ArticlePDF Available

Why 6-Iodouridine Cannot Be Used as a Radiosensitizer of DNA Damage? Computational and Experimental Studies

Authors:

Abstract and Figures

Previous density functional theory (DFT) studies on 6-brominated pyrimidine nucleosides suggest that 6-iodo-2'-deoxyuridine (6IdU) should act as a better radiosensitizer than its 5-iodosubstituted 2'-deoxyuridine analogue. In this work, we show that 6IdU is unstable in an aqueous solution. Indeed, a complete disappearance of the 6IdU signal was observed during its isolation by reversed-phase high-performance liquid chromatography (RP-HPLC). As indicated by the thermodynamic characteristics for the SN1-type hydrolysis of 6IdU obtained at the CAM-B3LYP/DGDZVP++ level and the polarizable continuum model (PCM) of water, 6-iodouracil (6IU) was already released quantitatively at ambient temperatures. The simulation of the hydrolysis kinetics demonstrated that a thermodynamic equilibrium was reached within seconds for the title compound. To assess the reliability of the calculations carried out, we synthesized 6-iodouridine (6IUrd), which was, unlike 6IdU, sufficiently stable in an aqueous solution at room temperature. The activation barrier for the N-glycosidic bond dissociation in 6IUrd was estimated experimentally using an Arrhenius plot. The stabilities in water calculated for 6IdU, 6IUrd, and 5-iodo-2'-deoxyuridine (5IdU) could be explained by the electronic and steric effects of the 2'-hydroxy group present in the ribose moiety. Our studies highlight the issue of the hydrolytic stability of potentially radiosensitizing nucleotides which, besides having favorable dissociative electron attachment (DEA) characteristics, must be stable in water to have any practical application.
Content may be subject to copyright.
Why 6Iodouridine Cannot Be Used as a Radiosensitizer of DNA
Damage? Computational and Experimental Studies
Karina Falkiewicz, Witold Kozak, Magdalena Zdrowowicz, Paulina Spisz, Lidia Chomicz-Manka,
Mieczyslaw Torchala, and Janusz Rak*
Cite This: J. Phys. Chem. B 2023, 127, 2565−2574
Read Online
ACCESS Metrics & More Article Recommendations *
Supporting Information
ABSTRACT: Previous density functional theory (DFT) studies
on 6-brominated pyrimidine nucleosides suggest that 6-iodo-2-
deoxyuridine (6IdU) should act as a better radiosensitizer than its
5-iodosubstituted 2-deoxyuridine analogue. In this work, we show
that 6IdU is unstable in an aqueous solution. Indeed, a complete
disappearance of the 6IdU signal was observed during its isolation
by reversed-phase high-performance liquid chromatography (RP-
HPLC). As indicated by the thermodynamic characteristics for the
SN1-type hydrolysis of 6IdU obtained at the CAM-B3LYP/
DGDZVP++ level and the polarizable continuum model (PCM)
of water, 6-iodouracil (6IU) was already released quantitatively at
ambient temperatures. The simulation of the hydrolysis kinetics
demonstrated that a thermodynamic equilibrium was reached
within seconds for the title compound. To assess the reliability of the calculations carried out, we synthesized 6-iodouridine (6IUrd),
which was, unlike 6IdU, suciently stable in an aqueous solution at room temperature. The activation barrier for the N-glycosidic
bond dissociation in 6IUrd was estimated experimentally using an Arrhenius plot. The stabilities in water calculated for 6IdU, 6IUrd,
and 5-iodo-2-deoxyuridine (5IdU) could be explained by the electronic and steric eects of the 2-hydroxy group present in the
ribose moiety. Our studies highlight the issue of the hydrolytic stability of potentially radiosensitizing nucleotides which, besides
having favorable dissociative electron attachment (DEA) characteristics, must be stable in water to have any practical application.
1. INTRODUCTION
Radiotherapy involves damaging genomic DNA by exposing
living cells to ionizing radiation.
1,2
Direct eects, where
ionizing photons reach the DNA molecules, are negligible for
sparsely ionizing radiation, for example, using X-rays, which is
commonly employed in radiotherapy.
3
Since water is the main
constituent of most cells, its radiolysis (producing mainly
hydroxyl radicals and hydrated electrons) accounts for most of
the secondary eects related to interactions between ionizing
radiation and living matter.
4
Hypoxic cells a hallmark of solid
tumors
5
are about 3 times more resistant to ionizing radiation
than normoxic ones.
6
To overcome the state of hypoxia which
is disadvantageous from a radiotherapy standpoint, modified
nucleosides (MNs) with radiosensitizing properties could be
introduced into anticancer therapy.
7
Note that some MNs can
easily penetrate cell membranes, be phosphorylated in the
cytoplasm, and then processed by DNA polymerases,
ultimately leading to their incorporation into the newly
biosynthesized DNA.
3,8,9
Such modified DNA then becomes
sensitive to hydrated electrons, one of the most abundant
radiolysis products of water, which after their attachment may
lead to a single-strand break (SSB).
The most thoroughly studied radiosensitizing nucleoside
analogues employing the above-mentioned mechanism of
DNA sensitization are 5-bromo- (5BrdU) and 5-iodo-2-
deoxyuridine (5IdU).
1012
The sensitizing eects of these
pyrimidines are related to an ecient dissociative electron
attachment (DEA) as a consequence of interactions between
MNs and hydrated electrons.
13
Indeed, hydrated electron
attachment may initiate the DEA process, which will produce
reactive and genotoxic nucleoside radicals. If these reactions
take place inside the DNA, the reactive species resulting from
DEA can lead to DNA strand breaks and finally to cell death.
14
Recently, several new uracil derivatives and 2-deoxyuridines
with radiosensitizing properties have been proposed by our
group.
1518
Most radiosensitizing nucleosides are 5-substituted pyrimi-
dines (5-Pyrs).
8,9,19
However, we have previously postulated
Received: January 24, 2023
Revised: February 14, 2023
Published: March 9, 2023
Articlepubs.acs.org/JPCB
© 2023 The Authors. Published by
American Chemical Society 2565
https://doi.org/10.1021/acs.jpcb.3c00548
J. Phys. Chem. B 2023, 127, 25652574
that 6-substituted uridine or cytidine analogs could work as
DNA radiosensitizers more eciently than 5-Pyrs.
20,21
In fact,
our computational studies have demonstrated that the uridine-
6-yl radical, forming with a low activation barrier DEA to the 6-
substituted pyrimidine nucleosides, can abstract a hydrogen
atom not only from the sugar moiety of an adjacent nucleoside
in a DNA strand but also from its own 2-deoxyribose.
20,21
Thus, the latter process should increase the amount of strand
breaks in DNA labeled with 6-substituted pyrimidines relative
to DNA labeled with 5-substituted ones.
In this work, we have attempted to experimentally follow the
concept indicated in our previous theoretical investiga-
tions,
20,21
i.e., we have tried to synthesize 6-iodo-2-
deoxyuridine (6IdU) and carry out radiolytic studies on its
aqueous solution coupled with the liquid chromatography
mass spectrometry (LCMS) identification of radioproducts.
However, it was observed that the presence of the iodine atom
at the C6 position of 2-deoxyuridine made the compound
susceptible to hydrolysis already at ambient temperatures.
Hence, this observation suggests that 6IdU cannot work as a
radiosensitizer since any radiosensitizer has to be relatively
stable in water to be transported to the cell cytoplasm and then
incorporated into the cellular DNA. To explain the high
observed propensity of 6IdU to hydrolysis, in addition to the
attempted synthesis of 6-iodo-2-deoxyuridine, we synthesized
6-iodouridine (6IUrd) and compared its hydrolytic behavior
and that of 5-iodo-2-deoxyuridine (5IdU) to that of 6IdU (see
Figure 1 for the structures of the studied species). The
mechanism of hydrolysis of the studied compounds was then
described at the density functional theory (DFT) level. The
thermodynamic stimuli and activation barriers were calculated
for the SN1 substitution of the sugar moiety at the C1position
with a water molecule. To validate the assumed level of theory,
the obtained computational barrier was compared to that
determined experimentally with the use of an Arrhenius plot
for a relatively stable 6IUrd nucleoside. The surprising ease of
hydrolysis of 6IdU compared to 6IUrd could be explained in
terms of both electronic factors concerning uracil and the lack
of a hydroxyl group at the C2position in the deoxyribose
moiety.
2. METHODS
2.1. Synthesis. 2.1.1. Synthesis of 2,3-O-Isopropylide-
neuridine 1(See Figure 2). Sulfuric acid (H2SO4, 0.65 mL)
was added dropwise to a stirred suspension of uridine (1 g,
4.13 mmol) in acetone (50 mL) at room temperature (RT),
and the resulting mixture was stirred for 3 h. Then, the
reaction was neutralized with Et3N and evaporated. The raw
product was purified by preparative column chromatography
using CHCl3/MeOH 15:1 as an eluent to produce the desired
compound 1(1.05 g) in a 90.2% yield.
2.1.2. Synthesis of 2,3-O-Isopropylidene-5-O-methox-
ymethyluridine 2(See Figure 2). Dimethoxymethane (13.33
mL, 150.5 mmol) and methanesulfonic acid (0.133 mL, 2.05
mmol) were added to a suspension of 1(0.5 g, 1.76 mmol) in
acetone (6.5 mL), and the resulting mixture was stirred
overnight. The stirred mixture was then poured into a 25%
aqueous solution of NH4OH and extracted with CHCl3. The
raw product was purified by preparative column chromatog-
raphy using CH2Cl2/MeOH 80:1 as an eluent to produce the
desired compound 2(385 mg) in a 66.7% yield.
2.1.3. Synthesis of 6-Iodo-2,3-O-isopropylidene-5-O-
methoxymethyluridine 3(See Figure 2). A solution of 2
(300 mg, 0.91 mmol) in anhydrous tetrahydrofuran (THF, 3.5
mL) was added dropwise to a stirred solution of lithium
diisopropylamide (LDA, 0.95 mL, 2.00 mmol, 2 M solution in
THF) in anhydrous THF (3.7 mL) at 78 °C. After stirring
for 1 h, iodine (232 mg, 0.91 mmol) in anhydrous THF (2.9
mL) was added and the mixture was stirred for an additional 5
h at 78 °C. Then, AcOH (0.45 mL) was added to quench the
reaction. After bringing to RT, the raw product was extracted
with AcOEt and the organic layer was washed with an aqueous
solution consisting of NaHCO3and brine. The product was
purified by preparative column chromatography using hexane/
AcOEt 7:6 as an eluent to produce the desired compound 3
(140 mg) in a 31.7% yield.
2.1.4. Synthesis of 6-Iodouridine 4(See Figure 2). A stirred
suspension of 3(100 mg, 0.22 mmol) in water (0.77 mL) was
treated with a 50% aqueous solution of trifluoroacetic acid
(TFA, 1.15 mL) at 0 °C. Then, the reaction mixture was
stirred for 48 h at RT. Following evaporation, the product was
purified by preparative column chromatography using CHCl3/
EtOH 10:1 as an eluent to produce the desired compound 4
Figure 1. Chemical structure of 5-iodo-2-deoxyuridine, 6-iodo-2-
deoxyuridine, and 6-iodouridine.
Figure 2. Synthesis of 6-iodouridine 4.
The Journal of Physical Chemistry B pubs.acs.org/JPCB Article
https://doi.org/10.1021/acs.jpcb.3c00548
J. Phys. Chem. B 2023, 127, 25652574
2566
(48 mg) in a 58.9% yield. Proton nuclear magnetic resonance
(1H NMR, Bruker AVANCE III, 500 MHz, DMSO), δ:13C
NMR (125 MHz, DMSO), δ: 162.5, 147.6, 116.1, 115.9,
102.6, 84.9, 72.0, 69.9, 62.3 (Figure S1). High-resolution mass
spectrometry (HRMS) (TripleTOF 5600+, SCIEX), m/z: [M
H]calcd for C9H11IN2O6368.9588, found 368.8943;
ultraviolet (UV) spectrum (water), λmax: 270 nm.
2.2. Kinetic Measurements and Calculations of
Kinetic Profiles. The vials with the reaction mixture (50
μL) containing 6-iodouridine in a phosphate buer (0.1 M, pH
= 7.0) were placed in an Eppendorf thermocycler for specified
periods of time (see time points in Figure S8) at the following
temperatures: 60, 65, 70, 75, and 80 °C. After heating, the
solutions were analyzed by high-performance liquid chroma-
tography (HPLC). The HPLC separations were performed on
a Dionex UltiMate 3000 System with a diode array detector,
which was set at 260 nm for monitoring the euents. The
analytes were separated on a Wakopak Handy ODS column
(4.6 mm ×150 mm) using an isocratic elution with 0.1%
formic acid in deionized water. The flow rate was set at 1 mL
min1. Each solution was injected into the HPLC instrument
using an autosampler. The experiment was carried out in
duplicate. The identification of the thermal analysis products
was performed by mass spectrometry (LCMS and LCMS/
MS experiments). The MS and MS/MS spectra are presented
in Supporting Information Figures S2S7. The LCMS
conditions were as follows: a TripleTOF 5600+ (SCIEX)
mass spectrometer (operating in a negative mode) coupled
with a Nexera X2 ultrahigh-performance liquid chromatog-
raphy (UHPLC) system; a Kinetex column (Phenomenex,
C18, 2.1 mm ×150 mm); a flow rate of 0.3 mL min1; an
isocratic elution with 0.1% formic acid in deionized water. All
MS and MS/MS analyses were performed using a spray voltage
of 4.5 kV and a source temperature of 300 °C.
The logarithmic values of the concentrations of 6IUrd at
time zero and at dierent time intervals were used to establish
the degradation plots. The degradation kinetic parameters (the
degradation rate constants) at 60, 65, 70, 75, and 80 °C were
derived from the plots. The predicted kinetics for the
degradation of 6IUrd at 25 °C was extrapolated from the
Arrhenius plot.
2.3. Quantum Chemical Calculations. The mechanisms
of hydrolysis of the three considered derivatives were studied
using the DFT. We applied the CAM-B3LYP functional
22
combined with the DGDZVP++ basis set.
23
The polarization
continuum model
24
(PCM) was used to mimic an aqueous
reaction environment. All geometries were optimized and
subjected to the frequency calculations for the standard state
(T= 298 K, p= 1 atm). The analysis of harmonic frequencies
confirmed that the stationary geometries were localized (all
force constants were positive for minima; all but one were
positive for the first-order transition states). The intrinsic
reaction coordinate (IRC)
25
procedure was used to verify that
the transition state connected the proper minima.
The changes in Gibbs free energy (ΔG) for particular
reactions were calculated as the dierence between the Gibbs
free energy of the products and substrates. The activation
barriers (ΔG*) were calculated as the dierence between the
Gibbs free energies of the transition state and the substrate. All
DFT computations were performed with the use of Gaussian
09.
26
2.4. Calculations of Equilibrium Concentrations and
Kinetics. The equilibrium concentrations of 5IdU, 6IdU, and
6IUrd at 25 °C were calculated using the following set of
equilibrium (see eqs 13and the Equilibrium Concentrations
Section in the Supporting Information)
++
nucleoside base sugarV
(1)
+ +
+ +
sugar 2 H O sugar (OH) H O
2 3
V
(2)
+ +base H O base (H) HO
2V
(3)
where Kdiss,Ka, and Kbstand for the equilibrium constants for
nucleoside dissociation eq 1 and cationic eq 2 and anionic eq 3
hydrolysis, respectively.
The cationic (Ka) and anionic (Kb) hydrolysis constants
(Table 1) were estimated employing the experimental values of
acid dissociation constants (Kref) for dimethoxycarbene
(MeO)2CHOH (pKa
27
=5.7) and uracil (pKa
28
= 9.45),
respectively, using eq 4.
29
=K K G
RT
p p 2.303
a/ b ref
(4)
where pKa/b and pKref stand for the negative decimal logarithm
of Ka/Kband Kref, respectively, while ΔΔGindicates the
change in Gibbs free energy for the reactions defined in Table
1.
These equilibrium constants were, in turn, used to calculate
some of the rate constants eq 5 that were required to simulate
the overall kinetics of hydrolysis using eq 5
=Kk
k
1
1
(5)
where k1and k1are the rate constants for forward and reverse
reactions, respectively, while Kis the respective equilibrium
constant.
The assumed elemental reactions running in the solution of
a nucleoside eqs 13led to the following set of ordinary
dierential equations eqs 615
[ ] = [ ][ ] [ ]
+
t
k k
d Nuc
d
2 R A 1 Nuc
(6)
[ ] = [ ][ ] [ ][ ]
[ ][ ] + [ ]
+
+ +
+
t
k k k
k
d H O
d3 R H O 4 R(OH) H O 7
H O OH 8 H O
3
2
2
3
3 2
2
(7)
Table 1. pKand KValues for the Given and Ref Compounds along with the Respective ΔΔG[kcal mol1]
reaction ΔΔGpK K
5IU+ U 5IU + U4.1 pKb= 7.53 Kb= 3.0 ×108
6IU+ U 6IU + U7.3 pKb= 9.91 Kb= 1.2 ×1010
deoxyribose++ (MeO)2CHOH deoxyribose + (MeO)2CHOH+9.1 pKa=12.36 Ka= 2.3 ×1012
ribose++ (MeO)2CHOH ribose + (MeO)2CHOH+15.5 pKa=17.09 Ka= 1.2 ×1017
The Journal of Physical Chemistry B pubs.acs.org/JPCB Article
https://doi.org/10.1021/acs.jpcb.3c00548
J. Phys. Chem. B 2023, 127, 25652574
2567
(8)
[ ] = [ ][ ] + [ ] [ ][ ] +
[ ][ ]
+
+ +
+
t
k k k k
d R
d2 R A 1 Nuc 3 R H O 4
R(OH) H O
2
2
3
(9)
[ ] = [ ][ ] [ ][ ]
+ +
t
k k
d R(OH)
d3 R H O 4 R(OH) H O
2
2
3
(10)
[ ] = [ ][ ] + [ ] [ ][ ] +
[ ][ ]
+
t
k k k k
d A
d2 R A 1 Nuc 5 A H O 6
A(H) OH
2
(11)
[ ] = [ ][ ] [ ][ ]
t
k k
d A(H)
d5 A H O 6 A(H) OH
2
(12)
[ ] = [ ][ ] + [ ][ ] +
[ ][ ] [ ][ ] + [ ]
[ ] [ ]
+ + +
t
k k k
k k
k
d H O
d5 A H O 6 A(H) OH 4
R(OH) H O 3 R H O 7 H O
OH 8 H O
2
2
3 2
2
3
2
2
(13)
Additional Equations for 6IUrd Only
[ ] = [ ][ ] + [ ][ ]
t
k k
d hydrNuc
d10 hydrNuc I 9 Nuc OH
(14)
[ ] = [ ][ ] + [ ][ ]
t
k k
d I
d10 hydrNuc I 9 Nuc OH
(15)
The abbreviations used in this system of kinetic equations
were the following: Nucnucleoside (substrate), A(H)
nucleic base, R(OH)sugar moiety, Aanionic form of
nucleic base, R+cationic form of sugar moiety, and
hydrNuc6-hydroxyuridine (6OHUrd). The system of dier-
ential equations, shown in eqs 615, was solved with the use of
the Octave program.
30,31
The initial concentration of a
nucleoside was assumed to be 103M, and the initial
concentrations of the remaining reactants were set to 0 M at
time = 0. Due to the small concentrations of reactants, we
assumed a constant concentration of water of 55.5 M over the
course of the studied processes.
32
3. RESULTS AND DISCUSSION
One of the basic reactions that nucleosides may undergo in an
aqueous solution is their hydrolysis resulting in the detachment
of a nucleobase (see Figure 3).
According to Kochetkov,
33
2-deoxynucleosides are hydro-
lyzed faster than ribonucleosides (about 1001000 times).
Indeed, the half-life of hydrolysis, τ1/2, for 2-deoxyuridine
(dU) amounts to 104 min, while uridine (Urd) is hydrolyzed
only to a small extent even in the 3 times longer period under
the same experimental conditions (5% hydrolysis in 5 h).
33
Moreover, the identity of a nucleobase plays an important role
in the stability of the N-glycosidic bond. Actually, the
hydrolysis of purine nucleosides is faster than that of
pyrimidine ones.
33
The rate of hydrolysis is also determined
by the type of substituent present in the nucleobase moiety. In
fact, the degree of cleavage of the N-glycosidic bond for 2-
deoxycitidine (dC) is 76%, for dU3%, and for 5XdU, where
X = F, Cl, or Br, 1317%, for a reaction occurring in 5%
trichloroacetic acid for 30 min at 100 °C.
33
Schroeder et al.
34
demonstrated that the experimental value of the activation
Gibbs free energy, ΔG*, for the heterolytic dissociation of the
C1′−N1 glycosidic bond in dU amounted to 30.5 kcal mol1.
In turn, Przybylski et al.
35
carried out computational studies for
this reaction using the B3LYP/6-31+G(d,p) level of theory
and obtained a value of ΔG*similar to that found
experimentally, i.e., 28.7 kcal mol1.
To explain the observed hydrolytic instability of 6IdU, we
initially decided to use four functionals, including the B3LYP
one, as recommended by Przybylski et al.
35
and Chen et al.;
36
we then calculated the free energy of hydrolysis for dU and
three other derivatives considered in the current work (see
Table 2). The B3LYP functional is a classical choice and has
been widely used in studies of the glycosidic bond cleavage in
nucleic acid-type systems.
3739
To describe this type of bond
dissociation, Chen et al.
36
also employed other DFT
functionals, including long-range corrected hybrid functionals
CAM-B3LYP
22
and wB97XD.
40
These authors obtained the
best accuracy at the CAM-B3LYP level (the lowest mean
absolute error among the considered functionals). As indicated
by the results gathered in Table 2, the B3LYP and CAM-
B3LYP functionals led to the most similar values compared to
the experimental estimate of ΔG*for dU. On the other hand,
calculations for 6IUrd proved that CAM-B3LYP provided the
value that was most compatible with the experimental
activation energy, as shown in the current study (see Table 2
and the Probable Mechanism of Hydrolysis Section). There-
fore, we settled to use CAM-B3LYP for the ultimate estimates
of the kinetic characteristics of the nucleosides considered in
this study.
3.1. Probable Mechanism of Hydrolysis. It has been
reported in the literature that attempts to obtain 6-aryl-2-
deoxyuridines through a 6-iodo-2-deoxyuridine (6IdU)
derivative (both hydroxyls protected with the 1,1,3,3-
tetraisopropoxydisiloxanylidene (TIPDS) group) had failed.
41
Indeed, treating the TIPDS-protected 6IdU with tetrabuty-
lammonium fluoride (TBAF) or ammonium fluoride (NH4F)
Figure 3. Hydrolysis of a nucleoside.
33
Table 2. Activation Free Energy [kcal mol1] (ΔG*) of the
Heterolytic Dissociation of the C1N1 Glycosidic Bond
(Figure 3) for the Selected DFT Functionals
a
system
functional dU 5IdU 6IdU 6IUrd
B3LYP 28.2 24.3 11.9 17.9
CAM-B3LYP 34.2 30.5 18.0 24.1
ωB97XD 37.2 33.5 21.5 28.7
experimental 30.5 26.8
a
All calculations were conducted using the PCM solvation model and
the DGDZVP++ basis set. The CAM-B3LYP results are highlighted
in bold.
The Journal of Physical Chemistry B pubs.acs.org/JPCB Article
https://doi.org/10.1021/acs.jpcb.3c00548
J. Phys. Chem. B 2023, 127, 25652574
2568
produced 6-iodouracil and other degradation products.
38
The
above-mentioned findings suggest the low stability of 6IdU.
Nevertheless, we attempted to synthesize 6IdU to verify
whether or not we could capture the moment of its
degradation. First, dU was treated with two equivalents of t-
butyldimethylsilyl chloride (TBDMSCl) in the presence of
imidazole. In the next step, the TBDMS-protected dU was
introduced into a mixture of LDA and iodine at 78 °C
yielding 2,5-di-t-butyldimethylsilyl-6-iodo-2-deoxyuridine. In
the last step, the deprotection of the 2and 5-OH groups was
performed with TBAF. The reaction was monitored by thin-
layer chromatography (TLC), which showed the appearance of
the desired product. We attempted to isolate this product with
RP-HPLC (in a gradient of acetonitrileACNand water)
but registered only 6IU as a degradation product of 6IdU.
Obviously, the interaction of 6IdU with water led to the swift
conversion of the synthesized nucleoside into the 6-substituted
base (6IU). To describe the probable mechanism of this
degradation, we decided to synthesize 6IUrd which, according
to the kinetic data, shown in Table 2, should have been
suciently stable in water to allow for its kinetic stability to be
assessed (ΔG*= 24.1 kcal mol1for 6IUrd vs ΔG*= 18.0 kcal
mol1for 6IdU at the CAM-B3LYP level). The 6IUrd, 4, was
synthesized through a modification of procedures available in
the literature
42,43
and the protocol of its synthesis is described
in detail in Section 2 (also see Figure 2). The synthesized
nucleoside was then subjected to kinetic studies. The
measurements were carried out at elevated temperatures so
that the reaction proceeded at convenient times (a significant
Figure 4. High-performance liquid chromatography (HPLC) traces of the thermal degradation of 6-iodouridine over temperatures ranging from 60
to 80 °C after 120 min of heating.
Figure 5. Arrhenius plot for 6IUrd hydrolysis proceeding in the 60
80 °C range.
Table 3. Thermodynamic and Activation Barriers [kcal
mol1] of 6OHUrd Formation (Figures S9S11)
substrates products ΔGΔG*
6IUrd, OH6OHUrd, I72.3 11.8
6IUrd, H2O 6OHUrd, HI 11.7 43.1
6IUrd, 2H2O 6OHUrd, I, H3O+10.4 39.2
The Journal of Physical Chemistry B pubs.acs.org/JPCB Article
https://doi.org/10.1021/acs.jpcb.3c00548
J. Phys. Chem. B 2023, 127, 25652574
2569
degradation at relatively short times). After specific periods of
time had elapsed, the reaction was “frozen” by a rapid lowering
of the temperature to an ambient one. Such a procedure
allowed the activation barrier of hydrolysis to be determined
experimentally. Figure 4 shows the HPLC traces for the
reaction mixture after its 120 min conditioning at several
elevated temperatures. Here, we also observed small amounts
of 6-hydroxyuridine (6OHUrd; see Figure 4) in addition to the
formation of 6-iodouracil. Both products were identified by
LCMS.
The above-described kinetic measurements revealed a
dependence of the 6IUrd concentration vs time at several
temperatures, which allowed the kinetic constants for each
considered temperature to be estimated (see Figure S8). In
turn, these estimated rate constants allowed us to construct an
Arrhenius plot (Figure 5) and, finally, to calculate the
activation energy at 298 K. Derived in such a way, the
experimental activation energy was equal to 26.8 kcal mol1
and remained in good agreement with the ΔG*value of 24.1
kcal mol1(see Table 1) calculated at the CAM-B3LYP level.
As indicated in Figure 4, the hydrolysis of 6IUrd led to two
stable products: 6IU (product of the hydrolysis of the N-
glycosidic bond, Figure 3) and 6OHUrd. Three possible
mechanisms yielding 6OHUrd were considered (Figures S9
S11). All of the analyzed reactions were thermodynamically
allowed (Table 3;ΔG< 0).
However, with one water molecule, the system ended up
with HI in addition to 6OHUrd, and the value of the activation
barrier reached 43.1 kcal mol1(Table 3). In the model
comprising two water molecules, the predicted barrier of 39.2
kcal mol1(Table 3) is somewhat lower than that calculated
for a single water molecule but was still too high to allow the
reaction to be completed over the experimental times (τ1/2 for
an activation energy of the order of 40 kcal mol1at 25 °C is
equal to ca. 7 ×109years). Finally, the activation barrier for
the nucleophilic substitution involving the hydroxyl anion
rather than water molecule(s) amounted to only 11.8 kcal
mol1(Table 3). Thus, although the concentration of OHin
the studied 6IUrd solution was below 107M, the direct attack
of OHon the 6IU moiety (SN2 mechanism) was probably
responsible for the formation of 6OHUrd as a side-product.
A general mechanism for nucleoside hydrolysis which
describes the hydrolysis of 6IdU, 5IdU, and 6IUrd specifically
is depicted in Figure 6. The main reaction (I) consisted of a
heterolytic dissociation of the C1′−N1 glycosidic bond, which
led to the carbocation of the sugar moiety and carbanion of the
substituted uracil. The subsequent steps involved the
hydrolyses of the cationic (II) and anionic (III) products of
Figure 6. Elemental reactions for the studied mechanism of hydrolysis of the three considered derivatives.
The Journal of Physical Chemistry B pubs.acs.org/JPCB Article
https://doi.org/10.1021/acs.jpcb.3c00548
J. Phys. Chem. B 2023, 127, 25652574
2570
the C1′−N dissociation. Moreover, 6-hydroxyuridine was
formed through the 6IUrd hydrolysis (V).
3.2. Simulation of Kinetic Profiles. The equations that
enabled the equilibrium concentrations in the reaction mixture
Figure 7. Concentrationtime plots for 5IdU (A), 6IdU (B), and 6IUrd (C). Nucnucleoside, Abase, Rsugar moiety, Aand R+anionic
and cationic forms of the base and sugar moieties.
The Journal of Physical Chemistry B pubs.acs.org/JPCB Article
https://doi.org/10.1021/acs.jpcb.3c00548
J. Phys. Chem. B 2023, 127, 25652574
2571
to be calculated are derived in the Equilibrium Concentrations
section in the Supporting Information, while the equilibrium
characteristics are summarized for 5IdU, 6IdU, and 6IUrd in
Table S2. However, one should note that these data do not
contain any information about the time which elapsed from the
dissolution of a nucleoside in water and the moment these
concentrations occurred. It is worth emphasizing that these
times were measures of the stability of the studied compounds
and would help decide about the possible usage of a nucleoside
as a radiosensitizer. Hence, to determine the time profiles of
hydrolysis, we performed kinetic simulations by integrating a
system of ordinary dierential eqs 615 and using rate
constants obtained as indicated in Section 2.4 for each of the
three studied nucleosides (see Table S1).
Figure 7 shows the concentration profiles plotted against the
time of hydrolysis reaction. For all studied systems, a decrease
in the concentration of a nucleoside (Nuc) and a proportional
increase in the concentration of a base (A) and sugar product
(R) were observed.
The equilibrium concentration of 6IU assumed a signifi-
cantly lower value than the initial concentration of 6IUrd as
was also indicated by the experimental data (see Figure 4). In
the case of 5IdU, the equilibrium was attained before the
complete hydrolysis of the substrate. Our kinetic simulations
led to τ1/2 being equal to 13.8 h and 1.4 s for 6IUrd and 6IdU,
respectively, while for 5IdU, τ1/2 was not defined. One could,
however, determine the time after which the equilibrium
concentrations were reached (42% of the substrate decom-
poses) to be ca. 126 years. The data obtained thus explain the
observed stabilities of 6IUrd and, in particular 5IdU, and the
significant instability of 6IdU in an aqueous solution. To
confirm the accuracy of the kinetic simulations performed here,
the calculated equilibrium concentrations shown in Table 4
were compared with those obtained through the kinetic
calculations. As demonstrated by these data, both concen-
trations were in excellent agreement.
At first glance, such a profound influence of subtle structural
dierences between the studied systems seems surprising, but
it could be explained by steric and electronic eects. Indeed,
since the hydration energies were almost identical for the
studied nucleosides, the lower stability of 6IdU relative to
5IdU was associated with the larger steric hindrance in the 6-
substituted derivative (the repulsion between the iodine atoms
and the sugar residue was much smaller in 5IdU); in turn, the
lower stability of the ribose cation relative to the 2-
deoxyribose one made the half-life of 6IdU significantly
shorter than that of 6IUrd. Indeed, the free energy of
nucleoside dissociation (eq I in Figure 6) was ca. 6.4 kcal
mol1lower at the CAM-B3LYP/PCM level for 6IdU than for
6IUrd. The presence of the OH group in the 2position of the
ribose moiety, exerting a negative inductive eect, was
probably responsible for the lower stability of their respective
carbocations.
4. SUMMARY
This paper presents a combination of theoretical and
experimental studies devoted to understanding the aqueous
stabilities of specific modified nucleosides with potentially
radiosensitizing properties. This investigation was prompted by
our previous findings that suggested that 6-substituted uracils
were better radiosensitizers than their 5-substituted analogues.
Our attempts to synthesize 6IdU were unsuccessful, which
was ascribed to the low stability of the compound in water (the
synthesized nucleoside was isolated and purified with RP-
HPLC using an ACN:water mobile phase).
The experimental activation energy and as a consequence,
the half-life and the calculated free activation energy remained
in good agreement for 6IUrd. The half-life of the latter
determined by kinetic simulations amounted to 13.8 h (see
Figure 7). On the other hand, predicted τ1/2’s of 1.4 s and 126
years were estimated for 6IdU and 5IdU, respectively. The
above-mentioned dierences in τ1/2 were likely due to
dissimilarities between steric and electronic eects of the
considered molecules.
Since slight structural dierences between nucleosides led to
huge variations in their stabilities in water, to design a working
radiosensitizer, one has to estimate the aqueous stability of
considered nucleosides alongside performing calculations of
their DEA profiles
3
before the actual synthesis. Therefore, the
proposed kinetic model should be employed in future studies
devoted to the rational engineering of new radiosensitizers.
ASSOCIATED CONTENT
*
Supporting Information
The Supporting Information is available free of charge at
https://pubs.acs.org/doi/10.1021/acs.jpcb.3c00548.
1H NMR spectrum; high-resolution mass spectra of 6-
iodouridine, 6-hydroxyuridine, and 6-iodouracil; reac-
tion profiles for 6OHUrd formation from 6IUrd and
OH, single water molecule and two water molecules;
calculations for equilibrium concentrations and table
Table 4. Equilibrium Concentrations [mol dm3] (Equilibrium; for Details, See the Equilibrium Concentration section in the
Supporting Information) and the Corresponding Values Resulting from the Integration of Eqs 615 (Kinetics)
5IdU 6IdU 6IUrd
reagent equilibrium kinetics
a
equilibrium kinetics
a
equilibrium kinetics
a
nucleoside [Nuc] 5.8 ×1045.8 ×1043.6 ×1010 9.0 ×1010 2.4 ×1010 4.1 ×109
base [A] 4.1 ×1044.3 ×1047.5 ×1047.5 ×1047.5 ×1047.5 ×104
base anion [A] 1.2 ×1051.1 ×1052.5 ×1042.5 ×1042.5 ×1042.5 ×104
sugar moiety [R] 4.2 ×1044.2 ×1041.0 ×1031.0 ×1031.0 ×1031.0 ×103
sugar moiety cation [R+] 2.2 ×1021 2.3 ×1021 1.1 ×1019 1.1 ×1019 2.1 ×1024 2.1 ×1024
[OH] 8.5 ×1010 7.8 ×1010 4.0 ×1011 4.0 ×1011 4.0 ×1011 4.0 ×1011
[H3O+] 1.2 ×1051.3 ×1052.5 ×1042.5 ×1042.5 ×1042.5 ×104
6OHUrd [hydrNuc] 1.6 ×104
iodide anion [I] 1.6 ×104
a
Small dierences between the equilibrium and kinetic concentrations of some regents were due to the finite integration time of the kinetic
equations.
The Journal of Physical Chemistry B pubs.acs.org/JPCB Article
https://doi.org/10.1021/acs.jpcb.3c00548
J. Phys. Chem. B 2023, 127, 25652574
2572
with their values; and system of dierential equations for
the kinetic profiles in the Octave program (PDF)
AUTHOR INFORMATION
Corresponding Author
Janusz Rak Laboratory of Biological Sensitizers, Department
of Physical Chemistry, Faculty of Chemistry, University of
Gdansk, 80-308 Gdansk, Poland; orcid.org/0000-0003-
3036-0536; Email: janusz.rak@ug.edu.pl
Authors
Karina Falkiewicz Laboratory of Biological Sensitizers,
Department of Physical Chemistry, Faculty of Chemistry,
University of Gdansk, 80-308 Gdansk, Poland
Witold Kozak Laboratory of Biological Sensitizers,
Department of Physical Chemistry, Faculty of Chemistry,
University of Gdansk, 80-308 Gdansk, Poland; orcid.org/
0000-0003-3253-5555
Magdalena Zdrowowicz Laboratory of Biological
Sensitizers, Department of Physical Chemistry, Faculty of
Chemistry, University of Gdansk, 80-308 Gdansk, Poland
Paulina Spisz Laboratory of Biological Sensitizers,
Department of Physical Chemistry, Faculty of Chemistry,
University of Gdansk, 80-308 Gdansk, Poland; Laboratory of
Intermolecular Interactions, Department of Bioinorganic
Chemistry, Faculty of Chemistry, University of Gdansk, 80-
308 Gdansk, Poland
Lidia Chomicz-Manka Laboratory of Biological Sensitizers,
Department of Physical Chemistry, Faculty of Chemistry,
University of Gdansk, 80-308 Gdansk, Poland
Mieczyslaw Torchala Laboratory of Biological Sensitizers,
Department of Physical Chemistry, Faculty of Chemistry,
University of Gdansk, 80-308 Gdansk, Poland; orcid.org/
0000-0002-4542-9156
Complete contact information is available at:
https://pubs.acs.org/10.1021/acs.jpcb.3c00548
Notes
The authors declare no competing financial interest.
ACKNOWLEDGMENTS
This work was supported by the Polish National Science
Center (NCN) under grant no. UMO-2014/14/A/ST4/
00405. The calculations were carried out using resources
provided by the Wroclaw Centre for Networking and
Supercomputing (https://wcss.pl), grant No. 209.
REFERENCES
(1) Santivasi, W. L.; Xia, F. Ionizing Radiation-Induced DNA
Damage, Response, and Repair. Antioxid. Redox Signaling 2014,21,
251259.
(2) Borrego-Soto, G.; Ortiz-Lopez, R.; Rojas-Martinez, A. Ionizing
Radiation-Induced DNA Injury and Damage Detection in Patients
with Breast Cancer. Genet. Mol. Biol. 2015,38, 420432.
(3) Rak, J.; Chomicz, L.; Wiczk, J.; Westphal, K.; Zdrowowicz, M.;
Wityk, P.; Zyndul, M.; Makurat, S.; Golon, Ł. Mechanisms of Damage
to DNA Labeled with Electrophilic Nucleobases Induced by Ionizing
or UV Radiation. J. Phys. Chem. B 2015,119, 82278238.
(4) Nair, C. K. K.; Parida, D. K.; Nomura, T. Radioprotectors in
Radiotherapy. J. Radiat. Res. 2001,42, 2137.
(5) Thomlinson, R. H.; Gray, L. H. The Histological Structure of
Some Human Lung Cancers and the Possible Implications for
Radiotherapy. Br. J. Cancer 1955,9, 539549.
(6) Dasu, A.; Denekamp, J. New Insights into Factors Influencing
the Clinically Relevant Oxygen Enhancement Ratio. Radiother. Oncol.
1998,46, 269277.
(7) Wardman, P. Chemical Radiosensitizers for Use in Radiotherapy.
Clin. Oncol. 2007,19, 397417.
(8) Chomicz, L.; Zdrowowicz, M.; Kasprzykowski, F.; Rak, J.;
Buonaugurio, A.; Wang, Y.; Bowen, K. H. How to Find Out Whether
a 5-Substituted Uracil Could Be a Potential DNA Radiosensitizer. J.
Phys. Chem. Lett. 2013,4, 28532857.
(9) Makurat, S.; Chomicz-Manka, L.; Rak, J. Electrophilic 5-
Substituted Uracils as Potential Radiosensitizers: A Density Func-
tional Theory Study. ChemPhysChem 2016,17, 25722578.
(10) Djordjevic, B.; Szybalski, W. Genetics of Human Cell Lines: III.
Incorporation of 5-Bromo- and 5-Iododeoxyuridine into the
Deoxyribonucleic Acid of Human Cells and Its Effect on Radiation
Sensitivity. J. Exp. Med. 1960,112, 509531.
(11) Shewach, D. S.; Lawrence, T. S. Deoxynucleoside Analogs in
Cancer Therapy; Humana Press: Totowa, New Jersey, USA, 2006; pp
289330.
(12) Park, Y.; Polska, K.; Rak, J.; Wagner, J. R.; Sanche, L.
Fundamental Mechanisms of DNA Radiosensitization: Damage
Induced by Low-Energy Electrons in Brominated Oligonucleotide
Trimers. J. Phys. Chem. B 2012,116, 96769682.
(13) Pan, X.; Cloutier, P.; Hunting, D.; Sanche, L. Dissociative
Electron Attachment to DNA. Phys. Rev. Lett. 2003,90, No. 208102.
(14) Goz, B. The Effects of Incorporation of 5-Halogenated
Deoxyuridines into the DNA of Eukaryotic Cells. Pharmacol. Rev.
1977,29, 249272.
(15) Meißner, R.; Makurat, S.; Kozak, W.; Limao-Vieira, P.; Rak, J.;
Denifl, S. Electron-Induced Dissociation of the Potential Radio-
sensitizer 5-Selenocyanato-2-Deoxyuridine. J. Phys. Chem. B 2019,
123, 12741282.
(16) Makurat, S.; Spisz, P.; Kozak, W.; Rak, J.; Zdrowowicz, M. 5-
Iodo-4-Thio-2-Deoxyuridine as a Sensitizer of X-ray Induced Cancer
Cell Killing. Int. J. Mol. Sci. 2019,20, No. 1308.
(17) Ameixa, J.; ArthurBaidoo, E.; Meißner, R.; Makurat, S.;
Kozak, W.; Butowska, K.; da Silva, F. F.; Rak, J.; Denifl, S. Low-
Energy Electron-Induced Decomposition of 5-Trifluoromethanesul-
fonyl-uracil: A Potential Radiosensitizer. J. Chem. Phys. 2018,149,
No. 164307.
(18) Makurat, S.; Zdrowowicz, M.; Chomicz-Manka, L.; Kozak, W.;
Serdiuk, I. E.; Wityk, P.; Kawecka, A.; Sosnowska, M.; Rak, J. 5-
Selenocyanato and 5-Trifluoromethanesulfonyl Derivatives of 2-
Deoxyuridine: Synthesis, Radiation and Computational Chemistry as
well as Cytotoxicity. RSC Adv. 2018,8, 2137821388.
(19) Wen, Z.; Peng, J.; Tuttle, P. R.; Ren, Y.; Garcia, C.; Debnath,
D.; Rishi, S.; Hanson, C.; Ward, S.; Kumar, A.; et al. Electron-
Mediated Aminyl and Iminyl Radicals from C5 Azido-Modified
Pyrimidine Nucleosides Augment Radiation Damage to Cancer Cells.
Org. Lett. 2018,20, 74007404.
(20) Golon, Ł.; Chomicz, L.; Rak, J. Electron-Induced Single Strand
Break in the Nucleotide of 5- and 6-Bromouridine. A DFT study.
Chem. Phys. Lett. 2014,612, 289294.
(21) Golon, Ł.; Chomicz, L.; Rak, J. The Radiosensitivity of 5- and
6-Bromocytidine Derivatives Electron Induced DNA Degradation.
Phys. Chem. Chem. Phys. 2014,16, 1942419428.
(22) Yanai, T.; Tew, D.; Handy, N. A New Hybrid Exchange-
Correlation Functional Using the Coulomb-Attenuating Method
(CAM-B3LYP). Chem. Phys. Lett. 2004,393, 5157.
(23) Godbout, N.; Salahub, D. R.; Andzelm, J.; Wimmer, E.
Optimization of Gaussian-Type Basis Sets for Local Spin Density
Functional Calculations. Part I. Boron Through Neon, Optimization
Technique and Validation. Can. J. Chem. 1992,70, 560571.
(24) Tomasi, J.; Mennucci, B.; Cammi, R. Quantum Mechanical
Continuum Solvation Models. Chem. Rev. 2005,105, 29993093.
(25) Hratchian, H. P.; Schlegel, H. B. Theory and Applications of
Computational Chemistry: The First 40 Years; Dykstra, C. E., Ed.;
Elsevier, 2005; pp 195249.
The Journal of Physical Chemistry B pubs.acs.org/JPCB Article
https://doi.org/10.1021/acs.jpcb.3c00548
J. Phys. Chem. B 2023, 127, 25652574
2573
(26) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.;
Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Mennucci,
B.; Petersson, G. A.et al. Gaussian 09, Revision D.01; Gaussian, Inc.:
Wallingford, CT, USA, 2013.
(27) Guthrie, J. P.; More O’Ferrall, R. A.; O’Donoghue, A. C.;
Waghorne, W. E.; Zrinski, I. Estimation of a pKa for Protonated
Dimethoxycarbene. J. Phys. Org. Chem. 2003,16, 582587.
(28) Ilyina, M. G.; Khamitov, E. M.; Mustafin, A. G.; Khursan, S. L.
A Theoretical Quantitative Estimation of Acidity of Uracil and Its
Derivatives Through the pKa Values. J. Chin. Chem. Soc. 2018,65,
14471452.
(29) Tran, N. L.; Colvin, M. E. The Prediction of Biochemical Acid
Dissociation Constants Using First Principles Quantum Chemical
Simulations. J. Mol. Struct.: THEOCHEM 2000,532, 127137.
(30) Bambase, M. E., Jr.; Nakamura, N.; Tanaka, J.; Matsumura, M.
Kinetics of Hydroxide-Catalyzed Methanolysis of Crude Sunflower
Oil for the Production of Fuel-Grade Methyl Esters. J. Chem. Technol.
Biotechnol. 2007,82, 273280.
(31) Jiang, C.; Garg, S.; Waite, T. D. Hydroquinone-Mediated
Redox Cycling of Iron and Concomitant Oxidation of Hydroquinone
in Oxic Waters under Acidic Conditions: Comparison with Iron-
Natural Organic Matter Interactions. Environ. Sci. Technol. 2015,49,
1407614084.
(32) Jencks, W. P.; Caplow, M.; Gilchrist, M.; Kallen, R. G.
Equilibrium Constants for the Synthesis of Hydroxamic Acids.
Biochemistry 1963,2, 13131320.
(33) Kochetkov, N. K. Organic Chemistry of Nucleic Acids: Part B;
Plenum Press: London and New York, 1972; pp 425445.
(34) Schroeder, G. K.; Wolfenden, R. Rates of Spontaneous
Disintegration of DNA and the Rate Enhancements Produced by
DNA Glycosylases and Deaminases. Biochemistry 2007,46, 13638
13647.
(35) Przybylski, J. L.; Wetmore, S. D. Designing an Appropriate
Computational Model for DNA Nucleoside Hydrolysis: A Case Study
of 2-Deoxyuridine. J. Phys. Chem. B 2009,113, 65336542.
(36) Chen, H-Y.; Yang, P-Y.; Chen, H-F.; Kao, C-L.; Liao, L-W.
DFT Reinvestigation of DNA Strand Breaks Induced by Electron
Attachment. J. Phys. Chem. B 2014,118, 1113711144.
(37) Millen, A. L.; Archibald, L. A.; Hunter, K. C.; Wetmore, S. D. A
Kinetic and Thermodynamic Study of the Glycosidic Bond Cleavage
in Deoxyuridine. J. Phys. Chem. B 2007,111, 38003812.
(38) Millen, A. L.; Wetmore, S. D. Glycosidic Bond Cleavage in
Deoxynucleotides A Density Functional Study. Can. J. Chem. 2009,
87, 850863.
(39) Li, X.; Sanche, L.; Sevilla, M. D. Base Release in Nucleosides
Induced by Low-Energy Electrons: A DFT Study. Radiat. Res. 2006,
165, 721729.
(40) Chai, J.-D.; Head-Gordon, M. Long-Range Corrected Hybrid
Density Functionals with Damped Atom-Atom Dispersion Correc-
tions. Phys. Chem. Chem. Phys. 2008,10, 66156620.
(41) Kögler, M.; De Jonghe, S.; Herdewijn, P. Synthesis of 6-Aryl-2-
Deoxyuridine Nucleosides via a Liebeskind Cross-Coupling Method-
ology. Tetrahedron Lett. 2012,53, 253255.
(42) Hayakawa, H.; Tanaka, H.; Miyasaka, T. Lithiation of 5,6-
Dihydrouridine: A New Route to 5-Substituted Uridines. Tetrahedron
1985,41, 16751683.
(43) Bello, A. M.; Poduch, E.; Fujihashi, M.; Amani, M.; Li, Y.;
Crandall, I.; Hui, R.; Lee, P. I.; Kain, K. C.; Pai, E. F.; Kotra, L. P. A
Potent, Covalent Inhibitor of Orotidine 5-Monophosphate Decar-
boxylase with Antimalarial Activity. J. Med. Chem. 2007,50, 915921.
The Journal of Physical Chemistry B pubs.acs.org/JPCB Article
https://doi.org/10.1021/acs.jpcb.3c00548
J. Phys. Chem. B 2023, 127, 25652574
2574
Article
Full-text available
We provide an experimentalist’s perspective on the present state-of-the-art in the studies of low-energy electron interactions with common radiosensitizers, including compounds used in combined chemo-radiation therapy and their model systems....
Article
Full-text available
Nucleosides, especially pyrimidines modified in the C5-position, can act as radiosensitizers via a mechanism that involves their enzymatic triphosphorylation, incorporation into DNA, and a subsequent dissociative electron attachment (DEA) process. In this paper, we report 5-iodo-4-thio-2′-deoxyuridine (ISdU) as a compound that can effectively lead to ionizing radiation (IR)-induced cellular death, which is proven by a clonogenic assay. The test revealed that the survival of cells, pre-treated with 10 or 100 µM solution of ISdU and exposed to 0.5 Gy of IR, was reduced from 78.4% (for non-treated culture) to 67.7% and to 59.8%, respectively. For a somewhat higher dose of 1 Gy, the surviving fraction was reduced from 68.2% to 54.9% and to 40.8% for incubation with 10 or 100 µM ISdU, respectively. The cytometric analysis of histone H2A.X phosphorylation showed that the radiosensitizing effect of ISdU was associated, at least in part, with the formation of double-strand breaks. Moreover, the cytotoxic test against the MCF-7 breast cancer cell line and human dermal fibroblasts (HDFa line) confirmed low cytotoxic activity of ISdU. Based on the results of steady state radiolysis of ISdU with a dose of 140 Gy and quantum chemical calculations explaining the origin of the MS detected radioproducts, the molecular mechanism of sensitization by ISdU was proposed. In conclusion, we found ISdU to be a potential radiosensitizer that could improve anticancer radiotherapy.
Article
Full-text available
Two classes of azido-modified pyrimidine nucleosides were synthesized as potential radiosensitizers; one class is 5-azidomethyl-2'-deoxyuridine (AmdU) and cytidine (AmdC), while the second class is 5-(1-azidovinyl)-2'-deoxyuridine (AvdU) and cytidine (AvdC). The addition of radiation-produced electrons to C5-azido nucleosides leads to the formation of π-aminyl radicals followed by facile conversion to σ-iminyl radicals either via a bimolecular reaction involving intermediate α-azidoalkyl radicals in AmdU/AmdC or by tautomerization in AvdU/AvdC. AmdU demonstrates effective radiosensitization in EMT6 tumor cells.
Article
Full-text available
5-trifluoromethanesulfonyl-uracil (OTfU), a recently proposed radiosensitizer, is decomposed in the gas-phase by attachment of low-energy electrons. OTfU is a derivative of uracil with a triflate (OTf) group at the C5-position, which substantially increases its ability to undergo effective electron-induced dissociation. We report a rich assortment of fragments formed upon dissociative electron attachment (DEA), mostly by simple bond cleavages (e.g., dehydrogenation or formation of OTf−). The most favorable DEA channel corresponds to the formation of the triflate anion alongside with the reactive uracil-5-yl radical through the cleavage of the O–C5 bond, particularly at about 0 eV. Unlike for halouracils, the parent anion was not detected in our experiments. The experimental findings are accounted by a comprehensive theoretical study carried out at the M06-2X/aug-cc-pVTZ level. The latter comprises the thermodynamic thresholds for the formation of the observed anions calculated under the experimental conditions (383.15 K and 3 × 10−11 atm). The energy-resolved ion yield of the dehydrogenated parent anion, (OTfU–H)−, is discussed in terms of vibrational Feshbach resonances arising from the coupling between the dipole bound state and vibrational levels of the transient negative ion. We also report the mass spectrum of the cations obtained through ionization of OTfU by electrons with a kinetic energy of 70 eV. The current study endorses OTfU as a potential radiosensitizer agent with possible applications in radio-chemotherapy.
Article
Full-text available
In this paper, we propose a computational scheme for the theoretical estimation of gas‐phase acidity of uracil and its derivatives. The calculation of the acidities (рKа) of the compounds under study was performed using quantum chemical calculations with the composite G3MP2B3 method. The solvent effect was taken into account within the polarizable continuum model and density functional theory calculations by the Polarizable continuum model (solvate model density) ‐ the exchange functional of Tao, Perdew, Staroverov, and Scuseria method. We evaluated the influence of the nonspecific solvation using two approaches (single‐point calculations versus full optimization of the studied structures) and found that the use of full optimization of the geometry of the compound for calculating the solvent effect significantly enhances the accuracy of numerical estimation of the рKа values. The mean absolute deviation decreases by 0.60 and 0.37 units of рKа in the case of the single‐point and full‐optimization approaches, respectively. The most pronounced advantage of the latter approach is its universality. Indeed, because of the implicit accounting of solvation, this computational scheme may be applied to the calculations of the рKа values of any class of compounds without reservations. The proposed computational scheme for the estimation of acidity opens new opportunities for further studies in the field of acidity.
Article
Full-text available
5-Selenocyanato-2′-deoxyuridine (SeCNdU) and 5-trifluoromethanesulfonyl-2′-deoxyuridine (OTfdU) have been synthesized and their structures have been confirmed with NMR and MS methods. Both compounds undergo dissociative electron attachment (DEA) when irradiated with X-rays in an aqueous solution containing a hydroxyl radical scavenger. The DEA yield of SeCNdU significantly exceeds that of 5-bromo-2′-deoxyuridine (BrdU), remaining in good agreement with the computationally revealed profile of electron-induced degradation. The radiolysis products indicate, in line with theoretical predictions, Se–CN bond dissociation as the main reaction channel. On the other hand, the DEA yield for OTfdU is slightly lower than the degradation yield measured for BrdU, despite the fact that the calculated driving force for the electron-induced OTfdU dissociation substantially overpasses the thermodynamic stimulus for BrdU degradation. Moreover, the calculated DEA profile suggests that the electron attachment induced formation of 5-hydroxy-2′-deoxyuridine (OHdU) from OTfdU, while 2′-deoxyuridine (dU) is mainly observed experimentally. We explained this discrepancy in terms of the increased acidity of OTfdU resulting in efficient deprotonation of the N3 atom, which brings about the domination of the OTfdU(N3–H)⁻ anion in the equilibrium mixture. As a consequence, electron addition chiefly leads to the radical dianion, OTfdU(N3–H)˙²⁻, which easily protonates at the C5 site. As a result, the C5–O rather than O–S bond undergoes dissociation, leading to dU, observed experimentally. A negligible cytotoxicity of the studied compounds toward the MCF-7 cell line at the concentrations used for cell labelling calls for further studies aiming at the clinical use of the proposed derivatives.
Article
Full-text available
Although 5-bromo-2'-deoxyuridine (5BrdU) possesses a significant radiosensitizing power in vitro, clinical studies did not confirm any advantages of radiotherapy employing 5BrdU. This situation calls for a continuous search for efficient radiosensitizers. Using the proposed mechanism of radiosensitization by 5BrdU, we propose a series of 5-substituted uracils, XYU, that should undergo an efficient dissociative electron attachment. The DFT calculated thermodynamic and kinetic data concerning the XYU degradations induced by electron addition suggests that some of the scrutinized derivatives have much better characteristics than 5BrdU itself. Synthesis of these promising candidates for radiosensitizers, followed by studies of their radiosensitizing properties in DNA context, and ultimately in cancer cells, are further steps to confirm their potential applicability in anticancer treatment.
Article
Full-text available
Breast cancer is the most common malignancy in women. Radiotherapy is frequently used in patients with breast cancer, but some patients may be more susceptible to ionizing radiation, and increased exposure to radiation sources may be associated to radiation adverse events. This susceptibility may be related to deficiencies in DNA repair mechanisms that are activated after cell-radiation, which causes DNA damage, particularly DNA double strand breaks. Some of these genetic susceptibilities in DNA-repair mechanisms are implicated in the etiology of hereditary breast/ovarian cancer (pathologic mutations in the BRCA 1 and 2 genes), but other less penetrant variants in genes involved in sporadic breast cancer have been described. These same genetic susceptibilities may be involved in negative radiotherapeutic outcomes. For these reasons, it is necessary to implement methods for detecting patients who are susceptible to radiotherapy-related adverse events. This review discusses mechanisms of DNA damage and repair, genes related to these functions, and the diagnosis methods designed and under research for detection of breast cancer patients with increased radiosensitivity.
Article
5-selenocyanato-2’-deoxyuridine (SeCNdU) is a recently proposed radiosensitizer based on 2’-deoxyuridine (dU) with the electron-affinic selenocyanato (–SeCN) side group attached at the C5 position of uracil. As for radiation damage electron interaction processes are an important source of DNA damage, we have studied low-energy dissociative electron attachment processes to SeCNdU in the gas phase. Negative ion formation has been obtained by means of mass spectrometry, where a rich fragmentation pattern is observed even at ~0 eV. The reaction pathways exhibiting the highest ion yields are C4N2O2H2Se●- and CN⁻, both involving a cleavage of the Se–CN bond. The heaviest fragment anion observed is C9N2O5H10Se●-, where besides the charged species also the hydrogen and cyano radical are formed. Further decomposition channels also yield the highly reactive hydroxyl radical, which possesses a high DNA damage potential. All observed channels have experimentally determined onsets at 0 eV, which are supported by calculations performed at the M06-2X/aug-cc-pVTZ level. The calculations comprise the thermochemical thresholds at standard and experimental (428.15 K, 3 × 10⁻¹¹ atm) conditions together with the adiabatic electron affinities. The present study shows that low-energy electrons very effectively decompose SeCNdU upon attachment of thermal electrons producing a large variety of charged fragments and radicals.
Article
Reactions of hydrolysis of glycosidic bonds in nucleosides, nucleotides, and nucleic acids have in the past played an important role in determination of the structure of these compounds. They are widely used nowadays to analyse the nucleotide composition of nucleic acids and to study their primary structure.
Article
Interactions of 1,4-hydroquinone with soluble iron species over a pH range of 3-5 are examined here. Our results show that 1,4-hydroquinone reduces Fe(III) in acidic conditions, generating semiquinone radicals ( ) that can oxidize Fe(II) back to Fe(III). The oxidation rate of Fe(II) by increases with increase in pH due to the speciation change of with its deprotonated form ( ) oxidizing Fe(II) more rapidly than the protonated form ( ). Although oxygenation of Fe(II) is negligible at pH < 5, O2 still plays an important role in iron redox transformation by rapidly oxidizing to form benzoquinone ( ). A kinetic model is developed to describe the transformation of quinone and iron under all experimental conditions. The results obtained here are compared with those obtained in our previous studies of iron- Suwannee River Fulvic Acid (SRFA) interactions in acidic solutions and support the hypothesis that hydroquinone moieties can reduce Fe(III) in natural waters. However, the semiquinone radicals generated in pure hydroquinone solution rapidly oxidize in the presence of dioxygen while the semiquinone radicals generated in SRFA solution are resistant to oxidation by dioxygen with the result that steady state semiquinone concentrations in SRFA solutions are 2-3 orders of magnitude greater than in solutions of 1,4-hydroquinone. As a result, semiquinone moieties in SRFA play a much more important role in iron redox transformations than is the case in solutions of simple quinones such as 1,4-hydroquinone. This difference in steady state concentration of semiquinone species has a dramatic effect on the cycling of iron between the +II and +III oxidation states with iron turnover frequencies in solutions containing SRFA 10-20 times higher than observed in solutions of 1,4-hydroquinone.