You are on page 1of 12

Dalton

Transactions
View Article Online
PAPER View Journal | View Issue

Photophysical, photooxidation, and biomolecule-


Published on 03 January 2022. Downloaded by Universidad de Granada on 8/22/2023 9:44:23 PM.

Cite this: Dalton Trans., 2022, 51,


interaction of meso-tetra(thienyl)porphyrins
1646 containing peripheral Pt(II) and Pd(II) complexes.
Insights for photodynamic therapy applications†
Isadora Tisoco,a Maria Carolina Donatoni,b Henrique Fernandes Vieira Victória,c
José Roberto de Toledo,c Klaus Krambrock,c Otávio Augusto Chaves,d
Kleber Thiago de Oliveira b and Bernardo Almeida Iglesias *a

We report the synthesis and characterization of two novel tetra-cationic porphyrins, containing Pt(II) or Pd
(II) polypyridyl complexes attached at the peripheral position of N4-macrocycle. Compounds were charac-
terized through elemental analysis, molar conductivity, cyclic voltammetry, and spectroscopy analysis.
Photophysical and photobiological parameters were also evaluated. Also, the binding capacity of each
porphyrin with human serum albumin (HSA) was determined by UV–Vis, steady-state, and time-resolved
fluorescence spectroscopy, combined with molecular docking calculations. The results suggest that the
interaction of these compounds is spontaneous, weak to moderate, and probably occurs at site III (subdo-
Received 21st October 2021, main IB) by non-covalent forces, including van der Waals and H-bonding. Moreover, porphyrins contain-
Accepted 30th December 2021
ing peripheral complexes improve their interactions with biomolecules, show good photostability, gene-
DOI: 10.1039/d1dt03565g rate reactive oxygen species under white light studied by electron paramagnetic resonance (EPR) analysis,
rsc.li/dalton and promote photo-damage of HSA.

1. Introduction tion mechanisms: type I which produces radical species, e.g.,


hydroxyl, superoxide, and hydroperoxide (•OH, O2•−, •OOH,
Porphyrins are organic macrocycles widely used as sensitizers respectively) by electron/hole transfer, while type II might gen-
in photodynamic processes, including photodynamic therapy erates singlet oxygen (1O2) by energy transfer processes.3 Due
of cancer (PDT) and antimicrobial photodynamic therapy to their high reactivity, ROS can oxidize different biological
(aPDT), due to their good photostability, high reactive oxygen substrates and biomacromolecules, including nucleic acids,
species (ROS) yield, and specific interaction with target bio- lipoproteins, and albumin.4–7
macromolecules.1 Different physicochemical factors determine The meso-substituted porphyrins containing thienyl units (S
both safe-use and applications of different photosensitizers atoms) can be an excellent alternative for the formation of
(PS) in the photobiology, e.g., number of charges, aggregation novel peripherical cationic metalloporphyrins due to the
state, and hydrophobicity character.2 The ROS formation is a effective complex stabilization based on bonds between thiol
fundamental role in PDT assays since the combination of group and soft metal ions, e.g., Pt(II), Pd(II), and Ru(II) ions.8
photosensitizer (PS), oxygen (O2), and adequate light will Cationic porphyrins containing polypyridyl peripheral tran-
produce reactive species. In this case, there are two main reac- sition metal complexes such as Pt(II) or Pd(II) are good ROS
generators and their applications in photooxidative processes
have already been reported in the literature.9–12 To date, there
a are few reports based on tetra-( pyridyl)porphyrin macrocycle
Department of Chemistry, Federal University of Santa Maria, Av. Roraima, Santa
Maria-RS, Brazil. E-mail: bernardopgq@gmail.com, bernardo.iglesias@ufsm.br as main core for photodynamic assays.13–18
b
Department of Chemistry, Federal University of São Carlos, Rod. Washington Luiz, Some circulatory proteins such as human or bovine serum
São Carlos-SP, Brazil albumins (HSA and BSA, respectively) are the most abundant
c
Department of Physics, Federal University of Minas Gerais, Av. Antônio Carlos, Belo plasma proteins in the mammal’s bloodstream, being widely
Horizonte-MG, Brazil
d applied for pharmacokinetic studies of potential drugs.19
Department of Chemistry, University of Coimbra, Rua Larga, Coimbra, Portugal
† Electronic supplementary information (ESI) available. See DOI: 10.1039/ Serum albumin plays several important roles, including regu-
d1dt03565g lation of plasma pressure, transport, and biodistribution of

1646 | Dalton Trans., 2022, 51, 1646–1657 This journal is © The Royal Society of Chemistry 2022
View Article Online

Dalton Transactions Paper

different types of both endogenous or exogenous compounds, = 0.15, λexc = 522 nm)24 as the Φf standard as the fluorescence
including hormones, fatty acids, vitamins, heme type, metab- yield and eqn (1) was used to determine the Φf values:
olites, and drugs.20 The binding capacity of albumins to metal
I ð1  10A Þstd η2
ions and coordination compounds reveals considerable inter- ΦF ¼ ΦFstd ð1Þ
est in its physiological role as a metal-binding protein, being Istd ð1  10A Þ η2std
one of the most intriguing and potentially useful properties where Φf, I, A, and η are the fluorescence quantum yield, inte-
available that impact the solubility and pharmacokinetics of gral area of fluorescence, absorbance in λexc, and refractive
inorganic compounds.21 Regarding the photooxidation of pro- index of selected solvents (DCM = 1.4241 and DMSO = 1.479).
teins, different techniques e.g., fluorescence, UV-Vis or elec- The subscript “std” refers to the standard molecule.
Published on 03 January 2022. Downloaded by Universidad de Granada on 8/22/2023 9:44:23 PM.

tron paramagnetic resonance (EPR) are often applied. In this Fluorescence lifetime decays (τf ) were recorded using time
way, it was shown by EPR spin trapping and fluorescence emis- correlated single photon counting (TCSPC) method with
sion the photooxidation capacity of some tetra-cationic Pd(II) DeltaHub controller in conjunction with Horiba spectrofluorom-
porphyrin derivatives, indicating that these compounds are eter. Data was processed with DAS6 and Origin® 8.5 software
able to produce singlet oxygen (1O2) species under white-light using exponential (mono-exponential) fitting of raw data.
irradiation resulting in photodamage of BSA, which indicates NanoLED (Horiba) source (1.0 MHz, pulse width <1.3 ns at
a good alternative for photodynamic processes.22 455 nm excitation wavelength) was used as an excitation source.
Due to the increase of the importance of porphyrins in PDT In this way, radiative (kr) and non-radiative (knr) constants can be
assays, the present work reports the synthesis of two novel determined by knowing the fluorescence quantum yield and fluo-
meso-tetra-(2-thienyl)porphyrins coordinated to peripheral Pt rescence lifetime, as following eqn (2) and (3):25
(II) or Pd(II) complexes, namely H2TTP, PtbpyTTP and
PdbpyTTP. Their characterization by elemental analysis, molar kr ¼ ϕf =τf ð2Þ
conductivity, cyclic voltammetry, and spectroscopy analysis knr ¼ ð1  ϕf Þ=τf ð3Þ
were conducted. The photophysical/photobiological properties,
including aggregation, solution stability, photostability, singlet
2.3. Aggregation study by UV-Vis analysis
oxygen quantum yield, and water/octanol coefficient partition
were also determined. Finally, the HSA-binding capacity was UV-Vis absorption experiments were conducted as a function
evaluated by spectroscopy methods combined with molecular of successive increase of porphyrin concentration (0 to 60 μM)
docking calculations, and the albumin photooxidation by EPR in DMSO solution and changes in the λmax in the 250–700 nm
analysis. range were monitored, according to the related literature.15

2.4. Stability in solution by UV-Vis assays

2. Experimental section The stability experiments in pure dimethyl sulfoxide (DMSO)


solution and in DMSO(5%)/Tris-HCl mixture buffer ( pH 7.4) of
2.1. General
related porphyrin was also monitored by absorption electronic
All chemical reagents were of analytical grade and purchased UV–Vis measurements at several days (up to one week). All
from Sigma-Aldrich® and Oakwood Chemical® (USA) without experiments were performed in duplicate and independently.
any further purification. The HSA was lyophilized powder and
fatty acid-free (Sigma-Aldrich®; purity ≥ 99%). The concen- 2.5. Photostability assays and singlet oxygen quantum yield
tration of the stock solutions of albumin was confirmed by (ΦΔ) determination
UV-Vis analysis through the Beer–Lambert equation with the The photostability experiments in dimethyl sulfoxide (DMSO)
molar absorptivity (ε) value of 35 700 M−1 cm−1 at 280 nm in solution of related porphyrin was also monitored by absorp-
Tris-HCl buffer ( pH 7.4) solution23 and the water used in all tion electronic UV–Vis measurements at different exposure
experiments was milliQ grade. times (0 to 30 min) under white-light LED array system (400 at
800 nm) at fluence rate of 50 mW cm−2 and a total light
2.2. Photophysical characterization dosage of 90 J cm−2. All experiments were performed in dupli-
UV-Vis absorption spectroscopy analysis was recorded using a cate and independently.
Shimadzu UV-2600 spectrophotometer (data interval, 1.0 nm) According to typical 1,3-diphenylisobenzofuran (DPBF)
in dimethyl sulfoxide (DMSO) as solvent at 250–800 nm range singlet oxygen quencher using in photodegradation experi-
(concentration of 2.0 µM). Steady-state fluorescence emission ments, the maximum volume of 1.0 mL which contained
spectra for the porphyrins in DMSO were measured in a 100 μM DPBF in DMSO was mixed with 0.5 mL (50 μM) of por-
Horiba Yvon-Jobin Fluoromax Plus (Em/Exc; slit 5.0 mm) at phyrins H2TTP, PtbpyTTP and PdbpyTTP. The flask was then
600–800 nm range (concentration of 1.0 µM). Fluorescence filled with 2.0 mL of DMSO to a final volume of 3.5 mL. In
quantum yield values (Φf ) values of the derivatives H2TTP, order to measure singlet oxygen generation, absorption UV-Vis
PtbpyTTP and PdbpyTTP were determined by comparing the spectra of each solution were recorded at different exposure
corrected fluorescence spectra with those obtained by meso- times (0 to 600 s, using red-light diode laser; Thera Laser DMC
tetra( phenyl)porphyrin (TPP) in dichloromethane solution (Φf – São Paulo; potency of 100 mW) according to the literature.26

This journal is © The Royal Society of Chemistry 2022 Dalton Trans., 2022, 51, 1646–1657 | 1647
View Article Online

Paper Dalton Transactions

The singlet oxygen production quantum yield (ΦΔ) was calcu- steady-state fluorescence emission measurements at 298.15,
lated applying eqn (4): 310.15, and 318.15 K in DMSO(5%)/Tris-HCl buffer ( pH 7.4) in
the 300 to 550 nm range. It is well known that fluorescence
ΦΔ ¼ Φstd
Δ k=k
std
 I std =I ð4Þ quenching might occurs by static or dynamic processes, which
where I /I = (1–10 )/(1–10 ), Φstd
std Astd A
is the singlet oxygen can be determined by Stern–Volmer analysis. To analyze the
Δ
quantum yield of standard sample (meso-tetra( phenyl)por- data from the fluorescence quenching experiments, the Stern–
phyrin TPP in DMF, Φstd 27 Volmer equation (8) was used:
Δ = 0.66), k and kstd are the photode-
gradation kinetic constants for porphyrins and TPP (standard), F0
respectively. Astd and A are the absorbances for TPP and por- ¼ 1 þ Ksv ½Q ¼ 1 þ kq τ0 ½Q ð8Þ
Published on 03 January 2022. Downloaded by Universidad de Granada on 8/22/2023 9:44:23 PM.

F
phyrin, respectively.
where F0 and F are the fluorescence intensities in the absence
and presence of a quencher, respectively. KSV, kq, τ0 and [Q]
2.6. Water/n-octanol partition coefficients (log POW)
denote Stern–Volmer constant, quenching rate constant, the
The partition coefficient of porphyrins was determined using original lifetime of HSA (5.67 × 10−9 s)30 and the concentration
octanol (3.0 mL) and water (3.0 mL), according to the litera- of quencher, respectively. According to eqn (8), the KSV values
ture.28 A mass of 0.5 mg of each porphyrin was added to the were calculated from the slope and kq is equal to KSV/τ0.
mixture and stirred vigorously for 24 h at room temperature. In order to verify a certain trend in the association of por-
Then, the mixture was centrifuged to separate the organic phyrins with the biomolecule, association constant (Ka) values
phase from the aqueous phase. The porphyrin concentrations were obtained by modified Stern–Volmer equation (9):
as well as their respective absorbances were determined by the
absorption spectrum in the UV-Vis region (250–800 nm), and F0 1 1
¼ þ ð9Þ
the Soret band was chosen for monitoring. The log POW value F0  F fKa ½Q f
was calculated from eqn (5), with Aorg and Aaq being the where F0 and F are the steady-state fluorescence intensities of
maximum absorbances of the Soret band in the organic and HSA in the absence and presence of each porphyrin derivative,
aqueous phases, and Vorg and Vaq being the final volumes of respectively. Ka, f and [Q] are the modified Stern–Volmer
the organic and aqueous phases, respectively. binding constants; the fraction of the initial fluorescence that
log P OW ¼ log½ðAorg =Aaq ÞðV aq =V org Þ ð5Þ is accessible to the quencher ( f ≈ 1.00) and porphyrin deriva-
tive concentration, respectively.
In order to obtain quantitative information on the thermo-
2.7. HSA-binding assays by UV-Vis absorption analysis dynamic parameters of the binding between HSA and each
UV-Vis absorption spectra for each porphyrin without and in porphyrin, van’t Hoff equation (10) was applied.31 Gibbs’ free
the presence of successive additions of HSA solution were energy was calculated using the same eqn (7), but Ka values
obtained at 298.15 K in DMSO(5%)/Tris-HCl buffer ( pH 7.4) instead Kb:
mixture solution, in the 250 to 700 nm range. The porphyrins ΔH° ΔS°
concentration was fixed in 1.0 μM and HSA concentration was ln Ka ¼  þ : ð10Þ
RT R
in the 0 to 60 μM range. The hyperchromicity (H%), bathochro-
mic shift (Δλ), binding constant (Kb), and Gibb’s free-energy
2.9. Time-resolved fluorescence decay with HSA
(ΔG°) values of the studied porphyrins were calculated accord-
ing to the literature, through Benesi–Hildebrand and free- Fluorescence lifetime decays (τf ) were recorded using time cor-
energy equations (eqn (6) and (7)).25,29 related single photon counting (TCSPC) method with
DeltaHub controller in conjunction with Horiba spectrofluo-
½HSA ½HSA 1 rometer. Data was processed with DAS6 and Origin® 8.5 soft-
¼ þ ð6Þ
jðεa  εf Þj jðεb  εf Þj Kb jðεb  εf Þj ware using exponential (mono-exponential) fitting of raw data.
ΔG° ¼ RT ln K b ð7Þ NanoLED (Horiba) source (1.0 MHz, pulse width <1.2 ns at
284 nm excitation wavelength) was used as an excitation
where [HSA], εf, εb, and εa are the HSA concentration (in µM), the source. The HSA and porphyrins concentration were fixed in
molar absorptivity of the porphyrins in the absence and presence 2.0 µM each in the solution.
of albumin, and apparent molar absorptivity of the porphyrins
(εa = Aobs/[porphyrin]) (in M−1 cm−1), respectively. For ΔG° calcu- 2.10. Molecular docking procedure
lation, R and T are the gas constant (1.987577 kcal K−1 mol−1) The crystallographic structure of HSA was obtained from
and temperature (298.15 K), respectively. Protein Data Bank, with access code 1N5U.32 The chemical
structure of the synthetic porphyrins H2TTP, PtbpyTTP, and
2.8. HSA-binding assays by steady-state fluorescence PdbpyTTP was built and minimized in terms of energy by
emission analysis Density Functional Theory (DFT), available in the Spartan’18
Additional binding parameters between HSA and each por- software (Wavefunction, Inc., Irvine, CA, USA).33 Molecular
phyrin (H2TTP, PtbpyTTP e PdbpyTTP) were obtained by docking calculations were performed with GOLD 2020.2 soft-

1648 | Dalton Trans., 2022, 51, 1646–1657 This journal is © The Royal Society of Chemistry 2022
View Article Online

Dalton Transactions Paper

ware (Cambridge Crystallographic Data Centre, Cambridge, pyrroline n-oxide, Oakwood Chemical®, USA). Aliquots were pre-
CB2 1EZ, UK).34 Hydrogen atoms were added to the albumin pared using 14 of the total volume containing a DMSO solution
following tautomeric states and ionization data inferred by with porphyrin at 57 μM, 12 of the total volume containing a
GOLD 2020.2 software at pH 7.4. A 10 Å radius around each of DMSO solution with the spin trap and 14 of the completed volume
the three main binding pockets (subdomains IIA, IIIA and IB – with the aqueous protein solution at concentration of 114 µM or
known as sites I, II, and III, respectively)32,35 were defined and just water. Then, the solutions were placed in glass capillaries
explored for in silico calculations. The standard ChemPLP with a volume of 50 μL and later inserted in a quartz tube to
function was used due to the redocking studies and best carry out the EPR measurements.
results obtained in previous work for other porphyrins.36 The
Published on 03 January 2022. Downloaded by Universidad de Granada on 8/22/2023 9:44:23 PM.

figures for the best docking pose were generated with PyMOL 2.12. HSA photooxidation by steady-state fluorescence
Delano Scientific LLC software.37 emission analysis
The photooxidation assays of HSA were conducted by steady-
2.11. EPR analysis
state fluorescence emission at room temperature. Stock solu-
Electron Paramagnetic Resonance (EPR) technique was used tions of HSA (5.0 μM) was prepared in Tris-HCl buffer ( pH 7.4)
for the characterization and identification of the reactive containing H2TTP, PtbpyTTP, and PdbpyTTP (at 50 μM in
oxygen species (ROS) generated under white-light illumination DMSO) and the solutions were irradiated with white-light
of porphyrins H2TTP, PtbpyTTP and PdbpyTTP at concen- source (fluence rate of 50 mW cm−2 and a total light dosage of
tration of 57 μM, mixed with aqueous solutions of HSA at con- 90 J cm−2) in a time span of 30 min. The albumin was excited
centration of 114 μM. at 290 nm and the fluorescence emission intensity was
In order to obtain which reactive oxygen species the por- measured at 334 nm. Also, plots of ln F0/F versus time for HSA
phyrins generate, EPR measurements were performed in gave a straight line from which the photodegradation rate con-
DMSO solutions using the spin trap methodology in a com- stant was calculated.
mercial MiniScope MS400 spectrometer (Magnettech,
Germany) operating in X band (microwave frequency approxi-
mately equal to 9.4 GHz). The experimental parameters used 3. Results and discussion
in the measurements were: 10 mW of microwave power, a
modulation field of 100 kHz with an amplitude of 0.2 mT, a 3.1. General characterization
field centered on 337 mT with an application interval of The free-base tetra-(2-thienyl)porphyrin H2TTP was previously
10 mT, a scan time of 60 s and 4096 spots. All EPR spectra were described and fully characterized by Momo and co-workers.38
measured at room temperature. The spin traps used in this The tetra-cationic porphyrin synthesis was carried out in dry
experiment were TEMP (2,2,6,6-tetramethylpiperidinol, Sigma- DMF with a little excess of cis-[M(bpy)Cl2] complex (M = PtII or
Aldrich®, USA), PBN (N-tert-butyl-α-phenylnitrone, Tokyo PdII) per thienyl group at 50 °C for 48 h and under an argon
Chemical Industry Company, Japan) and DMPO (5,5-dimethyl-1- atmosphere, according to the literature (Scheme 1).39 The

Scheme 1 Synthetic pathways for the tetra-cationic thienyl–porphyrins PtbpyTTP and PdbpyTTP.

This journal is © The Royal Society of Chemistry 2022 Dalton Trans., 2022, 51, 1646–1657 | 1649
View Article Online

Paper Dalton Transactions

obtained derivatives exhibited dark-brown aspect and yields of (see the ESI – Fig. S1 and Tables S1 and S2†). Due to a very
80–90%. These structures were characterized and confirmed high number of peaks in NMR and vibrational spectra, which
by CHN% analysis, molar conductivity, and cyclic voltammetry can lead to a complicated and erroneous assignment, the
NMR and FT-IR data of the derivatives PtbpyTTP and
PdbpyTTP were not presented.

3.2. Photophysical properties


The UV-Vis spectra of the porphyrins containing peripheral Pt(II)
Published on 03 January 2022. Downloaded by Universidad de Granada on 8/22/2023 9:44:23 PM.

or Pd(II) complexes in DMSO solution consist of an envelope of


absorption bands in the range of 250–700 nm, arising from the
characteristic absorptions of porphyrin (Soret and Q bands) and
polypyridyl metal derivatives (π → π* and metal-to-ligand charge
transfer transitions, MLCT) entities (Fig. 1a and Table 1).
Porphyrins generally present two fluorescence emission
transitions in the excited state (Fig. 1b). The different values
for the fluorescence quantum yields (Φf ) may be a clue con-
cerning the electronic coupling between the [M(bpy)Cl]+ unit
and the porphyrin ring (Table 1), and the decrease in the fluo-
rescence can be associated with an increase of non-radiative
processes.40 Time-resolved fluorescence was measured by excit-
ing the porphyrins at 455 nm with a NanoLED. The fluo-
rescence lifetimes (τf; Table 1) were obtained through a mono-
exponential fit (I = I0 e−t/τ), and the best fits are presented in
the ESI (Fig. S2–S4†). In this way, it was observed that both
free-base porphyrins without (H2TTP) or with coordination
complexes (PtbpyTTP and PdbpyTTP) presented the lower fluo-
rescence quantum yield (2.0 to 6.0%) and the shorter fluo-
rescence lifetime (2.0 to 2.20 ns range), in agreement to the
literature.40

3.3. Photobiological parameters


Table 2 summarizes the 1O2 generation, photostability, photo-
bleaching, and partition coefficients values. In general, both por-
phyrins did not tend to form aggregates in solution (in between
0.5 to 60 µM), being stable for a few days in DMSO solution and
remained photostable at least over a period of 30 min under
white-light LED irradiation (see ESI – Fig. S5–S13†).
For 1O2 quantum yields, the porphyrins H2TTP, PtbpyTTP,
Fig. 1 (a) Normalized UV-Vis absorption spectra of porphyrins H2TTP,
PtbpyTTP and PdbpyTTP, in DMSO solution and (b) normalized steady-
and PdbpyTTP are capable of producing singlet oxygen species
state fluorescence emission spectra of porphyrins H2TTP, PtbpyTTP and (Table 2). Compared to the standard meso-tetra( phenyl)por-
PdbpyTTP, in DMSO solution, at λexc = 522 nm. phyrin (TPP), the presence of thienyl units and peripheral Pt

Table 1 Photophysical parameters of thienyl–porphyrins in DMSO solution

Porphyrin

H2TTP PtbpyTTP PdbpyTTP

λ, nm 426 (5.54); 522 (4.24); 560 (3.99); 280 (5.15); 326 (4.89); 425 (5.52); 522 (4.21); 560 313 (4.88); 426 (5.49); 522 (4.25); 561
(log ε)a 596 (3.82); 660 (3.63) (3.96); 597 (3.78); 660 (3.63) (4.02); 597 (3.88); 659 (3.75)
λEm, nm 667, 730 (6.0) 618, 685 (2.0) 618, 678 (3.0)
(QY, %)b
τf (ns)c 2.02 2.15 2.20
kr (107 s−1)d 2.90 0.95 1.35
knr (107 s−1)d 46.5 45.6 44.0
a
In DMSO solution. b In DMSO solution (excited at 522 nm and using TPP in DCM as standard – Φf = 0.15). c In DMF solution (using NanoLED at
455 nm). d Using eqn (2) and (3).

1650 | Dalton Trans., 2022, 51, 1646–1657 This journal is © The Royal Society of Chemistry 2022
View Article Online

Dalton Transactions Paper

Table 2 Photobiological data of thienyl–porphyrins (II) or Pd(II) complexes can interfere in the intersystem crossing
by means of spin–orbit coupling, mainly due to the presence
Porphyrin of chemical groups with high electronic density.40 In the
H2TTP PtbpyTTP PdbpyTTP photobleaching study, it is possible to notice that the deriva-
tives PtbpyTTP and PdbpyTTP are less susceptible to photooxi-
ΦΔa 0.33 0.40 0.38
kpo (M−1 s−1)b 0.0014 0.0017 0.0016 dative processes, a fact that is reflected in the ΦPB values (see
kpb (s−1)c 0.87 0.78 0.84 Table 2 and ESI – Fig. S14–S16†). In addition, the log POW was
ΦPB (%)d 17.0 2.45 2.30 measured and the values found for the cationic derivatives
log POWe +2.95 +1.56 +1.51
containing the Pt(II) and Pd(II) polypyridyl complexes are in
Published on 03 January 2022. Downloaded by Universidad de Granada on 8/22/2023 9:44:23 PM.

a
Singlet oxygen generation, in DMSO solution (using TPP in DMF as accordance with the literature,18,28 with neutral tetra-thienyl
standard – Φf = 0.66). b DPBF photooxidation constant. derivative H2TTP showing a more hydrophobic character when
c
Photobleaching constant determined by photostability assays (white-
LED with fluence rate of 50 mW cm−2 and a total light dosage of 90 J compare to the tetra-cationic derivatives PtbpyTTP and
cm−2 for 30 min). d Photobleaching quantum yield determined by PdbpyTTP (Table 2).
photostability assays (white-LED with fluence rate of 50 mW cm−2 and
a total light dosage of 90 J cm−2 for 30 min). e Partition coefficients in
octanol/water. 3.4. HSA-binding study
The porphyrins under study were also evaluated according to
their interaction with albumin (HSA), due to the importance of
this protein in transporting drugs in the human
bloodstream.30,31 The interactive parameters are presented in
Table 3 The HSA-binding parameters for the interactions with thienyl–
porphyrins in DMSO(5%)/Tris-HCl pH 7.4 buffer mixture solution by Table 3 and the entire set of spectra are listed in the ESI
UV-Vis absorption analysis (Fig. S17 and S18†).
Firstly, by UV-Vis analysis, there is a hyperchromic profile
Porphyrin of the spectra with no significant bathochromic shift. The
H2TTP PtbpyTTP PdbpyTTP
intrinsic binding constant (Kb ≈ 103 M−1) and Gibb’s free-
energy (ΔG° < 0) values indicated that the porphyrins interact
a
H (%) 41.0 35.5 58.5 weakly to moderately and spontaneously with albumin, mainly
Δλ (nm)b 0.0 0.0 0.0
Kb (×103 M−1)c 1.57 ± 0.09 0.63 ± 0.02 1.62 ± 0.07 by non-covalent forces, probably by van der Waals and
ΔG° (kJ mol−1)d −18.20 −15.95 −18.30 H-bonding (Table 3). As an example, the UV-Vis HSA titration
of compound PtbpyTTP is presented in the Fig. 2.
a
Hyperchromicity (H%) = Afinal − Ainitial/Ainitial × 100. b Red shift.
c
Intrinsic binding constant determined by Benesi–Hidelbrand The additional binding parameters obtained by steady-state
equation (6). d Gibb’s free-energy determined by eqn (7). fluorescence emission corroborated with the trend described

Fig. 2 UV–Vis absorption spectra of porphyrin PtbpyTTP with increase HSA concentrations (0 to 60 µM), in a DMSO(5%)/Tris-HCl buffer ( pH 7.4)
solution. The insert shows the plot of [HSA]/(εa − εf ) versus [HSA].

This journal is © The Royal Society of Chemistry 2022 Dalton Trans., 2022, 51, 1646–1657 | 1651
View Article Online

Paper Dalton Transactions

Table 4 The HSA-binding parameters for the interactions with porphyrins H2TTP, PtbpyTTP, and PdbpyTTP, in DMSO(5%)/Tris-HCl pH 7.4 buffer
mixture solution by fluorescence emission analysis

Porphyrin T (K) Qa (%) KSV b (×104 M−1) kq c (×1012 M−1 s−1) Ka d (×103 M−1) ΔH° e (kJ mol−1) ΔS° e (kJ mol−1 K−1) ΔG° f (kJ mol−1)

H2TTP 298.15 51.30 1.69 ± 0.02 2.98 ± 0.35 4.46 ± 0.13 −20.80 ± 0.39 +0.070 ± 0.01 −20.85
310.15 1.40 ± 0.02 2.47 ± 0.35 4.38 ± 0.17 −21.60 ± 0.39 +0.069 ± 0.01 −21.65
318.15 1.36 ± 0.01 2.40 ± 0.17 4.27 ± 0.18 −22.10 ± 0.39 +0.069 ± 0.01 −22.15
PtbpyTPP 298.15 80.30 4.62 ± 0.04 8.15 ± 0.70 6.66 ± 0.25 −22.70 ± 0.70 +0.073 ± 0.02 −21.85
310.15 5.73 ± 0.06 10.10 ± 0.10 5.88 ± 0.14 −22.40 ± 0.70 +0.072 ± 0.02 −22.40
318.15 6.50 ± 0.07 11.5 ± 0.12 5.58 ± 0.13 −22.25 ± 0.70 +0.072 ± 0.02 −22.85
−21.70 ± 0.12 −20.40
Published on 03 January 2022. Downloaded by Universidad de Granada on 8/22/2023 9:44:23 PM.

PdbpyTTP 298.15 63.50 1.66 ± 0.02 2.93 ± 0.35 3.67 ± 0.18 +0.068 ± 0.04
310.15 2.32 ± 0.03 4.09 ± 0.53 6.35 ± 0.21 −23.15 ± 0.12 +0.073 ± 0.04 −22.60
318.15 3.25 ± 0.05 5.73 ± 0.88 5.78 ± 0.20 −22.90 ± 0.12 +0.072 ± 0.04 −22.95
a
Quenching (Q%) = F − F0/F0 × 100. b Stern–Volmer quenching constant calculated by eqn (8). c Bimolecular quenching rate constant calculated
by eqn (8). d Modified Stern–Volmer binding constant calculated by eqn (9). e Enthalpy and entropy changes by Van’t Hoff plot calculated by
eqn (10). f Gibbs free-energy values calculated by eqn (7).

above, as well as indicated that the interaction is enthalpically due to the bimolecular quenching rate kq values are about
and entropically driven (Table 4). A ground-state association three orders of magnitude higher than the diffusion rate con-
(static fluorescence quenching mechanism) was also detected stant in water (kdiff ≈ 7.40 × 109 M−1 s−1, at 305 K according to
Smoluchowski–Stokes–Einstein).41 For the metalloporphyrins
the increase in the KSV values with the increasing of tempera-
ture, suggest the possibility to also occur a dynamic process,
probably due to the high steric volume of these compounds.
All results are listed in the Table 4, as an example, the steady-
state fluorescence emission spectra and modified Stern–
Volmer plots for PtbpyTTP porphyrin at three different temp-
eratures are represented in the Fig. 3. All selected spectra and
graphs are listed in the ESI file (Fig. S19–S31†).
In order to identify the main fluorescence quenching
mechanism (static or dynamic), time-resolved fluorescence
decays were obtained for HSA without and in the presence
of H2TTP, PtbpyTTP, and PdbpyTTP in DMSO(5%)/Tris-HCl
pH 7.4 buffer mixture solution and lifetime values are pre-
sented in the Table 5. There is a significant decrease in the
τf value of HSA upon porphyrin addition, indicating that at
high porphyrin concentration a combined static and
dynamic fluorescence quenching mechanism is feasible for
the interaction HSA:H2TTP/PtbpyTTP/PdbpyTTP. All fluo-
rescence decays plots are listed in the ESI (Fig. S32–S35†).
Finally, the number of binding sites (n), calculated by
double-logarithmic equation,25 are between 0.85 and 1.45,
indicating that the proportion HSA : porphyrin is mainly
1 : 1, thus one porphyrin interact with only one single
albumin compound.

Table 5 Time-resolved fluorescence parameters and molecular


docking score (dimensionless) values for the interaction
HSA : porphyrins in DMSO(5%)/Tris-HCl buffer ( pH 7.4) solution

Time-resolved
fluorescence Molecular docking
Fig. 3 (a) Steady-state fluorescence emission spectra for HSA without
and upon successive additions of PtbpyTTP, in a DMSO(5%)/Tris-HCl Compound τf (ns) χ2 Site I Site II Site III
buffer ( pH 7.4) solution at 298.15 K. The concentration of compounds HSA 5.06 ± 0.001 1.04780 — — —
ranged from 0 to 60 μM. Insert graph shows the plot of F0/F versus [ por- HSA:H2TTP 3.89 ± 0.002 1.04934 53.2 4.09 80.9
phyrin] and (b) modified Stern–Volmer plots for HSA:PtbpyTTP adduct HSA:PtbpyTTP 3.74 ± 0.009 1.01757 36.2 1.27 66.8
at three different temperatures. HSA:PdbpyTTP 3.71 ± 0.001 1.04731 42.1 1.35 67.1

1652 | Dalton Trans., 2022, 51, 1646–1657 This journal is © The Royal Society of Chemistry 2022
View Article Online

Dalton Transactions Paper

To offer an atomic-level explanation on the binding capacity Waals interactions were detected as the main intermolecular
of each porphyrin into HSA pocket, in silico studies via mole- forces responsible for the association HSA : porphyrins, being
cular docking calculations were performed (Fig. 4). In this in agreement with the experimental data.
way, Table 6 shows the docking score value (dimensionless),
suggesting that the main binding site is subdomain IB (an 3.5. EPR analysis
external binding region), probably due to the high molecular ROS generation was identified and quantified via EPR tech-
surface area of the porphyrins, which impacts the fit capacity nique by spin trap methodology for a system upon white-light
of the compound. Additionally, hydrogen bonding and van der illumination (see Experimental section). To evaluate the gene-
Published on 03 January 2022. Downloaded by Universidad de Granada on 8/22/2023 9:44:23 PM.

Fig. 4 Best docking poses for the interaction (a) HSA:H2TTP, (b) HSA:PtbpyTTP and (c) HSA:PdbpyTTP, into site III. Selected amino acid residue,
H2TTP, PdbpyTTP and PdbpyTTP are in stick representation in cyan, purple, brown and beige, respectively. Elements’ color: hydrogen: white;
oxygen: red; nitrogen: dark blue, sulfur: yellow, chloro: green, Pt(II): gray and Pd(II): dark green.

This journal is © The Royal Society of Chemistry 2022 Dalton Trans., 2022, 51, 1646–1657 | 1653
View Article Online

Paper Dalton Transactions

Table 6 Molecular docking results for the interaction between HSA: the kinetics of spin adduct generation, it can be noted that all
H2TTP, HSA:PtbpyTTP, and HSA:PdbpyTTP in the site III porphyrins show similar behavior, with an accentuated pro-
duction of singlet oxygen in the first minutes of irradiation
Code Amino acid residues Interaction Distance (Å)
(<5 min). After this initial irradiation time, the radical concen-
H2TTP Arg-117 Hydrogen bonding 3.20 tration slowly decreased, due to the degradation of the spin
Phe-134 van der Waals 1.70
adducts caused by the high concentration of 1O2 generated
Tyr-138 van der Waals 2.90
Ile-142 van der Waals 3.20 during the photo-induced process.22,42 In the presence of the
Phe-149 van der Waals 2.40 HSA protein, it is possible to observe that the production of
Leu-154 van der Waals 3.50
the spin adducts is reduced in relation to the HSA-free por-
Published on 03 January 2022. Downloaded by Universidad de Granada on 8/22/2023 9:44:23 PM.

Phe-157 van der Waals 3.50


Ala-158 van der Waals 2.00 phyrins (Fig. 5). According to the observed results, it is poss-
Tyr-161 Hydrogen bonding 2.40 ible to claim that the derivatives generate 1O2 and can promote
Phe-165 van der Waals 2.10
oxidation in the HSA protein (Table 7).
Arg-186 van der Waals 2.70
PtbpyTTP Arg-114 van der Waals 3.40 To identify the ROS formed by type I mechanism, the PBN
Arg-117 van der Waals 2.30 spin trap (0.1 M) was used and the analysis revealed the for-
Tyr-138 van der Waals 1.70
mation of a methyl radical (•CH3, aN = 1.49 mT and aH(β) =
Ile-142 van der Waals 1.40
His-146 van der Waals 2.90 0.25 mT) (see ESI – Fig. S36†). This radical is characteristic of
Phe-157 van der Waals 3.60 the interaction of the hydroxyl radical (•OH) with the DMSO
Tyr-161 Hydrogen bonding 3.00
molecule.43–45 However, the concentration of •OH is much less
Arg-186 van der Waals 1.50
Lys-190 Hydrogen bonding 2.00 in relation to the formation of 1O2, expected behavior for this
PdbpyTTP Arg-114 van der Waals 3.60 type of photosensitizer, where the dominant ROS formation
Leu-115 van der Waals 3.20
mechanism is due to type II process (Fig. 6). All porphyrins
Arg-117 van der Waals 3.50
Tyr-138 van der Waals 3.00 showed the formation of the same adducts, with the difference
Ile-142 van der Waals 3.00 that the presence of the HSA protein contributes to a slight
His-146 van der Waals 3.50
increase in the concentration of PBN/•CH3 formed during
Tyr-161 Hydrogen bonding 2.20
Arg-186 Hydrogen bonding 3.50 white-light illumination. Even so, it is reckless to conclude
Lys-190 Hydrogen bonding 2.00 that protein binding improves the generation of ROS by the
type I mechanism.
Complementarily, the DMPO spin trap (0.3 M) was used
ration of 1O2 species, the hydroxy-TEMP spin trap in water (1.0 and added the formation of the superoxide radical (O2•−, aN =
M) was used. The concentration of spin adducts generated 1.33 mT, aH(β) = 1.04 mT and aH(γ) = 0.14 mT) indicating elec-
upon white-light illumination was determined by double inte- tron transfer between the porphyrins and molecular oxygen
gration of the EPR signal (Fig. 5) and compared with the signal present in the medium.46–49 All porphyrins, with or without
intensity of aqueous TEMPOL solution (1.0 mM). Considering HSA protein, exhibited a similar line shape in the EPR spectra
(see ESI – Fig. S37 and S38†).
Finally, Fig. 7 depicts the steady-state fluorescence emission
quenching of HSA in the presence of the studied tetra-cationic
porphyrins upon white-light irradiation conditions for 30 min
(HSA photooxidation assays) used as example compound
PtbpyTTP. The data show that under irradiation, tetra-cationic
porphyrins PtbpyTTP and PdbpyTTP present a similar behav-
ior in the photodegradation constants (kpd) (Table 8). All fluo-
rescence emission spectra are presented in the ESI – Fig. S39
and S40.†

Table 7 Spin adducts generation reaction constant formed by singlet


oxygen (1O2) capture by TEMP generated in the white-light irradiation
conditions using H2TTP, PtbpyTTP, and PdbpyTTP, without and in the
presence of HSA

Porphyrin

H2TTP PtbpyTTP PdbpyTTP


Fig. 5 Concentration of spin adduct formed by singlet oxygen (1O2) Without HSA −1
ks (min ) 0 to 5 min 11.0 10.0 13.0
capture by TEMP as a function of irradiation time for H2TTP, PtbpyTTP and ks (min−1) 0 to 2 min 13.0 12.0 16.0
PdbpyTTP porphyrins, in the absence (solid black, blue and red lines) and With HSA ks (min−1) 0 to 5 min 6.0 7.0 7.0
in the presence of HSA (dashed orange, pink and light-blue lines). ks (min−1) 0 to 2 min 10.0 13.0 11.0

1654 | Dalton Trans., 2022, 51, 1646–1657 This journal is © The Royal Society of Chemistry 2022
View Article Online

Dalton Transactions Paper


Published on 03 January 2022. Downloaded by Universidad de Granada on 8/22/2023 9:44:23 PM.

Fig. 6 Comparison of concentrations of PBN/•CH, PBN/•CH3 and PBN* spin adducts, formed by irradiating porphyrins H2TTP, PtbpyTTP and
PdbpyTTP dissolved in DMSO. In (a) a comparison of the concentration of spin adducts formed for porphyrins without HSA protein is presented. In
the other panels, comparisons for porphyrins (b) H2TTP, (c) PtbpyTTP and (d) PdbpyTTP as a function of the presence or absence of HSA albumin
are presented.

Table 8 Photooxidation rate constants in the white-light irradiation


conditions using H2TTP, PtbpyTTP, and PdbpyTTP, in the presence of
HSA, by steady-state fluorescence emission analysis

Porphyrin

H2TTP PtbpyTTP PdbpyTTP


a
Q (%) 28.6 35.6 34.2
kpd (min−1)b 9.30 16.0 15.0
a
Quenching = F − F0/F0 × 100%. b First-order kinetics profile.

4. Conclusions
In summary, the results indicated that meso-tetra-cationic(2-
thienyl)porphyrins PtbpyTTP and PdbpyTTP are interesting
scaffolds for the design of novel photosensitizer, presenting
relevant properties for future application as PDT and aPDT
agents. The insertion of the peripheral-coordinated Pt(II) or
Pd(II) complexes resulted in increased interaction with HSA
Fig. 7 HSA photooxidation assays (DMSO(5%)/Tris-HCl buffer pH = 7.4
at 298.15 K) in the presence of porphyrin PtbpyTTP using an excitation
when compared to the non-cationic porphyrin H2TTP. These
wavelength of 290 nm, under white-light irradiation conditions. Inset novel meso-tetra-cationic(2-thyenil)porphyrin derivatives inter-
plot: first-order kinetic graphical profile. act weakly to moderately and spontaneously with human

This journal is © The Royal Society of Chemistry 2022 Dalton Trans., 2022, 51, 1646–1657 | 1655
View Article Online

Paper Dalton Transactions

serum albumin, mainly by non-covalent forces, probably by K. T. Oliveira, J. M. S. Lopes, N. M. B. Neto, W. C. Moreira,
van der Waals and H-bonding forces, showing good binding L. R. Dinelli and A. A. Batista, Inorg. Chem., 2019, 58, 1030–
parameters to be feasible their biodistribution in the human 1039.
bloodstream. Additionally, there is a great capacity of these 9 V. A. Oliveira, B. A. Iglesias, B. L. Auras, A. Neves and
porphyrins to generate ROS upon white-light irradiation con- H. Terenzi, Dalton Trans., 2017, 46, 1660–1669.
ditions, including protein photodamage. Overall, porphyrins 10 G. Basso, J. F. Cargnelutti, A. L. Oliveira, T. V. Acunha,
PtbpyTTP and PdbpyTTP can be considered as potential and R. Weiblen, E. F. Flores and B. A. Iglesias, J. Porphyrins
promising candidates as bio-supramolecular compounds in Phthalocyanines, 2019, 23, 1041–1046.
the photobiological applications. 11 G. K. Couto, B. S. Pacheco, V. M. Borba, J. C. R. Junior,
Published on 03 January 2022. Downloaded by Universidad de Granada on 8/22/2023 9:44:23 PM.

T. L. Oliveira, N. V. Segatto, F. K. Seixas, T. V. Acunha,


B. A. Iglesias and T. Collares, J. Photochem. Photobiol., B,
Author contributions 2020, 202, 111725–111737.
12 T. T. Tasso, T. M. Tsubone, M. S. Baptista, L. M. Mattiazzi,
K. T. Oliveira and B. A. Iglesias idealized the work. B. A. T. V. Acunha and B. A. Iglesias, Dalton Trans., 2017, 46,
Iglesias conducted the characterization, photophysical and 11037–11045.
photobiological assays. H. F. V. Victória, J. R. de Toledo and 13 A. Naik, R. Rubbiani, G. Gasser and B. Spingler, Angew.
K. Krambrock conducted the EPR experiments. O. A. Chaves Chem., Int. Ed., 2014, 53, 6938–6941.
conducted the molecular docking calculations. K. T. Oliveira, 14 L. Schneider, M. Kalt, M. Larocca, V. Babu and B. Spingler,
B. A. Iglesias, O. A. Chaves and K. Krambrock wrote and Inorg. Chem., 2021, 60, 9416–9426.
correct the manuscript. 15 C. H. da Silveira, V. Vieceli, D. J. Clerici, R. C. V. Santos and
B. A. Iglesias, Photodiagn. Photodyn. Ther., 2020, 31,
101920.
Conflicts of interest 16 G. G. Rossi, K. B. Guterres, C. H. da Silveira, K. S. Moreira,
T. A. L. Burgo, B. A. Iglesias and M. M. A. de Campos,
The authors declare that they do not have conflict of interest.
Microb. Pathog., 2020, 148, 104455.
17 D. C. Jornada, R. Q. Garcia, C. H. da Silveira, L. Misoguti,
C. R. Mendonça, R. C. V. Santos, L. De Boni and
Acknowledgements B. A. Iglesias, Photodiagn. Photodyn. Ther., 2021, 35,
This research was supported by the Conselho Nacional de 102459.
Desenvolvimento Científico e Tecnológico (CNPq/grants 18 S. C. Pinto, T. V. Acunha, J. M. Santurio, L. B. Denardi and
409150/2018–5 and 304711/2018–7), Coordenação de B. A. Iglesias, Photodiagn. Photodyn. Ther., 2021, 36,
Aperfeiçoamento de Pessoal de Nível Superior (CAPES – 102550.
Finance Code 001), Fundação de Amparo à Pesquisa do Estado 19 T. Peters, All About Albumin: Biochemistry, Genetics, and
de São Paulo (FAPESP/grant 2013/07276-1, 2019/27176-8 and Medical Applications, Academic Press, San Diego, CA, 1996,
2020/06874-6) and Fundação de Amparo à Pesquisa do Estado de pp. 79–131.
Minas Gerais (FAPEMIG). O. A. C. thanks Fundação para a 20 P. Ascenzi and M. Fasano, IUBMB Life, 2009, 61, 1118–1122.
Ciência e a Tecnologia (FCT – Portuguese Foundation for Science 21 O. A. Chaves, L. B. Menezes and B. A. Iglesias, J. Mol. Liq.,
and Technology) for the PhD fellowship 2020.07504.BD. 2019, 294, 111581.
22 F. S. Santos, C. H. da Silveira, F. S. Nunes, D. C. Ferreira,
H. F. V. Victória, K. Krambrock, O. A. Chaves,
References F. S. Rodembusch and B. A. Iglesias, Dalton Trans., 2020,
49, 16278–16295.
1 K. Lang, J. Mosinger and D. M. Wagnerová, Coord. Chem. 23 T. Chatterjee, A. Pal, S. Dey, B. K. Chatterjee and
Rev., 2004, 248, 321–350. P. Chakrabarti, PLoS One, 2012, 7, e37468.
2 F. Ricchelli, J. Photochem. Photobiol., B, 1995, 29, 109–118. 24 (a) P. Foletto, F. Correa, L. Dornelles, B. A. Iglesias, C. H. da
3 J. M. Dąbrowski, Adv. Inorg. Chem., 2017, 70, 343–394. Silveira, P. A. Nogara, J. B. T. da Rocha, M. A. F. Faustino
4 B. Halliwell and J. M. C. Gutteridge, Free Radicals in Biology and O. E. D. Rodrigues, Molecules, 2018, 23, 2588;
and Medicine, Oxford Science, Oxford, 3rd edn, 1998, 1–35. (b) L. H. Z. Cocca, T. M. A. Oliveira, F. Gotardo, A. V. Teles,
5 W. E. Savive, Aust. J. Chem., 1971, 24, 1285–1293. R. Menegatti, J. P. Siqueira, C. R. Mendonça,
6 I. E. Borissevitch, T. T. Tominaga and C. C. Schmitt, L. A. M. Bataus, A. O. Ribeiro, T. F. M. Souza,
J. Photochem. Photobiol., A, 1998, 114, 201–207. G. R. L. Souza, P. J. Gonçalves and L. De Boni, J. Photochem.
7 N. S. Lebedeva, E. A. Malkova, T. E. Popova, A. E. Kutyrev, Photobiol., B, 2017, 175, 1–8; (c) R. N. Sampaio,
S. A. Syrbu, E. V. Parfenyuk and A. I. Vyugin, Spectrochim. W. R. Gomes, D. M. S. Araujo, A. E. H. Machado,
Acta, Part A, 2014, 118, 395–398. R. A. Silva, A. Marletta, I. E. Borissevitch, A. S. Ito,
8 T. H. O. Leite, G. Grawe, J. Honorato, B. N. Cunha, L. R. Dinelli, A. A. Batista, S. C. Zílio, P. J. Gonçalves and
O. R. Nascimento, P. S. de Vargas, C. Donatoni, N. M. Barbosa Neto, J. Phys. Chem. A, 2012, 116, 18–26.

1656 | Dalton Trans., 2022, 51, 1646–1657 This journal is © The Royal Society of Chemistry 2022
View Article Online

Dalton Transactions Paper

25 T. V. Acunha, B. M. Rodrigues, J. A. da Silva, 37 W. L. DeLano, PyMOL User’s Guide, DeLano Scientific LLC,
D. D. M. Galindo, O. A. Chaves, V. N. da Rocha, San Carlos, CA, USA, 2002.
P. C. Piquini, M. H. Köhler, L. De Boni and B. A. Iglesias, 38 P. B. Momo, C. Pavani, M. S. Baptista, T. J. Brocksom and
J. Mol. Liq., 2021, 340, 117223. K. T. de Oliveira, Eur. J. Org. Chem., 2014, 4536–4547.
26 R. C. Pivetta, B. L. Auras, B. de Souza, A. Neves, F. S. Nunes, 39 J. A. Naue, S. H. Toma, J. A. Bonacin, K. Araki and
L. H. Z. Cocca, L. D. Boni and B. A. Iglesias, J. Photochem. H. E. Toma, J. Inorg. Biochem., 2009, 103, 182–189.
Photobiol., A, 2017, 332, 306–315. 40 L. H. Z. Cocca, F. Gotardo, L. F. Sciuti, T. V. Acunha,
27 M. Pineiro, A. L. Carvalho, M. M. Pereira, A. M. D. A. Rocha B. A. Iglesias and L. De Boni, Chem. Phys. Lett., 2018, 708,
Gonsalves and L. Arnaut, Chem. – Eur. J., 1998, 4, 2299– 1–10.
Published on 03 January 2022. Downloaded by Universidad de Granada on 8/22/2023 9:44:23 PM.

2307. 41 M. Montalti, A. Credi, L. Prodi and M. T. Gandolfi,


28 F. M. Engelmann, S. V. O. Rocha, H. E. Toma, K. Araki and Handbook of Photochemistry, 3rd edn, CRC Press, Taylor &
M. S. Baptista, Int. J. Pharm., 2017, 329, 12–18. Francis, 2006.
29 H. A. Benesi and J. H. Hildebrand, J. Am. Chem. Soc., 1949, 42 G. A. Sáfar, R. N. Gontijo, C. Fantini, D. C. Martins,
71, 2703–2707. Y. M. Idemori, M. V. Pinheiro and K. Krambrock, J. Phys.
30 O. A. Chaves, T. V. Acunha, B. A. Iglesias, C. S. H. Jesus and Chem. C, 2015, 119, 4344–4350.
C. Serpa, J. Mol. Liq., 2020, 301, 112466. 43 M. G. Steiner and C. F. Babbs, Arch. Biochem. Biophys.,
31 O. A. Chaves, L. B. Menezes and B. A. Iglesias, J. Mol. Liq., 1990, 278, 478–481.
2019, 294, 111581. 44 K. Hensley, M. Aksenova, J. M. Carney, M. Harris and
32 M. Wardell, Z. Wang, J. X. Ho, J. Robert, F. Rucker, J. Ruble D. A. Butterfield, Amyloid, 8-peptide spin trapping II: evi-
and D. C. Carter, Biochem. Biophys. Res. Commun., 2002, dence for NeuroReport, 1995, vol. 6, pp. 493–496.
291, 813–819. 45 A. N. Saprin and L. H. Piette, Arch. Biochem. Biophys., 1977,
33 W. J. Hehre, A guide to molecular mechanics and quantum 180, 480–492.
chemical calculations, Wavefunction, INC., Irvine, USA, 2003. 46 A. E. Alegria and P. Riesz, Photochem. Photobiol., 1988, 48,
34 http://www.ccdc.cam.ac.uk/solutions/csd-discovery/com- 147–152.
ponents/gold/, accessed in October 2021. 47 M. M. Mossoba and P. L. Gutierrez, Biochem. Biophys. Res.
35 O. A. Chaves, M. R. L. Santos, M. C. C. de Oliveira, Commun., 1985, 132, 445–452.
C. M. R. Sant’Anna, R. C. Ferreira, A. Echevarria and 48 P. Bilski, K. Reszka, M. Bilska and C. F. Chignell, J. Am.
J. C. Netto-Ferreira, J. Mol. Liq., 2018, 254, 280–290. Chem. Soc., 1996, 118, 1330–1338.
36 V. Viecelli, O. A. Chaves, K. Araki, P. R. Martins and 49 J. A. Silvester, G. S. Timmins and M. J. Davies, Free Radicals
B. A. Iglesias, J. Braz. Chem. Soc., 2020, 31, 2282–2298. Biol. Med., 1998, 24, 754–766.

This journal is © The Royal Society of Chemistry 2022 Dalton Trans., 2022, 51, 1646–1657 | 1657

You might also like