ArticlePDF Available

The photochemistry of 8-bromo-2-deoxyadenosine. A direct entry to cyclopurine lesions

Authors:

Abstract and Figures

The UV photolysis of 8-bromo-2′-deoxyadenosine has been investigated in different solvents and in the presence of additives like halide anions. Photolytic cleavage of the C–Br bond leads to formation of the C8 radical. In methanol, subsequent hydrogen abstraction from the solvent is the main radical reaction; however, in water or acetonitrile intramolecular hydrogen abstraction from the sugar moiety, to give the C5′ radical, is the major path. This C5′ radical undergoes a cyclization reaction on the adenine and gives the aminyl radical. A rate constant of 1.8 × 105 s−1 has been measured by laser flash photolysis in CH3CN for this unimolecular process. Product studies from steady-state photolysis in acetonitrile have shown the conversion of 8-bromo-2′-deoxyadenosine to 5′,8-cyclo-2′-deoxyadenosine in 65% yield and in a diastereoisomeric ratio (5′R) ∶ (5′S) = 1.7. Evidence supporting that the equilibrium Br˙ + Br− ⇌ Br2˙− plays an important role in this synthetically useful radical cascade is obtained by regulating the relative concentrations of the two reactive oxidizing species.
Absorption spectrum obtained from the laser flash photolysis of an Ar-purged CH 3 CN solution containing 1 mM of 1, taken 19 ls after the pulse. Inset: Time dependence of absorption at 360 nm; the solid line represents the first-order kinetic fit to the data. solvent, or CH 3 CN saturated by air, the absorption at 360 nm decreased substantially indicating that the precursor radical(s) reacted efficiently with methanol and molecular oxygen. The results described above demonstrate that the C–Br bond in bromide 1 is efficiently cleaved by UV light producing Br @BULLET and the neutral r-type radical 2 (Scheme 2). In methanol as the solvent, intermolecular hydrogen abstraction by radical 2 is expected to be favored, as it is exothermic of ca. 20 kcal mol −1 , 9 to afford mainly the reduction product 7 (Table 1, entry 4). No evidence for formation or disappearance of radical 2 was obtained by laser flash photolysis experiments; however, it is reported that the rate constant for the reaction of phenyl radical with CH 3 OH is 4.4 × 10 6 M −1 s −1 . 10 In acetonitrile, hydrogen abstraction mainly occurs intramolecularly to give radical 3. This radical is calculated at the B3LYP/6-31G* level to be nearly planar p-type with a very low interconversion barrier, and its cyclization should afford the two aminyl radicals 12 and 13 in chair conformations. 2 The fate of radicals 12 and 13 mainly depends on the redox properties of the reaction partner. We suggest that these radicals are readily oxidized in the reaction mixture by transient oxidants (see below) followed by a rapid deprotonation to afford the corresponding compounds 4 and 5. In water, the above mentioned products are accompanied by large amounts of hydrated 5 -carboxyaldehyde 6, which should be due to the oxidation of radical 3 with formation of oxocarbenium 11 followed by reaction with the medium. Then, the question arises as to which is the oxidant acting during the course of the reaction. The simplest answer could be Br @BULLET atoms that are directly obtained from the photolysis of 1. The redox properties of Br @BULLET are well known with E @BULLET (Br @BULLET /Br − ) = 1.92 V. 11 Hence, the oxidation of C5 radical 3 by Br @BULLET is thermodynamically quite favorable. Based on CH 3 CH( @BULLET )OH radical for which E @BULLET (CH 3 CHO, H + /CH 3 CH( @BULLET )OH) = −1.25 V,
… 
Content may be subject to copyright.
FULL PAPER
PPS
www.rsc.org/pps
The photochemistry of 8-bromo-2-deoxyadenosine. A direct entry to
cyclopurine lesions
Liliana B. Jimenez,a,bSusana Encinas,aMiguel A. Miranda,*aMaria Luisa Navacchiaband
Chryssostomos Chatgilialoglu*b
aDepartamento de Qu´
ımica/Instituto de Tecnolog´
ıa Qu´
ımica UPV-CSIC, Universidad
Polit´
ecnica de Valencia, Camino de Vera s/n, E-46022, Valencia, Spain.
E-mail: mmiranda@qim.upv.es; Fax: 34 96 3877809; Tel: 34 96 3877804
bISOF, Consiglio Nazionale delle Ricerche, Via P. Gobetti 101, I-40129, Bologna, Italy.
E-mail: chrys@isof.cnr.it; Fax: 39 051 6398349; Tel: 39 051 6398309
Received 19th July 2004, Accepted 28th September 2004
First published as an Advance Article on the web 22nd October 2004
The UV photolysis of 8-bromo-2-deoxyadenosine has been investigated in different solvents and in the presence of
additives like halide anions. Photolytic cleavage of the C–Br bond leads to formation of the C8 radical. In methanol,
subsequent hydrogen abstraction from the solvent is the main radical reaction; however, in water or acetonitrile
intramolecular hydrogen abstraction from the sugar moiety, to give the C5radical, is the major path. This C5radical
undergoes a cyclization reaction on the adenine and gives the aminyl radical. A rate constant of 1.8 ×105s1has
been measured by laser flash photolysis in CH3CN for this unimolecular process. Product studies from steady-state
photolysis in acetonitrile have shown the conversion of 8-bromo-2-deoxyadenosine to 5,8-cyclo-2-deoxyadenosine
in 65% yield and in a diastereoisomeric ratio (5R):(5
S)=1.7. Evidence supporting that the equilibrium Br+
BrBr2
plays an important role in this synthetically useful radical cascade is obtained by regulating the relative
concentrations of the two reactive oxidizing species.
Introduction
Cyclopurine lesions, in particular 5,8-cyclo-2-deoxyadenosine
(5,8-cyclodAdo), are observed among the decomposition prod-
ucts of DNA when exposed to ionizing radiations or when
treated chemically by one-electron reductants.1Apart from the
usual glycosidic bond, in these moieties there is an additional
base-sugar linkage between the C8 and C5positions. From a
mechanistic point of view, it has been verified that the C5radical
initially generated by hydrogen abstraction, intramolecularly
attacks the aromatic ring of the base moiety to form cyclopurines
as the final products after oxidation.2
The chemical synthesis of the two diastereoisomeric forms of
5,8-cyclodAdo, i.e. (5R)- and (5S)-isomers, as well as their
incorporation on specific sites of DNA, is of considerable
importance in order to investigate the biological impact of
cyclopurines on the conformation and function of the double
helix. The existing approaches are mainly limited to low yield
multiple-step syntheses. The (5S)-isomer of 5,8-cyclodAdo
was prepared in seven steps starting from N6-benzoyl-dAdo in
an overall yield <10%.3Preparation of the (5R)-isomer was
further achieved by two additional steps, involving inversion of
configuration at the C5position.4Synthetic oligonucleotides
that contain these modified derivatives at selected sites were also
prepared,3,4 and some biochemical and biophysical features of
such lesions have already been obtained.5,6
Some of us recently investigated the reaction of hydrated elec-
trons (eaq
) with 8-bromo-2-deoxyadenosine (1) by radiolytic
methods.2It was found that 1is prompt to capture electrons and
rapidly loses the bromide ion, to give the corresponding radical
in the C8 position (2), which in turn abstracts intramolecularly
a hydrogen atom exclusively from the C5position to selectively
afford the desired C5radical 3(Scheme 1). This allowed for the
first time the study the fate of the 2-deoxyadenosin-5-yl radical
properly and in particular the cyclization step, which occurs
with a rate constant kc=1.6 ×105s1. Both species 3and
the cyclized aminyl radical are readily oxidized by Fe(CN)63,
the rate constants being 4.2 ×109and 8.3 ×108M1s1,
respectively, whereas the aminyl radical can also be reduced by
strong reductants.2
In the present paper, we describe a photochemical study of 1
under various conditions.7It will be shown that a synthetically
useful radical cascade process has been developed allowing for
a one-pot conversion of 1to 5,8-cyclodAdo in a very good
yield and in a diastereoisomeric ratio strongly dependent on the
experimental conditions.
Experimental
Chemicals
8-Bromo-2-deoxyadenosine was provided by Berry & Asso-
ciates. Sodium iodide, phosphate buffer, 1,4-diazabicyclo[2.2.2]
octane (DABCO) and tetrabutylammonium iodide were pur-
chased from Aldrich. Acetonitrile and methanol, both HPLC
grade, were from Scharlau. Water was purified through a
Millipore Milli-RO plus 30 system.
Instrumentation
Reverse-phase HPLC analysis was performed on a Waters
apparatus equipped with a Teknokroma column (Kromasil
N30396, C18, 5 lm packing), a Waters 2996 photodiode
array detector at fixed wavelength of 254 nm and a Waters
600 controller. Chromatographic system used for analytical
experiments consisted of acetonitrile and water as eluents [linear
gradient from 0 to 5% of acetonitrile (30 min), from 5 to 10% of
acetonitrile (40 min) and from 10 to 30% of acetonitrile (60 min)]
ataflowrateof0.7mLmin
1.
Steady-state photolysis
Solutions (10 mL) containing bromide 1(ca. 1mM)were
irradiated under N2. The irradiation sources were (A) a 125 W
medium-pressure mercury lamp and (B) a Rayonet multilamp
photoreactor (254 nm lamps). Different solvents were used
DOI:10.1039/b410939b
1042
Photochem. Photobiol. Sci.
, 2004, 3, 1042–1046
This journal is
©
The Royal Society of Chemistry and Owner Societies 2004
Downloaded by CNR Bologna on 09 May 2012
Published on 22 October 2004 on http://pubs.rsc.org | doi:10.1039/B410939B
View Online
/ Journal Homepage
/ Table of Contents for this issue
Scheme 1
(as indicated in Tables 1 and 2). The two diastereoisomers
of 5,8-cyclodAdo were purified on reverse-phase silica gel
(Fluka) eluting with 0, 1, 2 and 3% acetonitrile-containing
water. The UV-positive fractions were collected and lyophilized
to obtain the desired products as pure materials that were
spectroscopically characterized.2
Nanosecond laser flash photolysis
Laser flash photolysis experiments were performed by using
a Q-switched Nd:YAG laser (Quantel Brilliant, at 266 nm,
4 mJ pulse1, 5 ns fwhm) coupled to a mLFP-111 Luzchem
miniaturized equipment. All transient spectra were recorded
employing 10 ×10 mm quartz cells with 4 mL capacity that were
bubbled during 30 min with N2before acquisition. Photolysis of
1was performed in CH3CNandalsoinCH
3CN/MeOH (v/v
1 : 3). In the case of solutions in CH3CN data were also recorded
in the presence of air or O2. The absorbance of the samples
was kept between 0.20 and 0.60 at the laser wavelength. All the
experiments were carried out at room temperature (22 C).
Results and discussion
Bromide 1was UV-irradiated under a variety of experimental
conditions. The crude reaction mixtures were analyzed by
HPLC coupled with UV (diode-array) and MS (ion trap)
detection, using authentic samples as reference compounds for
the identification of the products. Table 1 and Chart 1 summarize
the experimental findings and the reaction products for which
yields are based on the consumption of starting bromide for a
better comparison.
Chart 1
Table 1 Yields (%) of products 49(see Chart 1) obtained by UV photolysis of 1a
Entry Solvent
Photolysis
conditions Conversion (%)
Overall yield
(%) 456789
1H
2O 1.5 hb61 93 21 6 17 10 34 5
2 1.5 hc95 100 31 7 34 19 9
3 1.5 hd48 88 14 6 68
4CH
3OH 1.5 h 97 82 5 2 2 62 11
5CH
3CN 0.5 h 100 96 41 24 17 14
6 1 h 100 90 36 23 18 13
7 2 h 100 86 29 26 14 17
aIrradiation of 1(10 mL tube1, 1 mM) under N2was performed with a multilamp photoreactor at 254 nm; in this system temperature changed from
20 to 50 C during the reaction course. Yields are based on the conversion of bromide 1. Estimated errors are less than 10% of the stated values. bpH
Changed from 7 to 4 during the reaction course. cpH 7; buffer solution with 10 mM KH2PO4/Na2HPO4.dIn the presence of DABCO (10 equiv);
pH changed from 9 to 8 during the reaction course.
Photochem. Photobiol. Sci.
, 2004, 3, 1042–1046 1043
Downloaded by CNR Bologna on 09 May 2012
Published on 22 October 2004 on http://pubs.rsc.org | doi:10.1039/B410939B
View Online
Table 2 Yields (%) of products 410 (see Chart 1) obtained by UV photolysis of 1in the presence of an additivea
Entry Solvent/additive
Photolysis
conditionsbConversion (%)
Overall
yield (%) 4567910
1H
2O/NaBr 3 h, A 58 98 34 13 27 13 7 4
23h,
cA 88 100 16 13 34 10 8 19
3H
2O/NaI 1.5 h, B 100 96 35 11 33 3 14 —
4 3 h, A 100 93 12 9 17 3 32 20
5 1.5 h, B +4h, A 100 100 3 9 24 5 36 23
67h,
cA 100 92 — — 33 3 11 45
7CH
3CN/Bu4N+I2 min, B 33 100 5 4 28 13 50
8 10 min, B 100 100 3 5 6 16 70
9 30 min, B 100 100 1 2 2 18 77
aYields are based on the conversion of bromide 1. Estimated errors are less than 10% of the stated values. bA: Irradiation of 1(10 mL tube1,1mM)
under N2was performed with a 125 W medium-pressure mercury lamp; in this system temperature was maintained at 20 C during the reaction
course. B: Irradiation of 1(10 mL tube1, 1 mM) was performed with a multilamp photoreactor at 254 nm; in this system temperature changed from
20 to 50 C during the reaction course. cBuffer solution with 10 mM KH2PO4/Na2HPO4.
Entry 1 in Table 1 shows the reaction in aqueous medium
and at natural pH. After 1.5 h of photolysis 61% of the starting
material was converted to products. The cyclic products (4and
5) were formed in 27% overall yield and in a ratio (5R):(5
S)=
3.5, together with 17% of hydrated 5-carboxyaldehyde (6)and
10% of reduced product (7). However, the reaction unexpectedly
also led to the formation of 8-bromoadenine (8)andadenine(9)
in 34 and 5% yields, respectively, which are formal products of
glycolysis of the corresponding nucleosides. Indeed, during the
reaction the pH changed from 7 to 4. In order to verify this
hypothesis, the reaction was carried out in a buffered solution
at pH 7 (KH2PO4/Na2HPO410 mM). Under these conditions
(entry 2) a higher conversion was observed, with the suppression
of 8-bromoadenine (8) and the increase of the other five products
including adenine (9), whereas the (5R):(5
S) ratio was slightly
increased.
We also considered that amines may act as electron donors
under our photolytic conditions. Entry 3 shows the reaction
in aqueous medium in the presence of 10 equiv. of DABCO
(10 mM). The efficiency of this reaction was relatively low
with complete suppression of cyclic nucleosides and formation
of large amounts of adenine. During the reaction course the
pH changed from 9 to 8. Similar results were obtained by
replacing DABCO by Et3N (data not shown). Entry 4 shows
the reaction in methanol where 97% consumption of starting
bromide was observed after 1.5 h of irradiation. The increase
of 2-deoxyadenosine (7) to the detriment of cyclic products and
hydrat ed 5-carboxyaldehyde is noteworthy. Entry 5 shows the
reaction in acetonitrile where complete consumption of starting
bromide was observed in a shorter time. Under these conditions,
the formation of cyclic products reached a 65% overall yield
with the suppression of hydrated aldehyde 6.Itisalsoworth
mentioning that the ratio (5R):(5
S)=1.7 in CH3CN is halved
compared to water. By increasing the irradiation time the overall
yield decreased (entries 6 and 7) due to the consumption of (5R)-
isomer (4), whereas the other three products are quite stable.
It is worth underlining the different photo-stability of the two
diastereoisomers (see below).
In order to obtain information about the reactive intermedi-
ates involved in these synthetically useful reactions, the photo-
reactivity of 1was investigated by using laser flash photolysis
techniques. Photolysis of the bromo derivative 1in CH3CN with
266 nm laser light, under an anaerobic atmosphere, did not result
in the ‘instantaneous’ formation of a transient. However, the
spectrum shown in Fig. 1 developed in 20 ls after the pulse. The
time profile for the transient formation with kmax =360 nm (inset)
followed first-order kinetics with a rate constant k=1.8 ×105s1
at room temperature. In analogy to the reaction of hydrated
electrons with 1, we assigned this transient to the conjugated
aminyl radical, and the observed rate to the cyclization of C5
radical 3(Scheme 1).3,8 Using CH3CN/CH3OH (v/v 1 : 3) as the
Fig. 1 Absorption spectrum obtained from the laser flash photolysis of
an Ar-purged CH3CN solution containing 1 mM of 1, taken 19 ls after
the pulse. Inset: Time dependence of absorption at 360 nm; the solid line
represents the first-order kinetic fit to the data.
solvent, or CH3CN saturated by air, the absorption at 360 nm
decreased substantially indicating that the precursor radical(s)
reacted efficiently with methanol and molecular oxygen.
The results described above demonstrate that the C–Br bond
in bromide 1is efficiently cleaved by UV light producing Br
and the neutral r-type radical 2(Scheme 2). In methanol as
the solvent, intermolecular hydrogen abstraction by radical 2is
expected to be favored, as it is exothermic of ca. 20 kcal mol1,
9to afford mainly the reduction product 7(Table 1, entry 4).
No evidence for formation or disappearance of radical 2was
obtained by laser flash photolysis experiments; however, it is
reported that the rate constant for the reaction of phenyl radical
with CH3OH is 4.4 ×106M1s1.10 In acetonitrile, hydrogen
abstraction mainly occurs intramolecularly to give radical 3.
This radical is calculated at the B3LYP/6-31G* level to be nearly
planar p-type with a very low interconversion barrier, and its
cyclization should afford the two aminyl radicals 12 and 13 in
chair conformations.2The fate of radicals 12 and 13 mainly
depends on the redox properties of the reaction partner. We
suggest that these radicals are readily oxidized in the reaction
mixture by transient oxidants (see below) followed by a rapid
deprotonation to afford the corresponding compounds 4and
5. In water, the above mentioned products are accompanied
by large amounts of hydrated 5-carboxyaldehyde 6,which
should be due to the oxidation of radical 3with formation of
oxocarbenium 11 followed by reaction with the medium.
Then, the question arises as to which is the oxidant acting
during the course of the reaction. The simplest answer could be
Bratoms that are directly obtained from the photolysis of 1.
The redox properties of Brare well known with E(Br/Br)=
1.92 V.11 Hence, the oxidation of C5radical 3by Bris
thermodynamically quite favorable. Based on CH3CH()OH
radical for which E(CH3CHO, H+/CH3CH()OH) =−1.25 V,
1044
Photochem. Photobiol. Sci.
, 2004, 3, 1042–1046
Downloaded by CNR Bologna on 09 May 2012
Published on 22 October 2004 on http://pubs.rsc.org | doi:10.1039/B410939B
View Online
Scheme 2
the driving force for reaction of 3with BrE(Br/Br)–
E(CH3CHO, H+/CH3CH()OH) =3.17 V is quite high and
should bring to a fast process. Similar consideration can be made
for aminyl radicals 12 or 13. Since the E(N+/N)0.36 V
(Nrefers to radical 12 or 13 and N+to the corresponding
cations),2thedrivingforceforthereactionof12 or 13 with Br
is 1.56 V. Under our experimental conditions, lMlevelofBr
should be reached very quickly as the reduction product of Br
and the equilibrium given in reaction (1) should start to play a
role. Rate and equilibrium constants of this reaction as well as
its temperature-dependence are well known. At 20 Cthevalues
are k1=1.1 ×1010 M1s1,k1=2.7 ×104s1and K1=3.6 ×
105M1whereas at 50 C the values are k1=1.9 ×1010 M1s1,
k1=9.2 ×104s1and K1=2.0 ×105M1.12 The equilibrium
defined by reaction (1) is important because it regulates the
relative concentrations of solvated reactive species. Basically this
means a conversion of the highly oxidizing Br[E(Br/Br)=
1.92 V] to a less oxidizing Br2
[E(Br2
/2Br)=1.66 V].11
Still, Br2
could play the role of an effective oxidant.
Br·+Brk1
k1
Br·
2(1)
In order to obtain further information about the reaction
mechanism and, in particular, to test the ability of halide anions
(Bror I)totrapBr
atoms in the reaction mixture, the reaction
in water was carried out in the presence of NaBr or NaI. Table 2
summarizes the experimental findings. Again yields are based on
the consumption of starting bromide for a better comparison.
Entry 1 in Table 2 shows the results of photolysis of de-aerated
aqueous solution containing 1 mM of 1in the presence of 2
equiv NaBr. The results were similar to the outcome in the
absence of NaBr (entry 1 in Table 1), although the suppression
of 8-bromoadenine (8) and the appearance of small amounts of
a new product, 2,5-dideoxycycloadenosine (10), can be noticed.
However, the same reaction carried out in buffer (pH 7) showed
an increase of this new product to the detriment of (5R)-isomer
4(entry 2). Entry 3 in Table 2 shows the reaction in the presence
of 2 equiv of NaI. Under prolonged photolysis the appearance of
10 mainly to the detriment of (5R)-isomer is again noteworthy
(entries 4 and 5). When the same solution was buffered (entry
6) the (5R)- and (5S)-isomers completely disappeared and the
dideoxy derivative 10 was obtained in a 45% yield. The effect of
iodide in the photolysis of 1in CH3CNwasalsotestedusingthe
soluble source Bu4N+I(entries 7 and 8). Under these conditions
the cyclonucleosides were formed in small quantities indicating
that C5radical is rapidly trapped prior to cyclization to give
5-carboxaldehyde (6), which further decomposes under the UV
irradition to afford the free adenine base in high yield. Thus,
adenine seems to be produced by photoinduced glycolysis of the
aldehyde.
In the presence of 2 mM NaBr, reaction (1) became the
predominating path of bromine atom since the equilibrium is
completely shifted to the right (K1=3.6 ×105M1). Rate and
equilibrium constants of reaction (2) as well as its temperature-
dependence are also well known.13 At 20 C the values are k2=
8.9 ×109M1s1,k2=6.5 ×104s1and K2=1.4 ×105M1,
which are similar to the corresponding data for reaction (1). It is
reasonable to assume similar data for the reaction of Bratoms
with I. Therefore, in the presence of 2 mM of halide anions,
Br2
or IBrshould play only the role of effective oxidants.
Moreover, since I2
[E(I2
/2I)=1.05 V] is weaker oxidant
than Br2
[E(Br2
/2Br)=1.66 V] it is reasonable to assume
that IBrhas an intermediate oxidizing ability, i.e.,E(IBr/I,
Br)1.36 V.11
I·+Ik2
k2
I·
2(2)
The formation of 2,5-dideoxycycloadenosine (10) is not
straightforward. This compound is clearly formed from the
two 5,8-cyclo-2-deoxyadenosines, (5R)-isomer being the more
reactive. It is favored in aqueous solution, in the presence of
halide anions and with increasing irradiation times. Evidence
supporting that higher temperature and phosphate catalysis
facilitate this process has also been obtained. Although a
detailed mechanistic investigation is beyond the scope of this
paper, a reasonable proposal is a reductive elimination or
photo-induced decomposition. Indeed, Br2
[E(Br2/Br2
)=
0.50 V] and I2
[E(I2/I2
)=0.21 V],11 are known to act as
reductants depending on the reaction partner. Current studies
are in progress to address these issues as well as to investigate
the importance of the reactions in DNA damage and repair.
Conclusions
Selective generation of C5radical in 2-deoxyadenosine has
been achieved by UV photolysis of C–Br in 8-bromo-2-
deoxyadenosine (1) followed by fast radical translocation. The
reactivity of C5radical has been studied in some details in
different solvents and in the presence of additives like halide
anions. An expedient one-pot procedure has also been developed
Photochem. Photobiol. Sci.
, 2004, 3, 1042–1046 1045
Downloaded by CNR Bologna on 09 May 2012
Published on 22 October 2004 on http://pubs.rsc.org | doi:10.1039/B410939B
View Online
that allows the conversion of 8-bromo-2-deoxyadenosine (1)to
5,8-cyclo-2-deoxyadenosine (5,8-cyclodAdo) in 65% and in a
diastereoisomeric ratio (5R):(5
S)=1.7 by UV photolysis in
acetonitrile. This radical cascade consists of photolytic cleavage
of the C–Br bond to give the C8 radical, a radical translocation
from C8 to C5position, a cyclization of C5radicaltothe
adenine moiety with a rate constant of 1.8 ×105s1to give an
aminyl radical and its final oxidation. The evidence supports
the suggestion that the equilibrium Br+BrBr2
plays an
important role by regulating the relative concentrations of the
two reactive oxidizing species has been obtained.
Acknowledgements
We thank Clara Caminal for providing us with a pure sample of
5-carboxyaldehyde-2-deoxyadenosine, which has been used as
reference in this work. Work supported in part by the European
Community’s Marie Curie Research Training Network under
contract MRTN-CT-2003-505086 [CLUSTOXDNA]. We also
thank the financial support given by the Spanish MCYT (BQU
2001-2725 and Ram´
on y Cajal project to S.E.), the Generalitat
Valenciana (Grupos 03/082, CTBPRB/2003/68 and Project
GV04A-349) and the UPV (Project PPI-06-03).
References
1(a) For example, see: M.-L. Dirksen, W. F. Brakely, E. Hol-
witt and M. Dizdaroglu, Effect of DNA conformation on
the hydroxyl radical-induced formation of 5,8-cyclopurine-2-
deoxyribonucleoside residues in DNA, Int. J. Radiat. Biol., 1988,
54, 195–204; (b) M. Dizdaroglu, P. Jaruga and H. Rodriguez,
Identification and quantification of 5,8-cyclo-2-deoxyadenosine in
DNA by liquid chromatography/mass spectrometry, Free Radical
Biol. Med., 2001, 30, 774–784; (c) P. Jaruga, M. Birincioglu, H.
Rodriguez and M. Dizdaroglu, Mass spectrometric assays for the
tandem lesion 5,8-cyclo-2-deoxyguanosine in mammalian DNA,
Biochemistry, 2002, 41, 3703–3711; (d) M. Birincioglu, P. Jaruga, G.
Chowdhury, H. Rodriguez, M. Dizdaroglu and K. S. Gates, DNA
base damage by the antitumor agent 3-amino-1,2,4-benzotriazine
1,4-dioxide (tirapazamine), J. Am. Chem. Soc., 2003, 125, 11607–
11615.
2(a) C. Chatgilialoglu, M. Guerra and Q. C. Mulazzani, Model
studies of DNA C5radicals. Selective generation and reactivity
of 2-deoxyadenosin-5-yl radical, J. Am. Chem. Soc., 2003, 125,
3839–3848; (b) R. Flyunt, R. Bazzanini, C. Chatgilialoglu and
Q. G. Mulazzani, Fate of the 2-deoxyadenosin-5-yl radical under
anaerobic conditions, J. Am. Chem. Soc., 2000, 122, 4225–4226.
3 A. Romieu, D. Gasparutto, D. Molko and J. Cadet, Site-specific
introduction of (5S)-5,8-cyclo-2-deoxyadenosineinto oligodeoxyri-
bonucleotides, J. Org. C h e m ., 1998, 63, 5245–5249.
4 A. Romieu, D. Gasparutto and J. Cadet, Synthesis and char-
acterization of oligonucleotides containing 5,8-cyclopurine 2-
deoxyribonucleosides: (5R)-5,8-cyclo-2-deoxyadenosine, (5S)-5,8-
cyclo-2-deoxyguanosine, and (5R)-5,8-cyclo-2-deoxyguanosine,
Chem. Res. Toxicol., 1999, 12, 412–421.
5(a) I. Kuraoka, C. Bender, A. Romieu, J. Cadet, R. D. Wood
and T. Lindahl, Removal of oxygen free-radical-induced 5,8-purine
cyclodeoxynucleosides from DNA by the nucleotide excision-repair
pathway in human cells, Proc. Natl. Acad. Sci. USA, 2000, 97,
3832–3837; (b) I. Kuraoka, P. Robins, C. Masutani, F. Hanaoka, D.
Gasparutto, J. Cadet, R. D. Wood and T. Lindahl, Oxygen free radical
damage to DNA: translesion synthesis by human DNApolymerase g
and resistance to exonuclease action at cyclopurine deoxynucleoside
residues, J. Biol. Chem., 2001, 276, 49283–49288.
6(a) P. J. Brooks, D. S. Wise, D. A. Berry, J. V. Kosmoski, M. J.
Smerdon, R. L. Somers, H. Mackie, A. Y. Spoonde, E. J. Ackerman,
K. Coleman, R. E. Tarone and J. H. Robbins, The oxidative DNA
lesion 5,8-(S)-cyclo-2-deoxyadenosine is repaired by the nucleotide
excision repair pathway and blocks gene expression in mammalian
cells, J. Biol. Chem., 2000, 275, 22355–22362; (b) K. Randerath,
G. D. Zhou, R. L. Somers, J. H. Robbins and P. J. Brooks, A
32P-postlabeling assay for the oxidative DNA lesion 5,8-cyclo-2-
deoxyadenosine in mammalian tissues. Evidence that four type
II I-compounds are dinucleotides containing the lesion in the 3
nucleotide, J. Biol. Chem., 2001, 276, 36051–36057.
7 It was reported in a note as unpublished results that photolysis at
254 nm of a de-aerated aqueous solution of 2afforded (5S)- and
(5R)-isomers of 1in 2 and 12% yields, respectively (ref. 3).
8 For the ribo analogues, see: C. Chatgilialoglu, M. Duca, C. Ferreri,
M. Guerra, M. Ioele, Q. G. Mulazzani, H. Strittmatter and B. Giese,
Selective generation and reactivityof 5-adenosinyl and 2-adenosinyl
radicals, Chem. Eur. J., 2004, 10, 1249–1255.
9 J. Berkowitz, G. B. Ellison and D. Gutman, Three methods to
measure RH bond energies, J. Phys. Chem., 1994, 98, 2744–2765.
10 X. Fang, R. Mertens and C. von Sonntag, Pulse radiolysis of
aryl bromides in aqueous solutions: Some properties of aryl and
arylperoxyl radicals, J. Chem. Soc., Perkin Trans. 2, 1995, 1033–1036.
11 P. Wardman, Reduction potentials of one-electron couples involving
free radicals in aqueous solution, J. Phys. Chem. Ref. Data, 1989, 18,
1637–1755.
12 Y. Liu, A. S. Pimentel, Y. Antoku, B. J. Giles and J. R.
Barker, Temperature-dependent rate and equilibrium constants for
Br.bul.(aq) +Br-(aq).dblarw. Br2-.bul.(aq), J. Phys. Chem. A, 2002,
106, 11075–11082.
13 Y. Liu, R. L. Sheaffer and J. R. Barker, Effects of temperature and
ionic strength on the rate and equilibrium constants for the reaction
I.bul.aq +I-aq.tautm. I2.bul.-2aq, J. Phys. Chem. A, 2003, 107,
10296–10302.
1046
Photochem. Photobiol. Sci.
, 2004, 3, 1042–1046
Downloaded by CNR Bologna on 09 May 2012
Published on 22 October 2004 on http://pubs.rsc.org | doi:10.1039/B410939B
View Online
... Schéma 17 : Cyclisation des analogues de nucléosides portant une chaîne linéaire [195,253,256,257] . Schéma 18 : Cyclisation du 8-bromoaciclovir [195,261] Schéma 20 : Réaction radicalaire amorcée photochimiquement des nucléosides portant un groupe thiophényle [196,262] Schéma 21 : Intermédiaire clé retrouvé dans la synthèse des analogues nucléosidiques [196,[262][263][264][265][266][267][268][269] Schéma 22 : Synthèse des cyclonucléosides dans lesquels le groupe thiophényle est porté par la base [196,263] Schéma 23 : Synthèse de cyclonucléosides en utilisant des dérivés 5'-céto comme précurseurs pour la formation sélective d'un radical 5'-nucleosidyle [196,270] Schéma 24 : Synthèse « one-pot » de cyclonucléosides à partir de 8-bromo-2'désoxynucléosides [196,[271][272][273][274] Schéma 29 : Synthèse du dérivé 8,2'-éthanocycloadénosine [196,283] ...
... [196,270] Schéma 23 : Synthèse de cyclonucléosides en utilisant des dérivés 5'-céto comme précurseurs pour la formation sélective d'un radical 5'-nucleosidyle [196,270] Une série de cyclonucléosides a été ainsi synthétisée en « one-pot » à partir de 8bromo-2'-désoxynucléosides en série purine par Chatgilialoglu et coll. [196,[271][272][273][274][275][276][277] (Schéma 24). ...
... La formation du radical, issue de la rupture de la liaison C-Br, suivie de la cyclisation et de l'oxydation du radical aminyle intermédiaire conduit aux cyclonucléosides (177), (178), (180) et (181) (Schéma 24). [196,[271][272][273][274][275][276] Schéma 24 : Synthèse « one-pot » de cyclonucléosides à partir de 8-bromo-2'-désoxynucléosides [196, 271- ...
Thesis
Les analogues synthétiques des nucléosides naturels constituant des acides nucléiques occupent une place importante dans le domaine du médicament comme principes actifs antiviraux ou anticancéreux. Ces nucléosides agissent comme « prodrogues » en perturbant la biosynthèse des acides nucléiques viraux ou des cellules cancéreuses après phosphorylation. Dans la recherche de nouveaux médicaments antiviraux, nous avons cherché à synthétiser de nouveaux analogues des nucléosides naturels, les 2-désoxy-adénosine et -guanosine, et de l’aciclovir et de ses dérivés (vanciclovir, ganciclovir…) qui sont très utilisés dans le traitement de l’Herpès. Des premiers travaux en série adénine et guanine, n’ont pas permis d’obtenir les dérivés cycliques recherchés dans lesquels la base et la chaîne latérale introduite en position 9 de la base sont liés par un atome d’oxygène se trouvant en position 8 pour former un nouveau cycle. Quatre analogues cycliques en série guanine ont été synthétisés dans lesquels la base et la chaîne latérale en position 8 sont liés soit par un hétéroatome (préparés par réaction de substitution nucléophile), soit par une liaison carbone-carbone (préparés par réaction radicalaire) et sont en cours d’évaluation antivirale.
... However, the CB potential of BiVO 4 (0.05 eV versus NHE) is more positive than the oxygen reduction potential (E°(O 2 /·O 2 − ) = −0.33 eV versus NHE), revealing that ·O 2 − cannot be formed in the photocatalytic degradation. The holes on the VB of AgBr with a weaker oxidizing ability (2.0 eV versus NHE) will not turn OH − into ·OH (E°(OH − /·OH) = 2.38 eV versus NHE) [48]. All these situations are not in accordance with the radical trapping tests. ...
Article
Full-text available
Successful construction of heterojunction can improve the utilization efficiency of solar light by broadening the absorption range, facilitating charge-carrier separation, promoting carrier transportation and influencing surface-interface reaction. Herein, visible-light-driven AgBr was deposited on the surface of lamellar BiVO4which was prepared by a facile hydrothermal process to improve charge carrier separation, and subsequent photocatalytic effectiveness. The catalyst with an optimal AgBr/BiVO4ratio exhibited a superbly enhanced photocatalytic decolorization ability (about 6.85 times higher than that of pure BiVO4) and high stability after four cycles. The unique photocatalytic mechanism of S-scheme carrier migration was investigated on the bases of radical trapping tests and photo/electrochemical characterizations. Results showed that the enhanced migration strategy and intimately interfacial collaboration guaranteed the effective charge carriers separation/transfer, leading to magnificent photocatalytic performance as well as excellent stability.
... The γ-radiolysis of aqueous solutions of 8-bromo-2'-deoxyadenosine (8-Br-2'-dA) in the presence of the K4Fe(CN)6 gives rise to the formation of cdA with an ratio R/S ratio of 6:1 and a yield of 67% (based on the starting material conversion) [22,23]. After ultraviolet light irradiation in acetonitrile, the yield of formation attained 65%, but with a diastereomeric ratio R/S of 1.7:1 [26]. The photolysis of 8-Br-2'-dG solution with ultraviolet light gave rise to cdG with a yield of 26% and an R/S ratio of 8:1 [21]. ...
Article
Full-text available
Purine 5’,8-cyclo-2’-deoxynucleosides (cPu) are tandem-type lesions observed among the DNA purine modifications and identified in mammalian cellular DNA in vivo. These lesions can be present in two diasteroisomeric forms, 5’R and 5’S, for each 2’-deoxyadenosine and 2’-deoxyguanosine moiety. They are generated exclusively by hydroxyl radical attack to 2′-deoxyribose units generating C5′ radicals, followed by cyclization with the C8 position of the purine base. This review describes the main recent achievements in the preparation of the cPu molecular library for analytical and DNA synthesis applications for the studies of the enzymatic recognition and repair mechanisms, their impact on transcription and genetic instability, quantitative determination of the levels of lesions in various types of cells and animal model systems, and relationships between the levels of lesions and human health, disease, and aging, as well as the defining of the detection limits and quantification protocols.
... However, beside the SSBs that are triggered by the cleavage of the C-Br bond further mechanisms have to be taken into account to understand the interaction between 8Br A and LEEs. The capture of an hydrated electron leads to the formation of 5',8-cycloadenosine [100,101,102] and moreover 12 ...
Thesis
Full-text available
In Germany more than 200.000 persons die of cancer every year, which makes it the second most common cause of death. Chemotherapy and radiation therapy are often combined to exploit a supra-additive effect, as some chemotherapeutic agents like halogenated nucleobases sensitize the cancerous tissue to radiation. The radiosensitizing action of certain therapeutic agents can be at least partly assigned to their interaction with secondary low energy electrons (LEEs) that are generated along the track of the ionizing radiation. In the therapy of cancer DNA is an important target, as severe DNA damage like double strand breaks induce the cell death. As there is only a limited number of radiosensitizing agents in clinical practice, which are often strongly cytotoxic, it would be beneficial to get a deeper understanding of the interaction of less toxic potential radiosensitizers with secondary reactive species like LEEs. Beyond that LEEs can be generated by laser illuminated nanoparticles that are applied in photothermal therapy (PTT) of cancer, which is an attempt to treat cancer by an increase of temperature in the cells. However, the application of halogenated nucleobases in PTT has not been taken into account so far. In this thesis the interaction of the potential radiosensitizer 8-bromoadenine (8BrA) with LEEs was studied. In a first step the dissociative electron attachment (DEA) in the gas phase was studied in a crossed electron-molecular beam setup. The main fragmentation pathway was revealed as the cleavage of the C-Br bond. The formation of a stable parent anion was observed for electron energies around 0 eV. Furthermore, DNA origami nanostructures were used as platformed to determine electron induced strand break cross sections of 8BrA sensitized oligonucleotides and the corresponding nonsensitized sequence as a function of the electron energy. In this way the influence of the DEA resonances observed for the free molecules on the DNA strand breaks was examined. As the surrounding medium influences the DEA, pulsed laser illuminated gold nanoparticles (AuNPs) were used as a nanoscale electron source in an aqueous environment. The dissociation of brominated and native nucleobases was tracked with UV-Vis absorption spectroscopy and the generated fragments were identified with surface enhanced Raman scattering (SERS). Beside the electron induced damage, nucleobase analogues are decomposed in the vicinity of the laser illuminatednanoparticles due to the high temperatures. In order to get a deeper understanding of the different dissociation mechanisms, the thermal decomposition of the nucleobases in these systems was studied and the influence of the adsorption kinetics of the molecules was elucidated. In addition to the pulsed laser experiments, a dissociative electron transfer from plasmonically generated ”hot electrons” to 8BrA was observed under low energy continuous wave laser illumination and tracked with SERS. The reaction was studied on AgNPs and AuNPs as a function of the laser intensity and wavelength. On dried samples the dissociation of the molecule was described by fractal like kinetics. In solution, the dissociative electron transfer was observed as well. It turned out that the timescale of the reaction rates were slightly below typical integration times of Raman spectra. In consequence such reactions need to be taken into account in the interpretation of SERS spectra of electrophilic molecules. The findings in this thesis help to understand the interaction of brominated nucleobases with plasmonically generated electrons and free electrons. This might help to evaluate the potential radiosensitizing action of such molecules in cancer radiation therapy and PTT.
Chapter
The environmental and health hazards created by industrial chemicals and consumer products must be minimized. For safer products to be designed, the relationships between structure and toxicity must be understood at the molecular level. Green chemistry combined with free radical research has the potential to offer innovative solutions to such problems. Some solutions are "greener then others", and many necessitate significant financial investment. New technology will only be adopted if real benefit can be shown and sometimes adaptation of existing methods is the best option. The efficiency of processes must be assessed, not only in terms of the final yield, but also cost, environmental impact and waste toxicity. This practical and concise guide showcases the sustainable methods offered by green free radical chemistry and summarizes the fundamental science involved. It discusses the pros and cons of free radical chemistry in aqueous systems for synthetic applications. All transformation steps are covered including initiation, propagation, and termination. Useful background knowledge is combined with examples, including industrial scale processes for pharmaceuticals and fine chemicals. The book helps chemists to choose appropriate methods for achieving maximum output using a modern, environmentally conscious approach. It shows that, armed with an elementary knowledge of kinetics, an understanding of the mechanistic and technical aspects, and some common sense, it is possible to harness free radicals for use in a broad range of applications. Streamlining Green Free Radical Chemistry is aimed at chemists, engineers, materials scientists, biochemists and biomedical experts, as well as undergraduate and postgraduate students. It encourages readers to question conventional methods and move towards the "Benign-by-Design" approach of the future. References to further reading are provided at the end of each chapter.
Chapter
The environmental and health hazards created by industrial chemicals and consumer products must be minimized. For safer products to be designed, the relationships between structure and toxicity must be understood at the molecular level. Green chemistry combined with free radical research has the potential to offer innovative solutions to such problems. Some solutions are "greener then others", and many necessitate significant financial investment. New technology will only be adopted if real benefit can be shown and sometimes adaptation of existing methods is the best option. The efficiency of processes must be assessed, not only in terms of the final yield, but also cost, environmental impact and waste toxicity. This practical and concise guide showcases the sustainable methods offered by green free radical chemistry and summarizes the fundamental science involved. It discusses the pros and cons of free radical chemistry in aqueous systems for synthetic applications. All transformation steps are covered including initiation, propagation, and termination. Useful background knowledge is combined with examples, including industrial scale processes for pharmaceuticals and fine chemicals. The book helps chemists to choose appropriate methods for achieving maximum output using a modern, environmentally conscious approach. It shows that, armed with an elementary knowledge of kinetics, an understanding of the mechanistic and technical aspects, and some common sense, it is possible to harness free radicals for use in a broad range of applications. Streamlining Green Free Radical Chemistry is aimed at chemists, engineers, materials scientists, biochemists and biomedical experts, as well as undergraduate and postgraduate students. It encourages readers to question conventional methods and move towards the "Benign-by-Design" approach of the future. References to further reading are provided at the end of each chapter.
Chapter
The environmental and health hazards created by industrial chemicals and consumer products must be minimized. For safer products to be designed, the relationships between structure and toxicity must be understood at the molecular level. Green chemistry combined with free radical research has the potential to offer innovative solutions to such problems. Some solutions are "greener then others", and many necessitate significant financial investment. New technology will only be adopted if real benefit can be shown and sometimes adaptation of existing methods is the best option. The efficiency of processes must be assessed, not only in terms of the final yield, but also cost, environmental impact and waste toxicity. This practical and concise guide showcases the sustainable methods offered by green free radical chemistry and summarizes the fundamental science involved. It discusses the pros and cons of free radical chemistry in aqueous systems for synthetic applications. All transformation steps are covered including initiation, propagation, and termination. Useful background knowledge is combined with examples, including industrial scale processes for pharmaceuticals and fine chemicals. The book helps chemists to choose appropriate methods for achieving maximum output using a modern, environmentally conscious approach. It shows that, armed with an elementary knowledge of kinetics, an understanding of the mechanistic and technical aspects, and some common sense, it is possible to harness free radicals for use in a broad range of applications. Streamlining Green Free Radical Chemistry is aimed at chemists, engineers, materials scientists, biochemists and biomedical experts, as well as undergraduate and postgraduate students. It encourages readers to question conventional methods and move towards the "Benign-by-Design" approach of the future. References to further reading are provided at the end of each chapter.
Article
At the density functional theory (DFT) level, addition reactions between the guanine-8-yl radical and its 3'/5' neighboring purine deoxynucleosides forming the purine-purine type intrastrand cross-links were studied. It is found that addition of the guanine-8-yl radical to the C8 site of its 5' neighboring deoxyguanosine or deoxyadenosine is a two-step reaction consisting of a structurally relatively unfavourable conformational transformation step, while the corresponding 3' C8 addition is straightforward and kinetically more efficient. The 3' C8 preference of the guanine-8-yl radical additions indicates the existence of an obvious sequence effect, which is completely opposite to that observed in the formation of pyrimidine radicals induced DNA intrastrand cross-links. The detrimental effects from steric hindrance and stabilizing weak interactions make these addition reactions markedly suppressed in double stranded DNA. This work broadens our knowledge about the possible types of DNA intrastrand cross-links.
Article
Brominated nucleobases sensitize double stranded DNA to hydrated electrons, one of the dominant genotoxic species produced in hypoxic cancer cells during radiotherapy. Such radiosensitizers can therefore be administered locally to enhance treatment efficiency within the solid tumor while protecting the neighboring tissue. When a solvated electron attaches to 8-bromoadenosine, a potential sensitizer, the dissociation of bromide leads to a reactive C8 adenosyl radical known to generate a range of DNA lesions. In the current work, we propose a multiscale computational approach to elucidate the mechanism by which this unstable radical causes further damage in genomic DNA. We employed a combination of classical molecular dynamics conformational sampling and QM/MM metadynamics to study the thermodynamics and kinetics of plausible reaction pathways in a realistic model, bridging between different timescales of the key processes and accounting for the spatial constraints in DNA. The obtained data allowed us to build a kinetic model that correctly predicts the products predominantly observed in experimental settings – cyclopurine and β-elimination (single strand break) lesions – with their ratio and yield dependent on the effective lifetime of the radical species. To date, our study provides the most complete description of purine radical reactivity in double stranded DNA, explaining the radiosensitizing action of electrophilic purines in molecular detail as well as providing a conceptual framework for the computational modeling of competing reaction pathways in biomolecules.
Article
Full-text available
Exposure of cellular DNA to reactive oxygen species generates several classes of base lesions, many of which are removed by the base excision-repair pathway. However, the lesions include purine cyclodeoxynucleoside formation by intramolecular crosslinking between the C-8 position of adenine or guanine and the 5' position of 2-deoxyribose. This distorting form of DNA damage, in which the purine is attached by two covalent bonds to the sugar-phosphate backbone, occurs as distinct diastereoisomers. It was observed here that both diastereoisomers block primer extension by mammalian and microbial replicative DNA polymerases, using DNA with a site-specific purine cyclodeoxynucleoside residue as template, and consequently appear to be cytotoxic lesions. Plasmid DNA containing either the 5'R or 5'S form of 5',8-cyclo-2-deoxyadenosine was a substrate for the human nucleotide excision-repair enzyme complex. The R diastereoisomer was more efficiently repaired than the S isomer. No correction of the lesion by direct damage reversal or base excision repair was detected. Dual incision around the lesion depended on the core nucleotide excision-repair protein XPA. In contrast to several other types of oxidative DNA damage, purine cyclodeoxynucleosides are chemically stable and would be expected to accumulate at a slow rate over many years in the DNA of nonregenerating cells from xeroderma pigmentosum patients. High levels of this form of DNA damage might explain the progressive neurodegeneration seen in XPA individuals.
Article
Aryl radicals were generated pulse-radiolytically in aqueous solutions by reacting the solvated electron with phenyl bromide and substituted phenyl bromides. The aryl radicals react rapidly (k 2.5 × 109 dm3 mol–1 s–1) with O2 giving rise to arylperoxyl radicals which display characteristic absorptions in the visible spectrum (490–600 nm, depending on the substituent). Aryl radicals react with alcohols by H-abstraction. Making use of the strong absorptions of the arylperoxyl radicals in the visible spectrum, rate constants of the reaction of the aryl radicals with various alcohols have been determined by competition kinetics [e.g. 4-methoxyphenyl: 1.1 × 107(2-PrOH), 5.8 × 106(EtOH), 1.1 × 106(MeOH), 5.2 × 105 dm3 mol–1 s–1(ButOH)]. The temperature dependence of the reaction of the 4-CN-phenyl radical was studied and reaction parameters of its reaction with O2(A= 4 × 1011 s–1; Ea= 12 kJ mol–1), propan-2-ol (A= 1.5 × 109 s–1; Ea= 13 kJ mol–1) and tert-butanol (A= 6 × 109 s–1; Ea= 25 kJ mol–1) determined.
Article
Reduction of an electron acceptor (oxidant), A, or oxidation of an electron donor (reductant), A2−, is often achieved stepwise via one-electron processes involving the couples A/A⋅− or A⋅−/A2− (or corresponding prototropic conjugates such as A/AH⋅ or AH⋅/AH2). The intermediate A⋅−(AH⋅) is a free radical. The reduction potentials of such one-electron couples are of value in predicting the direction or feasibility, and in some instances the rate constants, of many free-radical reactions. Electrochemical methods have limited applicability in measuring these properties of frequently unstable species, but fast, kinetic spectrophotometry (especially pulse radiolysis) has widespread application in this area. Tables of ca. 1200 values of reduction potentials of ca. 700 one-electron couples in aqueous solution are presented. The majority of organic oxidants listed are quinones, nitroaryl and bipyridinium compounds. Reductants include phenols, aromatic amines, indoles and pyrimidines, thiols and phenothiazines. Inorganic couples largely involve compounds of oxygen, sulfur, nitrogen and the halogens. Proteins, enzymes and metals and their complexes are excluded.
Article
Aqueous bromine atoms were produced by laser flash photolysis of 1,2-dibromoethane at 248 nm in solutions containing bromide ions. Forward and reverse rate constants of the title reaction were determined as functions of temperature. An analysis of potential sources of systematic errors shows that the measured forward and reverse rate constants have relative uncertainties (±σk/k) of 10 and 25%, respectively, over the temperature range from 10.5 to 50 °C. The Arrhenius parameters are (kf ± 10%) = 5.1 × 1012 exp (−1812/T) M-1 s-1 and (kr ± 25%) = 2.5 × 1010exp (−4068/T) s-1. The equilibrium constant is found from the ratio of kf/kr: (Keq ± 30%) = 2.0 × 102 exp (2256/T) M-1 or (3.9 ± 1.2) × 105 M-1 at 298 K. The reaction entropy and enthalpy are ΔSR° = 44 ± 6 J mol-1 K-1 and ΔHR° = −19 ± 2 kJ mol-1, respectively. The corresponding reaction reduction potential is ΔE° = 0.33 ± 0.01 V, in very good agreement with that calculated from half-cell potentials. In addition, preliminary rate constants for Br2-• + Br2-• → Br3- + Br- and the hydrogen abstraction reaction (Br• + BrCH2CH2Br → •CBrH−CH2−Br + H+ + Br-) are reported.
Article
Temperature-dependent forward (kf) and reverse (kr) constants for the title reaction were measured by analyzing the kinetics of formation and decay of I2•-. Over the temperature range 286−320 K, the Arrhenius parameters are kf = (2.4 ± 0.1) × 1013 exp[(−2324 ± 77)/T] M-1 s-1 and kr = (2.6 ± 0.4) × 1012 exp[(−5157 ± 198)/T] s-1. The equilibrium constant was found from the ratio of kf/kr: Keq = (9.2 ± 1.4) exp[(2833 ± 212)/T] M-1. Of particular interest, ionic strength effects on the rate constant of the title reaction are reported for the first time.
Article
In this paper, the authors compare and contrast three powerful methods for experimentally measuring bond energies in polyatomic molecules. The methods are: radical kinetics; gas phase acidity cycles; and photoionization mass spectroscopy. The knowledge of the values of bond energies are a basic piece of information to a chemist. Chemical reactions involve the making and breaking of chemical bonds. It has been shown that comparable bonds in polyatomic molecules, compared to the same bonds in radicals, can be significantly different. These bond energies can be measured in terms of bond dissociation energies.
Article
Reactions of hydroxyl radicals with DNA form a variety of base and sugar products and 8,5'-cyclopurine 2'-deoxyribonucleoside residues in DNA. Here we report the effect of DNA conformation on the yields of 8,5'-cyclopurine 2'-deoxynucleosides and the ratios of their (5'R)- and (5'S)-diastereomers. Calf thymus DNA in native (double-stranded DNA) or heat-denatured form (single-stranded DNA) was exposed to hydroxyl radicals generated by ionizing radiation in nitrous oxide-saturated phosphate buffer. Doses ranging from 10 to 40 Gy were used to ensure low levels of damage to DNA and thus to preserve its secondary structure in experiments with double-stranded DNA (ds-DNA). After irradiation, DNA was hydrolysed enzymatically to 2'-deoxyribonucleosides. The hydrolysates were dried, trimethylsilylated, and analyzed by capillary gas chromatography-mass spectrometry with selected-ion monitoring. An internal standard was used for quantitative measurements and added to DNA samples prior to enzymatic hydrolysis. The yields of 8,5'-cyclo-2'-deoxyadenosine and 8,5'-cyclo-2'-deoxyguanosine in single-stranded DNA (ss-DNA) were higher than those in ds-DNA. The (5'R)-diastereomers of both compounds were found to predominate over their (5'S)-diastereomers in ss-DNA. In contrast, the yields of the (5'S)-diastereomers in ds-DNA were slightly higher than those of the (5'R)-diastereomers. The G values of 8,5'-cyclo-2'-deoxyadenosine in ss-DNA and ds-DNA were 0.042 and 0.025, respectively. Those of 8,5'-cyclo-2'-deoxyguanosine in ss-DNA and ds-DNA were 0.038 and 0.017, respectively.
Article
Radiation-induced degradation of purine and pyrimidine nucleosides gave rise to carbon-bridged cyclocompounds. Such cyclonucleosides represent a class of tandem lesions in which modification of both the base and 2-deoxyribose has occurred. A solid-phase synthetic method was designed for the incorporation of both 5'R and 5'S diastereoisomers of 5',8-cyclopurine 2'-deoxyribonucleosides into oligodeoxynucleotides to facilitate the assessment of the biochemical and biophysical features of such lesions. We report the preparation of the phosphoramidite synthons of (5'R)-5', 8-cyclo-2'-deoxyadenosine (2), (5'S)-5',8-cyclo-2'-deoxyguanosine (3), and (5'R)-5',8-cyclo-2'-deoxyguanosine (4). Fully protected compounds 10, 18, and 25 were then inserted into several oligonucleotides by automated procedures. Analysis of modified DNA oligomers 26-31 by electrospray mass spectrometry and enzymatic digestions with exo- and endonucleases confirmed the base compositions and the integrity of free radical-induced tandem lesions 2-4 that were chemically inserted.