxidation or reduction via radiation chemical method
occurs in a random manner. In addition, oxidation or reduction
of the target molecules occurs from the reaction with an initially
generated oxidized orreduced solvent species, respectively. These
characteristics of radiation chemistry limit the time resolution
for kinetic research. Our research group attempted to apply pulse
radiolysis to the study of EET in DNA to determine the rate
constant of excess electron hopping (Fig. 1c), but the rate could
not be determined because of these limitations.
33
3.2 Excess electron hopping in DNA
Studies on excess electron hopping in DNA have been carried out
by means of product analysis of DNA with a tethered photo-
sensitizing electron donor, which injects excess electron to
a certain position of DNA. In 2002, Carell and co-workers inves-
tigated the distance dependence of EET in a DNA conjugated with
reduced avin and a TT cyclobutane dimer as the electron donor
to DNA and acceptor, respectively.
34
UV irradiation to a sequence
of consecutive T's is known to generate a cyclobutane pyrimidine
dimer lesion in DNA, which can be repaired by DNA photolyase
enzymes through the electron transfer from reduced and
deprotonated FADH
cofactor to a cyclobutane dimer.
3
In this
sense, this system mimics the DNA repair process by DNA pho-
tolyase. In their system, the T dimer and avin were separated by
A : T base pairs. Photoirradiation to reduced and deprotonated
avin causes the excess electron injection to T's,which is trapped
by the T dimer aer the hopping process in DNA. The reduction
of the T dimer causes cycloreversion, which can be detected by
Table 1 The intra- and inter-strand hole hopping rates between two G
separated by A, T or C and activation energy
20,21
Sequence nk
ht
/s
1
E
a
/eV
GA
n
G 1 4.8 10
7
0.18
2 9.7 10
4
0.43
3 1.4 10
4
GT
n
G 1 4.6 10
5
0.35
2 3.6 10
4
0.50
3 9.1 10
3
GA
n
C
a
1 1.4 10
6
0.30
2 4.5 10
4
0.53
GT
n
C
a
1 1.6 10
6
0.25
2 3.1 10
4
0.50
GC
n
G
a
1 (3.64.0) 10
8
0.220.25
a
Interstrand hole hopping rate.
1754 |Chem. Sci.,2017,8,17521762 This journal is © The Royal Society of Chemistry 2017
Chemical Science Perspective
HPLC. From the yield of the cycloreversion of the T dimer, they
estimated the distance dependence of the electron transfer yield,
from which the bvalue was estimated to be 0.11 ˚
A
1
.They
attributed this small bvalue to hopping of excess electron in
DNA. The plot of ln Eagainst ln N,whereEand Nare the e-
ciency of the charge transfer and the number of steps, respec-
tively, showed a linear relationship with a slope of approximately
2, which is in accordance with the one-dimensional random
walk model (EfN
2
). Furthermore, it was conrmed that the
excess electron induced to DNA could migrate over 24 ˚
A. They
also fabricated DNA hairpins with avin and T dimer as the
donor and acceptor, respectively, and showed a similar distance
dependence of EET.
3537
Using this system of a avin donor and T dimer acceptor,
Carell and co-workers further investigated EET in a PNA : DNA
hybrid.
38
In this hybrid system, the avin donor and T dimer
acceptor were introduced to PNA and DNA chains, respectively.
They found that the distance dependence of EET eciency was
similar to that of the DNA duplex system, indicating ecient
interstrand EET. Furthermore, their PNA : DNA hybrids did not
show a directional preference of EET (i.e.,5
0to 30or 30to 50).
39
Since the split rate of the T dimer is 10
6
s
1
, they suggested that
the rate-determining step of cycloreversion may be the splitting
rather than EET in DNA, indicating the fast EET in DNA. They
conrmed this speculation with experiments using BrU, BrA, and
BrG as the acceptor of the excess electron that was injected from
the avin donor to DNA.
40
They also investigated the eect of
G : C pair insertion in an A : T tract on EET yield. They found that
the DNA with the electron acceptor with higher electron accept-
ing ability (BrU) was more sensitive to the insertion of a G : C
pair, because the higher electron accepting ability results in
faster electron trapping than EET in DNA. From these results,
they suggested that the acceptor used in studies of the EET
process should possess high electron acceptor ability in order to
make the excess electron hopping a rate-determining step. More
recently, they suggested that the hopping rate through the (A : T)
4
pair is faster than 1.8 10
7
s
1
and slower than 1.4 10
8
s
1
based on the comparison of the EET rate through DNA with the
debromination rate of BrU or cleavage rate of the T dimer.
41
In their series of studies on EET using a avin donor, Carrell
et al. also revealed that the formation of the duplex structure is
essential for EET. In addition, they conrmed that N
2
O,
a solvated electron scavenger, has no eect on the EET. From
this nding, they excluded a contribution of the solvated elec-
tron, which can be generated by electron ejection from the
photoexcited donor to the solvent mediating the negative
charge to the acceptor.
42
Thus, the EET should be mediated by
nucleobases in DNA. These are also important conclusions of
their studies.
Rokita and Ito studied the EET using product analysis of the
photolysis of DNA conjugated with N,N,N0,N0-tetramethyl-1,5-
diaminonaphthalene (TMDN) and BrU, as the photosensitizing
electron donor and electron acceptor, respectively (Fig. 3).
43,44
Photoirradiation to TMDN causes the injection of the excess
electron to DNA. The arrival of the excess electron to BrU aer
hopping through DNA generates a radical anion of BrU, which
results in decomposition of the 50neighbour by generation of
the U radical, H abstraction, and hydrolysis. The yield of the
electron transfer to BrU was evaluated by electrophoresis. In
2003, they reported the small distance dependence of the
generation yield of reduced BrU, which is in accordance with
the results obtained by Carell et al., indicating that the excess
electron transfer occurred via a hopping mechanism. Further-
more, in 2004, Rokita and Ito further claried that T transfers
the excess electron more eciently than C for the same
distance. Analysis of the isotope eect of the solvent showed
that the contribution of the C radical anion to the excess charge
transfer is limited due to the protonation of the C radical anion.
Thus, T is a primary carrier of the excess electron. Furthermore,
it was indicated that the interstrand EET is slower than the
intrastrand EET, and the electron transfer in the 30to 50direc-
tion is faster than that of the reverse direction. This trend is
opposite to that observed for the HT in DNA. They pointed out
the importance of the contribution of MO to the charge carrier;
i.e. HOMO for the HT and LUMO for the EET. These ndings
were not observed by Carell et al. because the slower splitting
step of the reduced T dimer acted as the rate-determining step.
In summary, important information on the EET in DNA has
also been obtained through studies based on product analysis.
According to these results, DNA, especially T acts as the major
carrier of the excess electron, whereas C has a minor contri-
bution because of the protonation by G, which forms a pair with
C, and surrounding water. The distance dependence of the EET
yield associated with a bvalue of 0.10.3 ˚
A
1
, indicating that the
multistep hopping of the excess electron is important in the
long-range EET as in the case of the HT. This issue was
conrmed by Carell et al. who showed that the EET yield obeys
a one-dimensional random walk model. The tendencies
observed by Rokita are quite interesting issues to be claried
through more detailed investigations. However, in contrast to
the HT the rate constant of the EET has not been determined.
For determination of the rate constant, kinetic information
based on the time-resolved measurements is essential. In the
next section, an overview of time-resolved spectroscopic studies
for the EET is provided.
3.3 Excess electron injection to DNA
The kinetic properties of excess electron injection have been
studied by means of ultrafast spectroscopic methods. In 1999,
Fig. 3 DNA conjugated with TMDN and BrU as the electron donor to
DNA and acceptor, respectively. Reprinted with permission from
literature.
43
Copyright (2003) American Chemical Society.
This journal is © The Royal Society of Chemistry 2017 Chem. Sci.,2017,8,17521762 | 1755
Perspective Chemical Science
the research groups of Lewis and Wasielewski reported the
results of a femtosecond transient absorption study on the
electron injection from photoexcited stilbenediether (Sd) at the
loop position of DNA hairpin.
45,46
They found that the excess
electron injection to T occurs within 0.2 ps when T is located
next to Sd. They also conrmed that the electron injection to C
occurred with a rate constant of 3.3 10
11
s
1
, which is slower
than that to T (>2 10
12
s
1
) due to the smaller driving force of
the excess electron injection. In addition, they determined that
the electron transfer rate from excited Sd to the I : C pair was 1.4
10
12
s
1
, which is much faster than that of the G : C pair,
because of the enhanced electron accepting ability of C due to
base pairing. They also studied the distance dependence of the
excess electron injection using DNA hairpins with noncanonical
G : G base pairs between Sd and A : T pairs (Fig. 4) to ensure
that G acts as a barrier for the excess electron injection process
without a contribution of C, which is a possible electron
acceptor for photoexcited Sd. The electron injection rate
decreased to 5.7 10
9
and 4.4 10
8
s
1
when one and two
G : G pairs, respectively, were inserted between Sd and T. They
indicated that these rates are slower than the hole injection
from the singlet excited stilbenedicarboxamide (Sa) by a factor
of approximately 25, in spite of a similar driving force for the
charge separation. They attributed the observed dierence in
the charge injection rates to the longer donoracceptor distance
or weak donorbridgeacceptor electronic coupling.
In 2009, the research groups of Lewis and Fiebig reported the
charge injection process in DNA conjugated with aminopyrene
(APy) as an end-capping group and BrU.
47
The charge separation
time from photoexcited APy to T was 0.55 ps, which decreased to
0.27 ps when BrU was located at the rst or second neighbour of
APy. The eciency of the formation of reduced BrU, which is
formed by accepting the excess electron aer EET in DNA, was
evaluated by HPLC or MALDI TOF mass spectroscopy. The
distance dependence of the generation yield of the reduced
product was obtained as shown in Fig. 5, in which the HT yield
in SaA
n
Sd is also plotted. The plots were almost parallel to
each other, supporting the hopping mechanism of the excess
electron along DNA. These results are in accordance with those
by Carell et al. and Rokita and Ito described in the former
section.
Wagenknecht and co-workers also investigated the excess
electron injection process in DNA.
48
In 2004, they reported the
photoinduced process of pyrene-conjugated U and C (PydU and
PydC, respectively).
49,50
They showed that the excess electron
injection in PydC was weak pH dependent (pH ¼5 or 11), while
the injection process in PydU was completely inhibited at
higher pH. The weak pH dependence of PydC can be explained
on the basis of the protonation of the C radical anion in the
picosecond time-scale, even under a basic aqueous condition.
The degree of the proton association to the C radical anion
aected the charge recombination rate. By contrast, in the case
of PydU, the relatively higher energy of Pyc
+
dUc
than the pyrene
excited state inhibited the formation of Pyc
+
dUc
, and excited
PydU then undergoes proton coupled electron transfer to form
Pyc
+
dU(H)c. The authors pointed out that the observed pH
dependence is important to the EET in DNA, because the
protonated C radical anion, which should be generated by its
complement base G or surrounding water, will limit or termi-
nate the EET in DNA. That is, the C radical anion cannot act as
a charge carrier.
In a subsequent study in 2005, Wagenknecht and co-workers
reported the EET in DNA using PTZ as the electron donor and
BrU as the acceptor.
51
The eciency of the EET was evaluated
according to the strand cleavage yield. They conrmed that the
A : T base pairs transport the excess electron more eciently
Fig. 4 DNA hairpins conjugated with Sd at the loop position and
schematic energy diagram for the excess electron injection from the
photoexcited Sd. Reprinted with permission from literature.
46
Copy-
right (2002) American Chemical Society.
Fig. 5 The distance dependence of the generation yield of the
reduced product. The hole transport yield in SaA
n
Sd was also
plotted. Reprinted with permission from literature.
47
Copyright (2009)
American Chemical Society.
1756 |Chem. Sci.,2017,8,17521762 This journal is © The Royal Society of Chemistry 2017
Chemical Science Perspective
than do G : C pairs due to protonation of the C radical anion.
This result is consistent with that reported by Rokita and Ito
described above.
Our group has employed oligothiophenes as the photo-
sensitizing electron donor for nucleobases, because the electron
donor ability can be controlled by changing the substituents and
number of repeating units. In addition, strong absorption bands
in the radical cation state in the visible regions are good spectral
markers for a transient absorption study. We investigated the
charge-injection ability of oligothiophenes toward nucleobases
using dyad systems, as indicated in Fig. 6.
52,53
An ethylenedioxy-
substituent was introduced to some oligothiophenes, because it
has been reported that the oligomers of ethylenedioxythiophene
(EDOT) showed higher electron donor-ability than those of non-
substituted oligothiophenes.
54
Charge injection from the singlet
excited oligothiophenes to the nucleobases was indicated by the
uorescence quenching, and conrmed by the transient
absorption measurements during femtosecond laser ash
photolysis. The observed charge separation (CS) and recombi-
nation (CR) rates and corresponding driving forces (DG
CS
and
DG
CR
) are summarized in Table 2. The results indicated that the
CS rate becomes faster with an increase in the driving force, while
the CR rate becomes slower, indicating that the CS and CR
processes are in the normal and inverted regions of Marcus
theory, respectively, as shown in eqn (1).
22,23
kET ¼ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
p
ħ2lokBT
rjVj2X
meSðSm=m!Þ
exp ðloþDGþmħhu2
4lokBT!(1)
lo¼e21
2rD
þ1
2rA
1
r1
n21
3S(2)
S¼li
ħhui:(3)
In eqn (1), l
o
is the outer sphere reorganization energy given
by eqn (2), Vis the electronic coupling, Sis the electron-vibra-
tion coupling constant given by eqn (3), and huiis the averaged
angular frequency. In eqn (2), nis the refractive index. In eqn
(3), l
i
is the inner sphere reorganization energy. By applying the
Marcus theory to the observed driving force dependence of rate
constants, parameters such as V,l
o
, and l
i
were estimated to be
0.050, 0.20, and 1.10 eV, respectively (Fig. 7). Notably, for the
hole injection process, these values have been reported to be
0.043, 0.23, and 0.99 eV, respectively.
16
Thus, a similar total
reorganization energy was conrmed for hole and excess elec-
tron injection. The slightly faster excess electron injection rate
can be mainly attributed to the larger Vvalue. These results
conrm that the nucleobases act as electron acceptor when the
appropriate photosensitizing electron donor is employed,
although lots of chemists have paid attention to the oxidation
process of nucleobases in relation to DNA damage, as discussed
in the Introduction part.
3.4 Direct evaluation of excess electron transfer dynamics in
DNA
As summarized in the former sections, the excess electron
injection process has been investigated by both product anal-
ysis and transient absorption spectroscopy. However, the EET
process in DNA has mainly been studied using product analysis.
As observed in studies of the HT process in DNA, a donorDNA
acceptor system, of which the acceptor radical anion is detect-
able with transient absorption spectroscopy, will be useful for
the detailed analysis of the EET process in DNA. From this
perspective, our research group investigated the EET in DNA
hairpin conjugated with N-methyl-N-aminoglycine (
A
Py) and
diphenylacetylene (DPA), as the photosensitizing electron
donor and electron acceptor at the loop position, respectively.
55
To enhance the generation yield of the DPA radical anion,
Fig. 6 Structures of (a) 2E,3T, and 3E, and (b) dyads 115. Reprinted
with permission from literature.
52
Copyright (2014) American Chemical
Society.
Table 2 Driving forces (DG
CS
and DG
CR
) and rate constants (k
CS
and
k
CR
) of CS and CR in dyads 115. Reprinted with permission from
literature.
52
Copyright (2014) American Chemical Society
DG
CSa
(eV) k
CSb
(s
1
)DG
CRa
(eV) k
CRc
(s
1
)
10.01 6.4 10
11
3.63 1.1 10
11
20.32 1.0 10
12
3.32 6.7 10
10
30.54 7.7 10
11
3.10 2.9 10
11
40.63 1.0 10
12
3.01 3.8 10
11
50.73 2.5 10
12
2.91 3.4 10
11
60.75
d
3.93
d
70.44
d
3.62
d
80.22 3.5 10
10
3.40 6.5 10
9
90.13 5.8 10
10
3.31 6.5 10
9
10 0.03 8.7 10
10
3.21 7.3 10
9
11 0.27
d
3.39
d
12 0.04 1.1 10
11
3.08 8.3 10
10e
13 0.26 1.1 10
12
2.86 2.8 10
11e
14 0.35 1.8 10
12
2.77 3.7 10
11e
15 0.45 2.0 10
12
2.67 3.7 10
11e
a
DG
CS
and DG
CR
were calculated as indicated in literature.
52
b
Estimated error is less than 20%.
c
Estimated error is less than 10%.
d
Not observed.
e
Rate constant of the fast component.
This journal is © The Royal Society of Chemistry 2017 Chem. Sci.,2017,8,17521762 | 1757
Perspective Chemical Science
dihydrothymine (
D
T), which acts as a barrier for rapid charge
recombination,
56
was inserted between
A
Py and T. The genera-
tion of DPA radical anion was conrmed with nanosecond laser
ash photolysis, indicating that the EET time should be shorter
than 10 ns. It was conrmed that insertion of C in T's decreased
the generation of DPA radical anion. Thus, discussion based on
product analysis can also be assessed on the basis of the yield of
DPA radical anion. Furthermore, it was conrmed that the
excess electron could migrate even when
A
Py and DPA were
separated by 10 base pairs, i.e. 34 ˚
A.
In 2011, we investigated a series of donorDNAacceptor
conjugates, in which thiophene tetramer (4T) was employed as
a photosensitizing electron donor, for evaluation of the rate of
single-step excess electron hopping among nucleobases.
57
4T
has certain advantages such as sucient excited state oxidation
potentials to reduce nucleobases and strong absorption bands
upon oxidation, which can be distinguished from the excited
state and radical anion of the electron acceptor, i.e., DPA.
53
Furthermore, for the observation of the excess electron hopping
process within the time window possible with our current
instrument, rather short DNA dumbbells including a few
nucleobases were examined (Fig. 8a). The formation of duplex
structure was conrmed by means of the CD spectra and
melting temperature measurements, and supported by molec-
ular modelling.
The transient absorption spectra of the dumbbell DNA T
3
,
obtained by selective excitation of 4T with a 400 nm femto-
second laser pulse, are indicated in Fig. 8. Immediately aer the
laser pulse, the singlet excited 4T was observed. With the decay
of the singlet excited 4T, the generation of the radical cation of
4T was conrmed, indicating the excess electron injection from
the singlet excited 4T within 10 ps. However, the generation of
the DPA radical anion exhibited a two-step rising prole. The
rising rate of the faster component was the same as decay rate of
the singlet excited 4T, while the slow component reached
maximal absorbance about 200 ps aer the laser pulse excita-
tion. The fast component can be attributed to the single-step
charge separation between singlet excited 4T and DPA, whereas
the slow component can be attributed to the formation of DPA
radical anion by the excess electron aer hopping through the T
tract (Fig. 8b). Moreover, the generation rate of DPA radical
anion by the EET process became slower with increasing
number of T's between 4T and DPA, supporting the generation
of DPA radical anion by excess electron hopping in DNA. By
applying a one-dimensional random walk model, the T-to-T
hopping rate in consecutive T's was determined to be 4.4 10
10
s
1
. As described above, Lewis et al. reported that the single-step
hopping rate of HT was 4.3 10
9
and 1.2 10
9
s
1
for G-to-G
and A-to-A, respectively.
19
Our research group found that the A-
to-A hopping rate was 2 10
10
s
1
.
58
Therefore, it became clear
that the hopping rate of the excess electron among consecutive
T's is faster than the hole hopping rate among A's and G's.
In addition to the nicked-dumbbell DNA, we conrmed the
EET in DNA hairpin, in which DPA was placed at the
loop position and N,N0-dimethylaminopyrene (
A
Py0),
Fig. 7 DG(DG
CS
and DG
CR
) dependence of k
ET
(k
CS
Cand k
CR
B).
Numbers close to the marks indicate compounds. The solid and
hollow black squares are the 2T dyads. The solid and hollow red circles
are the dyads 15. The solid and hollow green circles are corre-
sponded to the dyads 810. The solid and hollow orange circles are
corresponded to the dyads 1215. The solid blue line was calculated
by eqn (1), (2), and (3) using l
S
,l
V
,V, and ħhuiof 0.20, 1.10, 0.050, and
0.19 eV, respectively. The solid pink line was calculated using l
S
,l
V
,V,
and ħhuiof 0.23, 0.99, 0.043, and 0.19 eV, respectively. Reprinted with
permission from literature.
52
Copyright (2014) American Chemical
Society.
Fig. 8 (a) Structures of dumbbell DNAs conjugated with 4T and DPA.
(b) Expected energy diagram. (c) Transient absorption spectra of T
3
during the laser ash photolysis using 400 nm femtosecond laser
pulse. Reprinted with permission from literature.
57
Copyright (2011)
American Chemical Society.
1758 |Chem. Sci.,2017,8,17521762 This journal is © The Royal Society of Chemistry 2017
Chemical Science Perspective
a photosensitizing electron donor, was attached at the 50end.
59
In this case, the EET dynamics indicated that
A
Py0did not
behave as an end-capping group but rather seems to be inter-
calated between two terminal base pairs; Lewis et al. conrmed
a similar structure with 50-pyrenecarboxamide tethered DNA.
60
The T-to-T hopping rate was estimated to be 6.1 10
10
s
1
,
which is slightly larger than the value estimated with the above
mentioned nicked-dumbbell DNA, but the hopping rate on the
order of 10
10
s
1
was conrmed.
3.5 T-to-T hopping rate depends on the driving force for
charge injection
As introduced in the former section, we have claried the T-to-T
hopping rate by studying donorDNAacceptor systems. As
a photosensitizing electron donor, we employed 4T and
A
Py0as
discussed in the previous section, which provided T-to-T
hopping rates of 4.4 10
10
and 6.1 10
10
s
1
, respectively. Our
further studies using a trimer and dimer of EDOT (3E and 2E,
respectively) as the photosensitizing electron donor yielded
T-to-T hopping rates of 2.2 10
11
and 2.6 10
11
s
1
, respec-
tively.
61
Thus, the reported EET rate constants for consecutive
T's are in the range of 10
10
to 10
11
s
1
depending on the pho-
tosensitizing electron donors (Table 3), which provided various
driving forces for excess-electron injection to DNA. It was
conrmed that a larger driving force for excess electron injec-
tion results in a faster T-to-T hopping rate. In fact, a nearly
linear relationship was conrmed between the T-to-T hopping
rate and driving force for excess electron injection (Fig. 9). The
intercept of the linear t ((3.8 1.5) 10
10
s
1
) should corre-
spond to the hopping rate for a non-energy assisted (DG
CS
¼0)
EET in DNA, i.e., the intrinsic hopping rate. Interestingly, the
intrinsic hopping rate agrees with the reported value for the
DNA sugar backbone and base motions, which occur with
periods as short as 30 ps at 303 K,
62,63
suggesting that the excess-
electron hopping is likely dominated by the structural dynamics
of DNA.
3.6 T-to-T hopping rate in an alternating AT sequence
Excess electrons should migrate through the LUMOs of C or T in
DNA due to their relatively high reduction potentials in DNA.
According to the results of product analysis, EET has been
conrmed to be a sequence-dependent process. Furthermore, it
was conrmed that the protonated C radical anion will limit or
terminate EET, as discussed in the next section in more detail.
Thus, T should contribute to EET signicantly. Several groups,
including our own, have reported the dynamics of intrastrand
EET in DNA through A : T sequences; however, investigations
on interstrand EET in alternating A : T sequences in DNA, in
which the interaction between the LUMOs of T's does not exist,
are still limited. For example, Carell and co-workers reported an
interstrand EET in PNA : DNA double strands based on product
analysis.
38
Their qualitative results showed that an interstrand
EET can eciently proceed in PNA : DNA double strands,
indicating that the EET in PNA : DNA is somewhat inuenced
by the specic stacking situation. By contrast, for interstrand
HT in DNA, Lewis's group reported that the eciency of inter-
strand HT in DNA is lower than that of intrastrand HT by
a factor of 4, due to the lack of interaction between the HOMOs
of A's, based on femtosecond laser ash photolysis studies.
66
We obtained similar results using nanosecond laser ash
photolysis in our previous report.
18
Thus, direct measurement
of the dynamics of EET in DNA by laser ash photolysis is
essential to achieve a quantitative understanding of interstrand
EET. Here, two series of functionalized DNA oligomers, Tn and
ATn, were synthesized with a strong electron-donating photo-
sensitizer (3E) and an electron acceptor (DPA) (Fig. 10).
64
Table 3 Excess electron hopping rates
57,59,61,64,65
Sequence Donor k
hop
/s
1
Note
TT 4T 4.4 10
10
Intrastrand T-to-T
A
Py06.1 10
10
Intrastrand T-to-T
3E 2.2 10
11
Intrastrand T-to-T
2E 2.6 10
11
Intrastrand T-to-T
TAT3E 1.1 10
11
Interstrand T-to-T
TCT3E 110
11
Assuming C as an excess electron carrier
3E 4.9 10
10
Assuming intrastrand T-to-T tunneling
Fig. 9 Dependence of ln k
hop
on DG
CS
. A, B, C, and 2-Tn were
estimated from DNA conjugated with 4T,
A
Py0,3E, and 2E, respectively,
as a photosensitizing electron donor. Reprinted with permission from
literature.
61
Copyright (2016) American Chemical Society.
This journal is © The Royal Society of Chemistry 2017 Chem. Sci.,2017,8,17521762 | 1759
Perspective Chemical Science
Transient absorption measurements during femtosecond laser
ash photolysis of ATn showed the generation of DPAc
for AT3,
whereas AT4 and AT5 did not result in the generation of DPAc
.
The absence of DPAc
can be attributed to the fast CR between
3Ec
+
and DNAc
. For AT3 and T3, the generation of 3Ec
+
DNA
DPAc
was conrmed, indicating that the inter- and intrastrand
EET is strongly limited by the CR rate. By applying the random
walk model, the inter- and intrastrand excess electron hopping
rates were determined. It was conrmed that the EET rate
constant of AT3 (1.1 10
11
s
1
) is 2-times lower than that of T3
(2.2 10
11
s
1
). The slower EET rate for intrastrand EET can be
attributed to the lack of p-stacking of T's in AT3. Thus, these
results indicated that excess electron hopping is aected by the
interaction between the LUMOs of nucleotides.
3.7 Eect of G and C
It has also been noted that pH aects the dynamics of EET in
DNA. This is because the protonated C radical anion, which can
be generated by proton transfer from the complementary base,
G, or from surrounding water molecules, will limit or terminate
EET.
30,44,50,51,67
Steenken and co-workers proposed a proton-
transfer reaction pathway for the G : C base pair radical anion
(G : Cc
base pair) as shown in Fig. 11,
68,69
which is a thermo-
dynamically favourable process and has been supported by
various experimental
70,71
and theoretical studies.
7274
Although
the rate constant of proton transfer in G : Cc
base pairs (k
PT
)
was theoretically calculated to be 10
11
s
1
,
75
it has not been
determined directly. Furthermore, limited information is
available for the eect of G : C base pairs on EET dynamics. To
clarify the dynamics of EET in DNA, a femtosecond laser ash
photolysis study of a donorDNAacceptor system was carried
out (Fig. 12).
65
In the transient absorption spectra of C4, the
transient absorption band at 580 nm, which can be attributed to
G(H)
: C(H)c, the product of the proton-transfer reaction of
G:Cc
base pairs, was observed as well as 3Ec
+
. From the global
analysis, the proton transfer rate was estimated to be 2.6 10
10
s
1
, which agreed with the theoretical estimation.
76
In addition,
generation of DPA was not conrmed with C3 and C4, indi-
cating that multiple C's completely terminated the EET in DNA.
Notably, the generation of DPAc
was conrmed from anal-
ysis of the transient absorption spectra of CTT, TCT, and TTC,
indicating that the single G : C pair in T's cannot completely
terminate the EET. The intervening G : C base pair in the
consecutive T's reduced the k
ET
value to 50%, regardless of the
position of the G : C base pair in the DNA oligomers. In addi-
tion, the yield of the formation of DPAc
with respect to the
initial 3Ec
+
generation showed a decrease of 30% in CTT, TCT,
and TTC. As a role of C in EET, two possibilities are pointed out:
(1) C acts as excess electron carrier and (2) C acts as a barrier for
excess electron (Table 3).
3.8 Detection of long range EET by electrochemical method
Because of the rapid CR, long-range EET through DNA has been
dicult to conrm with laser ash photolysis study of donor
DNAacceptor systems. Therefore, to observe long-range EET
through DNA we employed electrophotochemical systems using
donorDNAAu gold systems.
77
In this system, photoexcitation
of 3E yielded a charge separated state, i.e., excess-electron
injection from the singlet excited 3E,
1
3E*, to T. The injected
excess-electron migrated through T and was then trapped by the
Au electrode to yield a photocurrent. 3Ec
+
was reduced by
ascorbic acid, AA, as an electron-donating sacrice, to regen-
erate 3E. The redox levels of the components are presented in
Fig. 13. The signicant sequence dependence of photocurrent
generation suggests that T-to-T hopping is a dominant mecha-
nism for EET in DNA. Using this system, we conrmed EET
through DNA up to 40.8 ˚
A. We further conrmed the sequence
dependence of the process, as shown in Fig. 14. From the
experiments on DNAs with an alternating A : T sequence and
consecutive G : C sequence, the hopping models described in
the sections above were supported.
Fig. 10 Structures of DNA oligomers. The gap between the 50and 30
indicates a missing phosphate linker between two nucleotides in
nicked dumbbell structure. Reprinted with permission from litera-
ture.
64
Copyright (2015) John Wiley and Sons.
Fig. 11 Proposed proton-transfer reaction pathway for G : Cc
base
pair. Reprinted with permission from literature.
65
Copyright (2015)
American Chemical Society.
Fig. 12 Structures of 3E, DPA, and DNA oligomers (C3, C4, T3, CTT,
TCT, and TTC). The gap between the 50and 30indicates a missing
phosphate linker between two nucleobases in nicked dumbbell
structure. Reprinted with permission from literature.
65
Copyright
(2015) American Chemical Society.
1760 |Chem. Sci.,2017,8,17521762 This journal is © The Royal Society of Chemistry 2017
Chemical Science Perspective
4. Conclusions and remarks
In this perspective article, the EET in DNA were summarized
and a brief summary of the HT process was provided. Compared
to HT in DNA, our understanding of the EET process is still
limited, although the results of recent detailed investigations
have claried various aspects of EET. The following questions
point out examples of unknown issues about EET in DNA: (1)
how do multiple nucleobase insertions (i.e., G's between T's)
alter the EET hopping rate? We have already pointed out that
single C between T's does not terminate EET, but multiple C's
can terminate. How about G? (2) How far can the excess electron
migrate? In laser ash photolysis experiment, fast CR limits the
EET distance. But as indicated by the photoelectrochemical
method, excess electron potentially migrates over several tens
angstrom. Is EET over hundred angstrom possible? (3) To what
extent do mismatches change the EET hopping rate? Is the
mismatch eect similar to that for HT? (4) What eect does
DNA conformation, such as, A, B, C, and Z, have on the EET
rate? (5) What eect does protein association have on EET in
DNA? For example, would the EET rate in the chromosomes in
cellular nucleus be faster or not? This point should be essential
in DNA damage and repair in cell as indicated by the electro-
chemical studies.
78,79
These questions are fundamental for EET
in DNA. We hope that the detailed processes of EET in DNA will
be claried in the near future.
Note added after rst publication
This article replaces the version published on 23rd December
2016, in which incorrect contact details for the rst author were
presented through editorial error.
Acknowledgements
We are grateful to a number of collaborators, especially Prof.
Kiyohiko Kawai, Prof. Takashi Tachikawa, Dr Man Jae Park and
Dr Shih-Hsun Lin, SANKEN, Osaka University. This work has
been partly supported by a Grant-in-Aid for Scientic Research
(Projects 25220806, 25288035, and others) from the Ministry of
Education, Culture, Sports, Science and Technology (MEXT) of
Japanese Government.
References
1 C. J. Burrows and J. G. Muller, Chem. Rev., 1998, 98, 1109.
2 B. Armitage, Chem. Rev., 1998, 98, 1171.
3 T. Carell, Angew. Chem., Int. Ed. Engl., 1995, 34, 2491.
4 A. Sancar, Biochemistry, 1994, 33,2.
5 J. S. Taylor, Acc. Chem. Res., 1994, 27, 76.
6 D. D. Eley and D. I. Spivey, Trans. Faraday Soc., 1962, 58, 411.
7 H.-W. Fink and C. Schonenberger, Nature, 1999, 398, 407.
8 D. Porath, A. Bezryadin, S. de Vries and C. Dekker, Nature,
2000, 403, 635.
9 S. Priyadarshy, S. M. Risser and D. N. Beratan, J. Phys. Chem.,
1996, 100, 17678.
10 M. Taniguchi and T. Kawai, Phys. E., 2006, 33,1.
11 B. Giese, Acc. Chem. Res., 2000, 33, 631.
12 B. Giese, J. Amaudrut, A.-K. Kohler, M. Spormann and
S. Wessely, Nature, 2001, 412, 318.
13 M. Fujitsuka and T. Majima, Phys. Chem. Chem. Phys., 2012,
14, 11234.
14 F. D. Lewis, T. Wu, Y. Zhang, R. L. Letsinger, S. R. Greeneld
and M. R. Wasielewski, Science, 1997, 277, 673.
15 F. D. Lewis, T. Wu, X. Liu, R. L. Letsinger, S. R. Greeneld,
S. E. Miller and M. R. Wasielewski, J. Am. Chem. Soc., 2000,
122, 2889.
16 F. D. Lewis, R. S. Kalgutkar, Y. Wu, X. Liu, J. Liu, R. T. Hayes,
S. E. Miller and M. R. Wasielewski, J. Am. Chem. Soc., 2000,
122, 12346.
17 K. Kawai, T. Takada, S. Tojo, N. Ichinose and T. Majima, J.
Am. Chem. Soc., 2001, 123, 12688.
18 T. Takada, K. Kawai, X. Cai, A. Sugimoto, M. Fujitsuka and
T. Majima, J. Am. Chem. Soc., 2004, 126, 1125.
Fig. 13 DNA sequences and energetic diagram of photocurrent
generation of DNA-modied Au electrode in the presence of ascorbic
acid (AA). Potential versus Ag/AgCl. Reprinted with permission from
literature.
77
Copyright (2016) John Wiley and Sons.
Fig. 14 Normalized photocurrent density of DNA-modied electrode.
Error bars represent the standard deviation from ve experiments.
Reprinted with permission from literature.
77
Copyright (2016) John
Wiley and Sons.
This journal is © The Royal Society of Chemistry 2017 Chem. Sci.,2017,8,17521762 | 1761
Perspective Chemical Science
19 S. M. M. Conron, A. K. Thazhathveetil, M. R. Wasielewski,
A. L. Burin and F. D. Lewis, J. Am. Chem. Soc., 2010, 132, 14388.
20 T. Takada, K. Kawai, M. Fujitsuka and T. Majima, Chem.
Eur. J., 2005, 11, 3835.
21 Y. Osakada, K. Kawai, M. Fujitsuka and T. Majima, Proc.
Natl. Acad. Sci. U. S. A., 2006, 103, 18072.
22 R. A. Marcus, Annu. Rev. Phys. Chem., 1964, 15, 144.
23 R. A. Marcus and N. Sutin, Biochim. Biophys. Acta, Rev.
Bioenerg., 1985, 811, 265.
24 T. Takada, M. Fujitsuka and T. Majima, Proc. Natl. Acad. Sci.
U. S. A., 2007, 104, 11179.
25 T. Takada, Y. Takeda, M. Fujitsuka and T. Majima, J. Am.
Chem. Soc., 2009, 131, 6656.
26 T. Tachikawa, Y. Asanoi, K. Kawai, S. Tojo, A. Sugimoto,
M. Fujitsuka and T. Majima, Chem.Eur. J., 2008, 14, 1492.
27 C. A. M. Seidel, A. Schulz and M. H. M. Sauer, J. Phys. Chem.,
1996, 100, 5541.
28 A. Messer, K. Carpenter, K. Forzley, J. Buchanan, S. Yang,
Y. Razskazovskii, Z. Cai and M. D. Sevilla, J. Phys. Chem. B,
2000, 104, 1128.
29 Z. Cai, Z. Gu and M. D. Sevilla, J. Phys. Chem. B, 2000, 104,
10406.
30 Z. Cai, X. Li and M.D. Sevilla, J. Phys. Chem. B, 2002, 106, 2755.
31 M. Fujitsuka and T. Majima, J. Phys. Chem. Lett., 2011, 2, 2965.
32 R. Yamagami, K. Kobayashi and S. Tagawa, Chem.Eur. J.,
2009, 15, 12201.
33 K. Kawai, T. Kimura, K. Kawabata, S. Tojo and T. Majima,
J. Phys. Chem. B, 2003, 107, 12838.
34 C. Behrens, L. T. Burgdorf, A. Schwogler and T. Carell,
Angew. Chem., Int. Ed., 2002, 41, 1763.
35 C. Behrens, M. Ober and T. Carell, Eur. J. Org. Chem., 2002,
3281.
36 C. Behrens and T. Carell, Chem. Commun., 2003, 1632.
37 S. Breeger, U. Hennecke and T. Carell, J. Am. Chem. Soc.,
2004, 126, 1302.
38 M. K. Cichon, C. H. Haas, F. Grolle, A. Mees and T. Carell,
J. Am. Chem. Soc., 2002, 124, 13984.
39 C. Haas, K. Kraeling, M. Cichon, N. Rahe and T. Carell,
Angew. Chem., Int. Ed., 2004, 43, 1842.
40 A. Manetto, S. Breeger, C. Chatgilialoglu and T. Carell,
Angew. Chem., Int. Ed., 2006, 45, 318.
41 D. Fazio, C. Trindler, K. Heil, C. Chatgilialoglu and T. Carell,
Chem.Eur. J., 2011, 17, 206.
42 S. Breeger, M. von Meltzer, U. Hennecke and T. Carell,
Chem.Eur. J., 2006, 12, 6469.
43 T. Ito and S. E. Rokita, J. Am. Chem. Soc., 2003, 125, 11480.
44 T. Ito and S. E. Rokita, Angew. Chem., Int. Ed., 2004, 43, 1839.
45 F. D. Lewis, X. Liu, Y. Wu, S. E. Miller, M. R. Wasielewski,
R. L. Letsinger, R. Sanishvili, A. Joachimiak, V. Tereshko
and M. Egli, J. Am. Chem. Soc., 1999, 121, 9905.
46 F. D. Lewis, X. Liu, S. E. Miller, R. T. Hayes and
M. R. Wasielewski, J. Am. Chem. Soc., 2002, 124, 11280.
47 P. Daublain, A. K. Thazhathveetil, Q. Wang, A. Trifonov,
T. Fiebig and F. D. Lewis, J. Am. Chem. Soc., 2009, 131, 16790.
48 H.-A. Wagenknecht, Angew. Chem., Int. Ed., 2003, 42, 2454.
49 N. Amann, E. Pandurski, T. Fiebig and H.-A. Wagenknecht,
Chem.Eur. J., 2002, 8, 4877.
50 M. Raytchev, E. Mayer, N. Amann, H.-A. Wagenknecht and
T. Fiebig, ChemPhysChem, 2004, 5, 706.
51 C. Wagner and H.-A. Wagenknecht, Chem.Eur. J., 2005, 11,
1871.
52 S.-H. Lin, M. Fujitsuka, M. Ishikawa and T. Majima, J. Phys.
Chem. B, 2014, 118, 12186.
53 M. J. Park, M. Fujitsuka, K. Kawai and T. Majima, Chem.Eur.
J., 2012, 18, 2056.
54 M. Turbiez, P. Fr`
ere and J. Roncali, J. Org. Chem., 2003, 68,
5357.
55 K. Tainaka, M. Fujitsuka, T. Takada, K. Kawai and
T. Majima, J. Phys. Chem. B, 2010, 114, 14657.
56 J. N. Wilson, Y. Cho, S. Tan, A. Cuppoletti and E. T. Kool,
ChemBioChem, 2008, 9, 279.
57 M. J. Park, M. Fujitsuka, K. Kawai and T. Majima, J. Am.
Chem. Soc., 2011, 133, 15320.
58 T. Takada, K. Kawai, M. Fujitsuka and T. Majima, J. Am.
Chem. Soc., 2006, 128, 11012.
59 M. J. Park, M. Fujitsuka, H. Nishitera, K. Kawai and
T. Majima, Chem. Commun., 2012, 48, 11008.
60 K. Siegmund, P. Daublain, Q. Wang, A. Trifonov, T. Fiebig
and F. D. Lewis, J. Phys. Chem. B, 2009, 113, 16276.
61 S.-H. Lin, M. Fujitsuka and T. Majima, J. Phys. Chem. B, 2016,
120, 660.
62 G. B. Schuster, Acc. Chem. Res., 2000, 33, 253.
63 P. N. Borer, S. R. LaPlante, A. Kumar, N. Zanatta, A. Martin,
A. Hakkinen and G. C. Levy, Biochemistry, 1994, 33, 2441.
64 S.-H. Lin, M. Fujitsuka and T. Majima, Chem.Eur. J., 2015,
21, 16190.
65 S.-H. Lin, M. Fujitsuka and T. Majima, J. Phys. Chem. B, 2015,
119, 7994.
66 F.D.Lewis,P.Daublain,B.Cohen,J.Vura-Weis,V.Sharovich
andM.R.Wasielewski,J. Am. Chem. Soc., 2007, 129, 15130.
67 R. Huber, T. Fiebig and H.-A. Wagenknecht, Chem.
Commun., 2003, 1878.
68 S. Steenken, J. P. Telo, H. M. Novais and L. P. Candeias, J. Am.
Chem. Soc., 1992, 114, 4701.
69 S. Steenken, Chem. Rev., 1989, 89, 503.
70 R. Yamagami, K. Kobayashi and S. Tagawa, J. Am. Chem. Soc.,
2008, 130, 14772.
71 A. Szyperska, J. Rak, J. Leszczynski, X. Li, Y. J. Ko, H. Wang
and K. H. Bowen, ChemPhysChem, 2010, 11, 880.
72 X.Li,Z.CaiandM.D.Sevilla,J. Phys. Chem. B, 2001, 105, 10115.
73 N. A. Richardson, S. S. Wesolowski and H. F. Schaefer, J. Am.
Chem. Soc., 2002, 124, 10163.
74 A. Gupta, H. M. Jaeger, K. R. Compaan and H. F. Schaefer, J.
Phys. Chem. B, 2012, 116, 5579.
75 H.-Y. Chen, C.-L. Kao and S. C. N. Hsu, J. Am. Chem. Soc.,
2009, 131, 15930.
76 S. C. N. Hsu, T.-P. Wang, C.-L. Kao, H.-F. Chen, P.-Y. Yang
and H.-Y. Chen, J. Phys. Chem. B, 2013, 117, 2096.
77 S.-H. Lin, M. Fujitsuka and T. Majima, Angew. Chem., Int. Ed.,
2016, 55, 8715.
78 M. A. Grodick, N. B. Muren and J. K. Barton, Biochemistry,
2015, 54, 962.
79 A. R. Arnold, M. A. Grodick and J. K. Barton, Cell Chem. Biol.,
2016, 23, 183.
1762 |Chem. Sci.,2017,8,17521762 This journal is © The Royal Society of Chemistry 2017
Chemical Science Perspective
... Marguet et al. state that the interpretation of Ref. [12] was oversimplified [13]. Moreover, the charge-transfer dynamics in DNA have also been studied by time-resolved spectroscopy [14]. More interestingly, the development of multipulse configuration has been extended to UV regions and its application in 2D ES was achieved by Prokhorenko and coworkers [15] in 2016. ...
Article
Full-text available
Investigating exciton dynamics within DNA nucleobases is essential for comprehensively understanding how inherent photostability mechanisms function at the molecular level, particularly in the context of life’s resilience to solar radiation. In this paper, we introduce a mathematical model that effectively simulates the photoexcitation and deactivation dynamics of nucleobases within an ultrafast timeframe, particularly focusing on wave-packet dynamics under conditions of strong nonadiabatic coupling. Employing the hierarchy equation of motion, we simulate two-dimensional electronic spectra (2DES) and calibrate our model by comparing it with experimentally obtained spectra. This study also explores the effects of base stacking on the photo-deactivation dynamics in DNA. Our results demonstrate that, while strong excitonic interactions between nucleobases are present, they have a minimal impact on the deactivation dynamics of the wave packet in the electronic excited states. We further observe that the longevity of electronic excited states increases with additional base stacking and pairing, a phenomenon accurately depicted by our excitonic model. This model enables a detailed examination of the wave packet’s motion on electronic excited states and its rapid transition to the ground state. Additionally, using this model, we studied base stacks in DNA hairpins to effectively capture the primary exciton dynamics at a reasonable computational scale. Overall, this work provides a valuable framework for studying exciton dynamics from single nucleobases to complex structures such as DNA hairpins.
... Thymine-adenine-thymine (TAT) is a representative base triplet, which consists of one charge donor and two charge acceptors [11][12][13][14][15][16]. Adenine (A) is oxidized and can be used as a hole carrier, while thymine (T) is reduced and regarded as an electron carrier [17,18]. T and A form a dimer through the Watson-Crick structure, and the third base T is parallel to A and connected by hydrogen bonds to form a stable base triplet structure [19][20][21][22]. ...
Article
Full-text available
The short-range charge transfer of DNA base triplets has wide application prospects in bioelectronic devices for identifying DNA bases and clinical diagnostics, and the key to its development is to understand the mechanisms of short-range electron dynamics. However, tracing how electrons are transferred during the short-range charge transfer of DNA base triplets remains a great challenge. Here, by means of ab initio molecular dynamics and Ehrenfest dynamics, the nuclear–electron interaction in the thymine-adenine-thymine (TAT) charge transfer process is successfully simulated. The results show that the electron transfer of TAT has an oscillating phenomenon with a period of 10 fs. The charge density difference proves that the charge transfer proportion is as high as 59.817% at 50 fs. The peak position of the hydrogen bond fluctuates regularly between −0.040 and −0.056. The time-dependent Marcus–Levich–Jortner theory proves that the vibrational coupling between nucleus and electron induces coherent electron transfer in TAT. This work provides a real-time demonstration of the short-range coherent electron transfer of DNA base triplets and establishes a theoretical basis for the design and development of novel biological probe molecules.
... This discovery, made in 1993, opened the field of charge transport through nucleic acids. Since then, the conductivity of DNA has been demonstrated in numerous experiments (Fujitsuka and Majima 2017;Genereux and Barton 2010;Wagenknecht 2006). Beyond the aim of understanding the biological relevance, the main focus has been on the distances that can be covered in the process and the different types and mechanisms of charge transport. ...
... Although so far the high complexity of the system has partially hindered a complete characterization of the charge transfer processes in DNA, recent technical advances in spectroscopic instrumentation contributed to enhance our understanding of the kinetics of such phenomena. That is, starting from the first direct measurement of the photoinduced hole transport in DNA by time-resolved spectroscopy [3], several experimental works have contributed to shed light on this subject [4][5][6][7][8][9][10][11]. In particular, DNA hole transfer kinetics through π-stack arrays have been proved to depend not only from the redox potential of the single nucleobases but also on the DNA specific sequence and conformation [12] as revealed by means of time-resolved spectroscopic data. ...
Article
Full-text available
In this paper, we extend the previously described general model for charge transfer reactions, introducing specific changes to treat the hopping between energy minima of the electronic ground state (i.e., transitions between the corresponding vibrational ground states). We applied the theoretical–computational model to the charge transfer reactions in DNA molecules which still represent a challenge for a rational full understanding of their mechanism. Results show that the presented model can provide a valid, relatively simple, approach to quantitatively study such reactions shedding light on several important aspects of the reaction mechanism.
... Extensive investigation by researchers all over the world have discovered the amazing biochemical mechanisms, charge transfer, and semiconducting characteristics in DNA [5,6]. In the microscopic dimension compatible to a quantum well superlattice, the knowledge gained from semiconductor science can be applied and used to supplement to a better understanding of DNA [7][8][9] or vice versa. ...
Preprint
Full-text available
This study presents a quantum well model using transfer matrix to analyse the charge transfer characteristic in DNA nanostructure sequences. The unconfined state or unbound state above the quantum well are used to investigate the carrier behaviours in nanostructure of semiconductor. Similar analysis and simulations can be applied to the understanding of charge transfer in DNA nanostructure with sequence of periodic and quasi-periodic sequences. The model was validated with experiments using photoreflectance spectroscopy on nanostructure in semiconductor superlattice. In addition, the results published by the study of Thermoelectric effect and its dependence on molecular length and sequence in single DNA molecules[1] were used to compare with the simulations by the quantum well model. The results show agreement and can provide new insight into charge transfer and transport in DNA nanostructure with different types of sequences.
... This figure is adapted from Ref. 34. transport over nanometer scales because the charge transfer dynamics is limited by hopping among specific DNA nucleobases or nucleobase "islands" (typically purines for hole transport). 30,31 Achieving ratcheted charge transport in nucleic acids or other macromolecules may create novel opportunities to realize moleculebased circuitry in self-assembling matter. 32,33 In contrast to prior studies, which focused mainly on solid-state electronics, the ratchets explored here are based on a bottom-up approach with selfassembling soft matter. ...
Article
Ratcheted multi-step hopping electron transfer systems can plausibly produce directional charge transport over very large distances without requiring a source–drain voltage bias. We examine molecular strategies to realize ratcheted charge transport based on multi-step charge hopping, and we illustrate two ratcheting mechanisms with examples based on DNA structures. The charge transport times and currents that may be generated in these assemblies are also estimated using kinetic simulations. The first ratcheting mechanism described for nanoscale systems requires local electric fields on the 10⁹ V/m scale to realize nearly 100% population transport. The second ratcheting mechanism for even larger systems, based on electrochemical gating, is estimated to generate currents as large as 0.1 pA for DNA structures that are a few μm in length with a gate voltage of about 5 V, a magnitude comparable to currents measured in DNA wires at the nanoscale when a source–drain voltage bias of similar magnitude is applied, suggesting an approach to considerably extend the distance range over which DNA charge transport devices may operate.
Article
Full-text available
This is an Open Access Journal / article distributed under the terms of the Creative Commons Attribution License (CC BY-NC-ND 3.0) which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited. All rights reserved. High dilutions (HD) of drugs used in homeopathy are mostly devoid of original drug molecules. Water structure has been reported to carry the information of original drug molecules. We have defined the water structure in terms of free water molecules, hydrogen bond strength and number of hydrogen bonds. We have also reported that charge transfer interaction (CT) has been associated with HDs. In the present study we have analysed two drugs Cannabis sativa and Colchicum autumnale, and two potencies 6cH and 30cH by electronic and vibrational spectroscopy. The UV-Vis spectra of each potency show two peaks one at 200 nm and another 220 nm wave length. The first peak belongs to the absorbance by clathrate hydrate crystal (CHC), the second peak at higher wave length has been assigned to the CT interaction. In the CT interaction dissolved oxygen serve as electron acceptor and water or ethanol as electron donor. Dissolved oxygen might have been introduced in the solvent medium (EtOH-water) of the potencies during their preparation by mechanical agitation or succussion. CT interaction appears to ABSTRACT RESEARCH ARTICLE 48209 be a common factor for all homeopathic potencies. We have quantified free water molecules in the potencies from FTIR-spectra in the wave number region 3700 cm-1 to 3550 cm-1 .
Article
Full-text available
Surface‐enhanced Raman scattering (SERS) is among the most widely applied analytical techniques due to its easy execution and extreme sensitivity. Target molecules can be detected and distinguished based upon the fingerprint spectra that arise when absorbed on the SERS substrates surface, particularly on the SERS‐active hotspots. Thus, rational fabricating the enhancing substrates plays a key role in broadening SERS application. Programmable DNA functionalized plasmonic nanoassemblies, where DNA acts as both structure basis and functional unit, combine the specificity of DNA recognition, and modulate the assembly of plasmonic nanoparticles (NPs). Specifically designed DNA not only improves the selectivity to target molecules but also promotes the sensitivity of the optical signals through precisely regulating the distance between the molecule and the substrate. A variety of DNA‐functionalized SERS sensors have been reported and obtained well performance in the analysis of heavy metal ions in water, toxins, pesticide residues, antibiotics, hormones, illicit drugs, or other small molecules. This review places an emphasis on the design and sensing strategies of the DNA‐functionalized plasmonic nanoassemblies, as well as basic principles of Raman enhancement, and recent advances for environmental analysis. The current challenges and potential trends in the development of DNA‐functionalized SERS sensors for environmental pollutant monitoring in complicated scenarios are subsequently discussed. This review places an emphasis on the introduction of the programmable DNA‐functionalized plasmonic nanoassemblies as SERS sensors for environmental analysis, where the specifically designed DNA acts as both structure basis and functional unit, combining the specificity of DNA recognition and the sensitivity of SERS detection via modulating the assembly of plasmonic nanoparticles.
Article
Full-text available
Assessment of the DNA photo-oxidation and synthetic photocatalytic activity of chromium polypyridyl complexes is dominated by consideration of their long-lived metal-centered excited states. Here we report the participation of the excited states of [Cr(TMP)2dppz]³⁺ (1) (TMP = 3,4,7,8-tetramethyl-1,10-phenanthroline; dppz = dipyrido[3,2-a:2′,3′-c]phenazine) in DNA photoreactions. The interactions of enantiomers of 1 with natural DNA or with oligodeoxynucleotides with varying AT content (0–100%) have been studied by steady state UV/visible absorption and luminescence spectroscopic methods, and the emission of 1 is found to be quenched in all systems. The time-resolved infrared (TRIR) and visible absorption spectra (TA) of 1 following excitation in the region between 350 to 400 nm reveal the presence of relatively long-lived dppz-centered states which eventually yield the emissive metal-centered state. The dppz-localized states are fully quenched when bound by GC base pairs and partially so in the presence of an AT base-pair system to generate purine radical cations. The sensitized formation of the adenine radical cation species (A•+T) is identified by assigning the TRIR spectra with help of DFT calculations. In natural DNA and oligodeoxynucleotides containing a mixture of AT and GC of base pairs, the observed time-resolved spectra are consistent with eventual photo-oxidation occurring predominantly at guanine through hole migration between base pairs. The combined targeting of purines leads to enhanced photo-oxidation of guanine. These results show that DNA photo-oxidation by the intercalated 1, which locates the dppz in contact with the target purines, is dominated by the LC centered excited state. This work has implications for future phototherapeutics and photocatalysis.
Article
Remote reductive repair of thymine dimers in a DNA duplex by transfer of excess electrons over a distance of up to roughly 24 Å (n=7) has been attributed to thermally activated hopping (see scheme). Possible consequences for humans: the harmful effect of UV irradiation responsible for the development of skin cancer could potentially be reduced by compounds that bind to DNA and trigger long-range electron transport.
Article
Remote reductive repair of thymine dimers in a DNA duplex by transfer of excess electrons over a distance of up to roughly 24 Å (n=7) has been attributed to thermally activated hopping (see scheme). Possible consequences for humans: the harmful effect of UV irradiation responsible for the development of skin cancer could potentially be reduced by compounds that bind to DNA and trigger long-range electron transport.
Article
Given its well-ordered continuous π stacking of nucleobases, DNA has been considered as a biomaterial for charge transfer in biosensors. For cathodic photocurrent generation resulting from hole transfer in DNA, sensitivity to DNA structure and base-pair stacking has been confirmed. However, such information has not been provided for anodic photocurrent generation resulting from excess-electron transfer in DNA. In the present study, we measured the anodic photocurrent of a DNA-modified Au electrode. Our results demonstrate long-distance excess-electron transfer in DNA, which is dominated by a hopping mechanism, and the photocurrent generation is sequence dependent.
Article
The DNA double helix has captured the imagination of many, bringing it to the forefront of biological research. DNA has unique features that extend our interest into areas of chemistry, physics, material science, and engineering. Our laboratory has focused on studies of DNA charge transport (CT), wherein charges can efficiently travel long molecular distances through the DNA helix while maintaining an exquisite sensitivity to base pair π-stacking. Because DNA CT chemistry reports on the integrity of the DNA duplex, this property may be exploited to develop electrochemical devices to detect DNA lesions and DNA-binding proteins. Furthermore, studies now indicate that DNA CT may also be used in the cell by, for example, DNA repair proteins, as a cellular diagnostic, in order to scan the genome to localize efficiently to damage sites. In this review, we describe this evolution of DNA CT chemistry from the discovery of fundamental chemical principles to applications in diagnostic strategies and possible roles in biology.
Article
The dynamics of excess-electron transfer in DNA have attracted the attention of scientists from all kinds of research fields due to their importance in biological processes. To date, several studies on excess-electron transfer in consecutive adenine (A) :thymine (T) sequences in donor-DNA-acceptor systems have been published. However, the reported excess-electron transfer rate constants for consecutive Ts are in the range of 10(10)-10(11) s(-1) depending on photosensitizing electron donors, which provided various driving forces for excess-electron injection to DNA. In this study, we employed a strongly electron-donating photosensitizer, a dimer of 3,4-ethylenedioxythiophene (2E), and an electron acceptor, diphenylacetylene (DPA), to synthesize a series of modified DNA oligomers (2-Tn) to investigate excess-electron transfer dynamics in donor-DNA-acceptor by using femtosecond laser flash photolysis. The relation between free energy change for charge injection and excess-electron transfer rate among consecutive Ts provided the intrinsic excess-electron hopping rate constant ((3.8 ± 1.5) x 10(10) s(-1)) in DNA, which is consistent with the fluctuation frequency of the DNA sugar backbone and bases (3.3 x 10(10) s(-1)). Thus, we discussed the structural fluctuation effect on the excess-electron hopping in DNA.
Article
This paper presents the results of an investigation into the sequence-dependent excess-electron transfer (EET) dynamics in DNA, which plays an important role in DNA damage/repair. There are many published studies on EET in consecutive adenine:thymine (A:T) sequences (Tn), but those in alternating A:T sequences (ATn) remain limited. Here, two series of functionalized DNA oligomers, Tn and ATn, were synthesized with a strongly electron-donating photosensitizer, a trimer of ethylenedioxythiophene (3 E), and an electron acceptor, diphenylacetylene (DPA). Laser flash photolysis experiments showed that the EET rate constant of AT3 is two times lower than that of T3 due to the lack of π-stacking of Ts in AT3. Thus, it was indicated that excess-electron hopping is affected by the interaction between LUMOs of nucleotides.
Article
Charge transfer and proton transfer in DNA have attracted wide attention due to their relevance in biological processes and so on. Especially, excess-electron transfer (EET) in DNA has strong relation to DNA repair. However, our understanding on EET in DNA still remains limited. Herein, by using a strongly electron-donating photosensitizer, trimer of 3,4-ethylenedioxythiophene (3E), and an electron acceptor, diphenylacetylene (DPA), two series of functionalized DNA oligomers were synthesized for investigation of EET dynamics in DNA. The transient absorption measurements during femtosecond laser flash photolysis showed that guanine:cytosine (G:C) base pair affects the EET dynamics in DNA by two possible mechanisms: the excess-electron quenching by proton transfer with the complementary G after formation of C(●-) and the EET hindrance by inserting a G:C base pair as a potential barrier in consecutive thymines (Ts). In the present paper, we provided useful information based on the direct kinetic measurements, which allowed us to discuss EET through oligonucleotides for the investigation of DNA damage/repair.
Article
Intermolecular static and dynamic fluorescence quenching constants of eight coumarin derivatives by nucleobase derivatives have been determined in aqueous media. One common sequence of the quenching efficiency has been found for the nucleobases. The feasibility of a photoinduced electron transfer reaction for the nucleobase-specific quenching of fluorescent dyes is investigated by the calculation of the standard free energy changes with the Rehm-Weller equation. A complete set of one-electron redox potential data for the nucleobases are determined electrochemically in aprotic solvents for the first time, which are compared with values obtained by various other methods. Depending on the redox properties of the fluorescent dyes, the sequences of the quenching efficiencies can be rationalized by the orders of electrochemical oxidation potentials (vs NHE) of nucleosides (dG (+1.47 V) < dA ( dC approximate to dT < U (greater than or equal to +2.39 V)) and reduction potentials (de (( -2.76 V) < dA < dC approximate to dT < U (-2.07 V)). The correlation between the intermolecular dynamic quenching constants and the standard free energy of photoinduced electron transfer according to the classical Marcus equation indicates that photoinduced electron transfer is the rate-limiting step. However, an additional, water-specific gain of free energy between -0.5 and -0.9 eV shows that additional effects, like a coupled proton transfer and a hydrophobic effect, have to be considered, too. Furthermore, the capability of the nucleobases to form ground state complexes with fluorescent dyes is influenced by their redox potentials. The relevance of these observations to current efforts for DNA sequencing with a detection by laser-induced fluorescence and their application to other dyes are discussed.
Article
The unique characteristics of DNA charge transport (CT) have prompted an examination of roles for this chemistry within a biological context. Not only can DNA CT facilitate long-range oxidative damage of DNA, but redox-active proteins can couple to the DNA base stack and participate in long-range redox reactions using DNA CT. DNA transcription factors with redox-active moieties such as SoxR and p53 can use DNA CT as a form of redox sensing. DNA CT chemistry also provides a means to monitor the integrity of the DNA, given the sensitivity of DNA CT to perturbations in base stacking as arise with mismatches and lesions. Enzymes that utilize this chemistry include an interesting and ever-growing class of DNA-processing enzymes involved in DNA repair, replication, and transcription that have been found to contain 4Fe-4S clusters. DNA repair enzymes containing 4Fe-4S clusters, that include endonuclease III (EndoIII), MutY, and DinG from bacteria, as well as XPD from archaea, have been shown to be redox-active when bound to DNA, share a DNA-bound redox potential, and can be reduced and oxidized at long-range via DNA CT. Interactions between DNA and these proteins in solution, in addition to genetics experiments within Escherichia coli, suggest that DNA-mediated CT can be used as a means of cooperative signaling among DNA repair proteins that contain 4Fe-4S clusters as a first step in finding DNA damage, even within cells. On the basis of these data, we can consider also how DNA-mediated CT may be used as a means of signaling to coordinate DNA processing across the genome.
Article
Charge transfer in DNA has attracted great attention of scientists because of its importance in biological processes. However, our knowledge on excess-electron transfer in DNA still remains limited when compared to numerous studies of hole transfer in DNA. To clarify the dynamic of excess-electron transfer in DNA by photochemical techniques, new electron-donating photosensitizers should be developed. Herein, a terthiophene and two 3,4-ethylenedioxythiophene oligomers were used as photosensitizers in dyads including natural nucleobases as electron acceptors. The charge separation and recombination processes in the dyads were investigated by femtosecond laser flash photolysis, and the driving force dependence of these rate constants was discussed on the basis of the Marcus theory. From this study, the conformation effect on charge recombination process was found. We expect that 3,4-ethylenedioxythiophene oligomers are useful in investigation of excess-electron transfer dynamics in DNA.