Journal of Luminescence: Sciencedirect

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 15

Journal of Luminescence 210 (2019) 104–118

Contents lists available at ScienceDirect

Journal of Luminescence
journal homepage: www.elsevier.com/locate/jlumin

Study of energy transfer mechanism in the EuIII and GdIII homobimetallic T


complexes containing the anti-inflammatory drug naproxen and N,N-donors
ligands
Emmanuel M. Gomesa, Douglas F. Francob, Sérgio L. Scarparib, Marcos V. Colaçoc,
Monica S. Ferreirad, Ricardo O. Freired, Lippy F. Marquesa,

a
Grupo de Materiais Inorgânicos Multifuncionais (GMIM), Instituto de Química, Universidade do Estado do Rio de Janeiro, Rio de Janeiro, RJ 20550-013, Brazil
b
Instituto de Química, Universidade Estadual Paulista Júlio de Mesquita Filho, UNESP, Araraquara, SP 14801-970, Brazil
c
Laboratório de Física Médica, Instituto de Física, Universidade do Estado do Rio de Janeiro, Rio de Janeiro, RJ 25550-013, Brazil
d
Pople Computational Chemistry, Departamento de Química, Universidade Federal de Sergipe, São Cristóvão, SE 49100-000, Brazil

ARTICLE INFO ABSTRACT

Keywords: In this work, we present the synthesis, solid state characterization and complete photoluminescence study of new
Photophysical properties important homobimetallic lanthanide complexes containing the non-steroidal anti-inflammatory drug (NSAID)
Luminescence naproxen. The analytical and spectroscopic techniques reveals the formation of eight compounds of general
Lanthanides formula [Ln2(nap)6(H2O)4] (Eu 1 and Gd 2), [Ln2(nap)6(bpy)2] (Eu 3 and Gd 4), [Ln2(nap)6(4,4'-dmbpy)2] (Eu 5
Naproxen
and Gd 6) and [Ln2(nap)6(phen)2] (Eu 7 and Gd 8), where: nap = naproxen ligand, bpy =2,2'-bipyridine, 4,4'-
dmbpy = 4,4'-dimethyl-2,2'-bipyridine and phen = 1,10-phenanthroline. Using the RM1 model, the molecular
structures of the EuIII complexes were calculated, with your optimized ground state geometries used to obtain all
details involved in the energy transfer process. From the respective GdIII complexes were obtained the lowest
ligand triplet states, proving that the photoluminescence in the EuIII naproxen complexes is proposed to be a
ligand sensitized luminescence process. On the other hand, the position of the triplet states also explains the non-
effective energy transfer in the TbIII naproxen complexes. The presence of N,N-donors ligands (bpy, 4,4'-dmbpy
and phen) results in an 3–4-fold increase in the quantum efficiency when compared with the EuIII complex
without nitrogen ligands. The high values of emission quantum efficiency (η ~ 70 - 98%) show the EuIII com-
plexes can be potential candidates as emitters in biologic assays.

1. Introduction potential applications in new materials science, especially in the de-


velopment of Light Conversion Molecular Devices (LCMDs) [11,12].
Lanthanides have very unique photophysical properties which make Among the various classes of ligands, the carboxylates form complexes
them a popular choice for several applications, as magnetic [1], cata- readily with lanthanides, and many luminescent lanthanide complexes
lysts [2], solar cells [3], lasers [4], telecommunications [5], Organic from carboxylic acids have been synthesized and characterized, pre-
Light Emitting Diodes (OLEDs) [6] and luminescent probes for bio- senting a wide structural variety, including the metal-organic frame-
analysis and bio-imaging [7,8]. However, lanthanide ions cannot be works [13], homo/heterobimetallic [14,15] and monometallic [16]
excited efficiently by the absorption of light for quantum mechanical compounds. In particular, the naproxen (nap), is a molecule belonging
reasons [9,10] (e.g. the f → f absorption is Laporte forbidden, resulting to non-steroidal anti-inflammatory drug (NSAID) class, also known as
in low absorption coefficients). In order to overcome this drawback, anaprox, naprelan or naprosyn. This NSAID help reduce pain, fever and
suitable chromophores are employed as antennas (or sensitizers) that inflammation, acting mainly in the reduction of prostaglandin, sub-
have the capability to transfer energy to these ions, resulting in highly stance that usually causes inflammation. Taking into account its func-
photoluminescent coordination compounds [11]. In this context, the tionality in the coordination chemistry and in the energy transfer pro-
synthesis and characterization of lanthanide coordination compounds is cess, this molecule can act as an efficient ligand to obtain stable and
an area of current interest due to their intriguing structures and luminescent lanthanide complexes, because it has a carboxylate group


Corresponding author.
E-mail address: lippymarquesuerj@gmail.com (L.F. Marques).

https://doi.org/10.1016/j.jlumin.2019.01.066
Received 12 November 2018; Received in revised form 21 January 2019; Accepted 22 January 2019
Available online 15 February 2019
0022-2313/ © 2019 Elsevier B.V. All rights reserved.
E.M. Gomes, et al. Journal of Luminescence 210 (2019) 104–118

and still the aromatic portion, essential for the observation of the an- following Eq. [28,29]:
tenna effect. This fact becomes evident, since that some examples of the
3 c 3A 0
quantification and identification this NSAID are performed through the
J
= 7 ) 5D 2
luminescence of the naproxen complexes formed [17,18]. In fact, the 4e 2 3 FJ U ( 0 (1)
literature addresses sensing examples of others NSAIDs as ibuprofen where χ is the Lorentz local field correction term, given by χ = n
[19–21], ketoprofen [22], ortofen [22], indomethacin [22] and di- (n + 2)2/9 and 7F J U ( ) 5D0 is a squared reduced matrix element with
2

flunisal [23,24], using the 4–4f transitions of lanthanide ions. value of 0.0032 for the 5D0 → 7F2 transition and 0.0023 for the 5D0 →
However, there are no studies focused on the elucidation of energy 7
F4 one. The refractive index (n) has been assumed equal to 1.5. In this
transfer in the complexes derived from these anti-inflammatories. From work, the 5D0 → 7F6 transition was not observed experimentally; thus,
these photophysical studies we can establish which system is the most the experimental Ω6 parameter could not be estimated. The sponta-
adequate to promote the sensing of these NSAIDs. For example, among neous emission coefficient, A01 = 0.31 × 10-11(n)3(ν01)3, leading to an
the EuIII and TbIII ions, only the first is capable of being sensitized by the estimated value around 50 s-1 for the refractive index (n) defined above.
anti-inflammatory naproxen, contrary to what we observe for ibuprofen In Eq. (1), the A0λ term, where λ = 2 and 4, represents the spontaneous
[25,26]. Thus, due to the great appeal in the construction of new sys- emission coefficients of the 5D0 → 7F2 and 5D0 → 7F4 transitions, which
tems capable of promoting the sensing of bioactive molecules, we can be calculated from 5D0 → 7F1 reference transition (magnetic dipole
propose here, the synthesis and characterization of eight new homo- mechanism), therefore this transition is practically insensitive to che-
bimetallic [Ln2(nap)6(H2O)4] (Eu 1 and Gd 2), [Ln2(nap)6(bpy)2] (Eu 3 mical environment changing, Eq. (2).
and Gd 4), [Ln2(nap)6(4,4'-dmbpy)2] (Eu 5 and Gd 6) and
[Ln2(nap)6(phen)2] (Eu 7 and Gd 8) naproxen compounds (where: bpy v01 S0
A0 = (A01)
= 2,2'- bipyridine, 4,4'-dmbpy = 4,4'-dimethyl-2,2'-bipyridine and phen = v0 S01 (2)
1,10-phenanthroline). As the main purpose, we will investigate the en- 5
Where S01 and S0λ are the areas under the curves of the D0 → F1 and 7
ergy transfer process, for the eight homobimetallic naproxen com- 5
D0 → 7Fλ transitions, with ν01 and ν0λ being their energy barycenters
plexes, and the role of N,N-donors ligands in the emission process. For respectively.
this, the several spectroscopic properties as: energy transfer (WET),
back-transfer (WBT), radiative (Arad) and nonradiative (Anrad) decay
2.2. General synthesis of the homobimetallic naproxen complexes
rates, Ωλ intensity parameters (λ = 2, 4 and 6), quantum efficiency (η)
and quantum yield (q) of EuIII compounds were theoretically modeled
The preparation of all the complexes was carry out by rection be-
using the electronic and spectroscopic semiempirical models and
tween the naproxen (previously deprotonated with NaOH), N, N-donors
compared with those experimental values.
ligands (bpy, 4,4'-dmbpy and phen) and the corresponding LnIII chlorides
(Ln = EuIII and GdIII), in aqueous/ethanolic medium and at room
2. Experimental
temperature (Scheme 1).

2.1. Materials and measurements


2.2.1. Synthesis of the [Ln2(nap)6(H2O)4]complexes
All synthetic procedures were performed in air, and 2,2'-bipyridine A solution of sodium naproxenate (Nanap) was prepared by the
(bpy), 1,10-phenanthroline (phen) and 4,4'-dimethyl-2,2'-bipyridine (4,4'- addition of NaOH aqueous solution (0.40 mL, 1 mol L-1) to an aqueous
dmbpy) were obtained either from Aldrich® or Fluka® and used as re- suspension (30 mL) containing 82 mg (0.40 mmol) of naproxen. To this
ceived. Naproxen (nap) was obtained from Brazilian Foundation known solution was added the respective lanthanide chloride aqueous solu-
as Fundação Presidente Antônio Carlos (FUPAC) with 98% purity. The tions (10 mL, 0.13 mmol) until total precipitation of the complexes. The
EuCl3·6H2O and GdCl3.6H2O salts were prepared by dissolving euro- precipitates were washed three times with hot distilled water, filtered
pium and gadolinium oxides in hydrochloric acid solution and then and dried in a desiccator over anhydrous calcium chloride.
dried. Elemental analyses for C, H and N were carried out using a [Eu2(nap)6(H2O)4] 1: Yield: 88%. Anal. Calc. for C84H86O22Eu2: %C:
Perkin-Elmer 2400CHN analyzer. The metal contents were determined 57.6, %H: 4.95, %Eu:17.3; Found: %C: 57.1, %H: 4.92, %Eu: 17.9.
by complexometric titrations with standard EDTA solution and using [Gd2(nap)6(H2O)4] 2: Yield: 83%. Anal. Calc. for C84H86O22Gd2:
xylenol orange as indicator [27]. FTIR spectra were recorded with a %C: 57.2, %H: 4.92, %Gd: 17.8; Found: %C: 57.9, %H: 4.88, %Gd: 17.3.
Perkin Elmer Spectrum One model, using the ATR mode in the wave-
number range of 4000 - 400 cm-1, with an average of 128 scans and 2.2.2. Synthesis of the [Ln2(nap)6(bpy)2]complexes
4 cm-1 of spectral resolution. Powder diffraction analysis (PXRD) was An ethanolic solution (10 mL) containing 25 mg (0.13 mmol) of 2,2'-
performed using a Panalytical Xpert Pro (the generator was set at 40 kV bipyridine (bpy) was slowly added to an aqueous suspension (30 mL)
and 40 mA), CuKα radiation (λ = 1.54187 Å). The optics is consists in containing 82 mg of naproxen (0.40 mmol), 50 mg of LnCl3·6H2O
primary beam Soller slits (2.29°), fixed divergence slit (0.5°), anti- (0.13 mmol) and NaOH aqueous solution (0.40 mL, 1 mol L-1). The
scatter slit height (1.52 mm), where the polycrystalline compounds mixture was stirred at room temperature for 12 h and a white solid was
were previously grounded in an agate mortar and they were deposited formed. The solid was filtered, washed three times with hot distilled
sample holder plate, with nearly zero-background. The diffraction data water and dried in a desiccator over anhydrous calcium chloride.
of all compounds were collected by overnight scans in the 2θ range of 6- [Eu2(nap)6(bpy)2] 3: Yield: 85%. Anal. Calc. for C104H94O18N4Eu2:
105° with steps of 0.0167° and 4 s rotation time over 360°. Speed scan is %C: 62.7, %H: 4.76, %N: 2.81, %Eu:15.2; Found: %C: 63.1, %H: 4.85,
about 200 s/°. Detector has active angle of 2.122°. The luminescence %N: 2.85%, %Eu: 15.7.
excitation and emission spectra were recorded using a Jobin-Yvon Model [Gd2(nap)6(bpy)2] 4: Yield: 88%. Anal. Calc. for C104H94O18N4Gd2:
Fluorolog FL3-22 spectrophotometer equipped with a R928 Hamamatsu %C: 62.3, %H: 4.73, %N: 2.80, %Gd: 15.7; Found: %C: 62.5, %H: 4.73,
photomultiplier and 450 W xenon lamp as excitation source and the %N: 2.81, %Gd: 15.3.
spectra were corrected with respect to the Xe lamp intensity and
spectrometer response. Measurements of emission decay were per- 2.2.3. Synthesis of the [Ln2(nap)6(4,4'-dmbpy)2]complexes
formed with the same equipment by using a pulsed Xe (3 μs bandwidth) Compounds 5 and 6 were obtained by applying the same synthetic
source. Experimental intensity parameters, Ωλ, for the [Eu2(nap)6 procedure as described for 3 and 4, except 4,4'-dimethyl-2,2'-bipyridine
(H2O)4] 1, [Eu2(nap)6(bpy)2] 3, [Eu2(nap)6(phen)2] 5 and [Eu2(nap)6 (4,4'-dmbpy) (25 mg, 0.13 mmol) was used instead of 2,2'-bipyridine
(4,4′-dmbpy)2] 7 were determined from the emission spectra using the (bpy).

105
E.M. Gomes, et al. Journal of Luminescence 210 (2019) 104–118

Scheme 1. Synthetic route to obtain the lanthanide complexes 1–8.

[Eu2(nap)6(4,4'-dmbpy)2] 5: Yield: 89%. Anal. Calc. for C108H102O18 (criteria to be increased by 100 times), GEO-OK (override interatomic
N4Eu2: %C: 63.3, %H: 5.02, %N: 2.74, %Eu: 14.8; Found: %C: 63.9, % distance and other safety checks), XYZ (cartesian coordinate system was
H: 5.05, %N: 2.76, %Eu: 15.2. used), BFGS (use the BFGS geometry optimizer), GNORM= 0.25 (stop
[Gd2(nap)6(4,4'-dmbpy)2] 6: Yield: 90%. Anal. Calc. for C108H102O18 when gradient norm drops below 0.25) and ALLVEC (print all eigen-
N4Gd2: %C: 63.0, %H: 4.99, %N: 2.72, %Gd: 15.2; Found: %C: 63.4, % vectors).
H: 5.01, %N: 2.75, %Gd: 15.5. For all calculated ground state geometries, we have predicted their
singlet and triplet excited states using configuration interaction single
2.2.4. Synthesis of the [Ln2(nap)6(phen)2]complexes (CIS) based on the intermediate neglect of differential overlap/spec-
Compounds 7 and 8 were obtained by applying the same synthetic troscopic (INDO/S) technique [42,43]. We have used a point charge of
procedure as described for 3 and 4, except 1,10-phenanthroline (phen) +3 e to represent the trivalent lanthanide ion. All calculations were
(25 mg, 0.13 mmol) was used instead of 2,2'-bipyridine (bpy). performed using the software ORCA [44].
[Eu2(nap)6(phen)2] 7: Yield: 91%. Anal. Calc. for C108H94O18N4Eu2:
%C: 63.5, %H: 4.64, %N: 2.75, %Eu: 14.9; Found: %C: 64.0, %H: 4.66, 2.3.2. Intensity parameters calculation
%N: 2.78, %Eu: 15.3. The intensity parameters, Ωλ (λ = 2, 4, and 6), are calculated by
[Gd2(nap)6(phen)2] 8: Yield: 88%. Anal. Calc. for C108H94O18N4Gd2: Judd-Ofelt theory [45,46], implemented in the LUMPAC software [47].
%C: 63.2, %H: 4.62, %N: 2.73, %Gd: 15.3; Found: %C: 63.5, %H: 4.66, According to this theory the intensity parameters are calculated
%N: 2.75, %Gd: 15.1. through Eq. (3):

Note. For all reactions a pH control was maintained around 7.5, since + 1(odd ) t (all)
B tp 2

any over-alkalization can lead to hydrolysis of lanthanide ions. = (2 + 1)


t= 1 p= t
(2t + 1) (3)
In this equation the B tp term is calculated using the Eq. (4):
2.3. Theoretical studies
B tp = B edtp + B dctp (4)
The theoretical calculations followed a well-defined protocol
Where the first term refers only to the forced electric dipole (ed) con-
[30,31] which will be described in detail below.
tribution whereas the second term refers only to the dynamics coupling
(dc) contribution. Both contributions are given by Eqs. (5) and (6),
2.3.1. Ground state geometries and excited state calculations
respectively.
There are two possibilities for the quantum chemical calculation of
the electronic structure of systems containing EuIII ions. The first one is 2 t+1
B edtp = r t, t
the use of first principles methods such as Hartree-Fock or DFT (Density E p
(5)
Functional Theory) making use of effective core potentials (ECP) [32,33]
to treat the EuIII and the second is the use of semiempirical methods, ( + 1)(2 + 3)
1
2
among which we can mention the Sparkle model [34–38] or the B dctp = r (1 ) f C( ) f t
p t, +1
2 +1 (6)
RM1model with explicit treatment of s, p and d type orbitals [39]. Based
on two works published previously by our group [40,41], in which In the Eqs. (5) and (6) the term ΔE corresponds to the difference of
detailed studies of these two approaches were presented, we choose the energy between the ground state barycenters and the first excited state
semiempirical approach and considering that our newest method de- configuration of opposite parity. The radial integrals, r t + 1 , were taken
veloped for trivalent lanthanide ions has proved to be more accurate from reference [48], with an extrapolation for the quantity r 8 . The
and more general in comparison with the Sparkle models [34–38] we value of radial integral obtained for this quantity was 110.0323 a.u. The
chose to proceed with the calculation of the ground state geometry of terms θ(t,λ) are numeric factors associated with each lanthanide ion
the systems using the RM1 model [39]. and are estimated from radial integrals of Hartree-Fock calculations
All calculations were performed using the software MOPAC2016 [49]. The quantity (1 – σλ) is a shielding field due to 5s and 5p filled
[12]. The keywords used were: RM1 (The RM1 Hamiltonian), PRECISE orbitals of lanthanide ions, which have a radial extension larger than

106
E.M. Gomes, et al. Journal of Luminescence 210 (2019) 104–118

those of the 4f orbitals [49]. The term f C ( ) f is a tensor operator of The factor L is the ligand state bandwidth-at-half-maximum (in
rank (λ = 2, 4, and 6) and t, + 1s the Kronecker delta function. As such, cm-1), and Δ is the difference of energy between the donor and acceptor
B dctp , is equal to zero when t is different from the λ + 1. states involved in the energy transfer process.
The parameters pt (t = 1, 3, 5, and 7), given by Eq. (7), are the so- As can be seen in Eqs. (10) and (11), the energy transfer rate de-
called odd-rank ligand field parameters and contains a sum over the pends on the distance between the acceptor and donor states and the
surrounding atoms. energy transfer process. This distance is called RL. To calculate the RL, it
1/2 gj is necessary to estimate the coefficients of the molecular orbitals of the
4
t
p = e2 j (2 j )
t+1 Y pt* ( j , j) atom i (ci), which contribute to the state of the ligand (triplet or
2t + 1 Rtj + 1 (7)
j singlet). Thus, the RL is calculated by:
The Eq. (7) is proposed by the Simple Overlap Model (SOM) and c 2R
i i L, i
each term are detailed on references [50,51]. RL =
c2
i i (14)
The parameter pt (t = 1, 3, 5, and 7), like the parameter pt , also
depends on the coordination geometry and on the chemical environ- The energy back-transfer rates (WBT) are obtained by multiplying
ment around the lanthanide ion, and is given by Eq. (8): the transfer rate (WET) by the Boltzmann factor, exp( / kB T ) , con-
1/2 sidering the room temperature. Δ refers to the energy difference be-
4 j
t
p = Y pt* ( j , j) tween the donor and acceptor levels, and kB is Boltzmann constant.
2t + 1 j
Rj t + 1 (8)
The limitations in the intensity parameters calculation consist in 2.3.4. Emission radiative (Arad) and non-radiative (Anrad) rates calculation
determining the quantities, pt and pt . As a result, it is necessary to use The probability of spontaneous emission, considering the mechan-
the experimental intensity parameters. So, the theoretical intensity isms of magnetic dipole and forced electric dipole, is given by:
parameters were obtained using a new procedure developed by our 64 4 3 n (n2 + 2) 2
group, based on electronic densities and super-displacements for an A (5D0 7
FJ) = Sed + n3Smd
3h (2J + 1) 9 (15)
adjustment of the intensity parameters associated to a single set of
charge factors (g) and polarizabilities (α) [52]. 6
Arad = A (5D0 7
FJ )
(16)
2.3.3. Energy transfer and back-transfer rates J =0

The theoretical model used to quantify the energy transfer rates 5 7


The transitions D0 → FJ (J = 0, 3 and 5) are forbidden by mag-
between the binder part and the lanthanide ion was developed by Malta netic dipole and forced electric dipole mechanisms, that is, their con-
et al. [53]. The calculations were performed using the software tributions are equal to 0.
LUMPAC [47]. According to this model, the energy transfer rates, WET, In the Eq. (15), ν is the energy difference between the states 5D0 and
can be determined from the sum of two terms: 7
FJ (in cm-1), 2J + 1 is the degeneracy of the initial state, h is the Planck
mm
WET = WET em
+ W ET (9) constant, and n is the refractive index of the medium that in the present
study we assumed to be equal to 1.5. Sed and Smd are the contributions
Where the first term corresponds to the energy transfer rate obtained of the mechanism by electric and magnetic dipole, respectively. Sed and
from the multipolar mechanism and can be calculated with Eq. (10): Smd is given by Eqs. (17) and (18) respectively:
mm 2 e 2SL 2
Sed = e 2 5
D0 U ( ) 7F 2
WET = = F J U( ) J J
(2J + 1) G = 2,4,6 (17)
4 e 2SL ed 2
+ F J U( ) J 2
(2J + 1) GRL6 (10) Smd = J L + 2S J 2
4mc 2 (18)
And the second term, that refers to the energy transfer rates ob- The squared matrix elements D0 are equal to 0.0032
5
U ( ) 7F
2
J
tained from the exchange mechanism and are calculated by Eq. (11): (λ = 2), 0.0023 (λ = 4), and 0.0002 (λ = 6), respectively, for EuIII
2 [54].
em 8 e 2 (1 0)
2
2 With respect to the non-radiative emission rate (Anrad), there are no
W ET = F J S J µZ (k ) sm (k )
3 (2J + 1) RL4 m k analytical equations to calculate this quantity. Then, Anrad is de-
(11) termined by the lifetime (τ) of the 5D0 → 7F2 transition using the fol-
lowing equation:
In the Eq. (10) G is the degeneracy of the initial state of the ligand;
ed
are the electric dipole contributions to the intensity parameters, 1
Anrad = Arad
considering only the contributions of the parameters B edtp . The para- (19)
meters are calculated by Eq. (12): The lifetime (τ) is determined experimentally and Arad calculated by
r 2 Eq. (15). The calculations were performed using the software LUMPAC
= ( + 1) 2
3 C( ) 3 2 (1 )2 [47].
(RL + 2) (12)
In the Eqs. (10) and (11), J is the total angular momentum quantum 2.3.5. Emission quantum yield calculation
number of the lanthanide ion and α specifies the spectroscopic term of The emission quantum yield is defined as the ratio between the
the 4f orbitals. The quantities J U ( ) J 2 are reduced matrix ele- emitted and absorbed light intensities and is calculated by Eq. (20):
(λ)
ments of the tensor operators U ; SL is the dipole strength associated
Anrad
with the ϕ → ϕ′ transitions in the ligands. The quantity F corresponds to
5
D0
q=
the temperature-dependent factor and contains a sum of Frank Condon S0 (20)
factors and is calculated by Eq. (13): 5
Where 5D0 is the D0 level population. S0 correspond to the S0 singlet
1 In2
2 level population and φ correspond to absorption rate. The normalized
F= exp In2 population levels, ηj, are obtained from the appropriate rate equations
L L (13) given by:

107
E.M. Gomes, et al. Journal of Luminescence 210 (2019) 104–118

d S1 of the vibrational mode ν(C˭N) [56]. However, this vibrational mode


= (106 + 105 + W(S1 + W(S1 + W(S1 D0) ) S1 appears in a rich region of the FTIR spectrum, where are active several
5 5 5
D4) D1)
dt
+ 10 4 + W(5D4 + W(5D1 + W(5D0 other vibrational modes, as the asymmetric and symmetric stretching
S0 S1) 5D4 S1) 5D1 S1) 5D0 (21)
vibrations of the COO- groups. Thus, in the FTIR spectra of complexes
d T1
3–8, other characteristic bands, such as out-of-plane ρ(CH) pyridine
= (105 + W(T1 5
D4) + W(T1 5
D1) + W(T1 5
D0) ) T1 + 10 8 S1 vibrations [57] (absent in 1 and 2 complexes) are useful to confirm the
dt
+ W(5D4 + W(5D1 + W(5D0 coordination of the corresponding nitrogen ligands to the LnIII centers.
T1) 5D4 T1) 5D1 T1) 5D0 (22)
Such bands appear at 756 cm-1 for ρ(CH)bpy (in complexes 3 and 4),
d 5 822 cm-1 for ρ(CH)dmbpy (in complexes 5 and 6) and 733 cm-1 for
D4
= (106 + W(5D4 S1) + W(5D4 T1) ) 5D4 + W(S1 5
D4 ) S1 ρ(CH)phen (in complexes 7 and 8) (see Table 1 and spectra inserted in
dt
+ W(T1 the square, in Fig. 1b-d).
5
D4) T1 (23)

d 5
3.2. X-ray powder diffraction data
D1
= (106 + W(5D1 S1) + W(5D1 T1) ) 5D1 + 106 5
D4
dt
Despite several attempts to obtain the single crystals of the na-
+ W(S1 5 + W(T1 5
(24)
D1) S1 D1) T1
proxen complexes, we did not achieve success. These attempts included
hydro/solvothermal conditions, recrystallization (complex in the
d 5
D0
= (W(5D0 S1) + W(5D0 T1) + Arad + Anrad ) 5
D0 + 106 5
D1
powder form) in several solvents (ethanol, methanol, acetone, di-
dt methylformamide and dimethylsulfoxide) and solvent exchange in the
+ W(S1 5 + W(T1 5
(25)
D0) S1 D0) T1 reactions. To the best of our knowledge, are known in the literature,
only one work involving crystal structures of lanthanide complexes
5 + 5 + 5 + 7 =1 (26) containing the anti-inflammatory drug naproxen as ligand [58]. Thus,
D4 D1 D0 FJ

polycrystalline X-ray diffraction patterns (PXRD) were obtained for:


S0 + T1 + S1 =1 (27)
[Ln2(nap)6(H2O)4] (Eu 1 and Gd 2), [Ln2(nap)6(bpy)2] (Eu 3 and Gd 4),
5 5
In this study we considered the excited states S1, T1, D4, D1 and [Ln2(nap)6(4,4'-dmbpy)2] (Eu 5 and Gd 6) and [Ln2(nap)6(phen)2] (Eu 7
5
D0 and all calculations were performed using the software LUMPAC and Gd 8) (Fig. 2a-d, respectively). From PXRD patterns, we can note
[47]. that synthesized complexes have different crystallinity degrees: the
congeners compounds 1–2, 5–6 and 7–8 exhibit a higher crystallinity
3. Results and discussion degrees, evidenced by its better defined peaks. On the other hand, for
the compounds [Ln2(nap)6(bpy)2] (3 and 4) the absence of any dif-
3.1. Infrared spectroscopy fraction pattern demonstrates that such compounds have a low crys-
talline nature. These data show that the complexes [Eu2(nap)6(H2O)4] 1
The infrared spectra (ATR mode) of the studied (1–8) compounds and [Gd2(nap)6(H2O)4] 2 form a group isostructural, with the same
and sodium salt (Nanap) are shown in Fig. 1a–d, while the infrared occurring for the compounds [Ln2(nap)6(4,4'-dmbpy)2] (Eu 5 and Gd 6)
spectra of the free ancillary ligands: bpy, 4,4'-dmbpy and phen is shown and [Ln2(nap)6(phen)2] (Eu 7 and Gd 8). In addition, the difference in
in Fig. S1, in Supplementary material. The similarity of the lanthanide reflection intensities among the PXRD patterns was due to the variation
complexes IR spectra indicates the formation of four groups of iso- in polycrystalline samples preferred orientation during the experi-
structural compounds: [Ln2(nap)6(H2O)4] (Eu 1 and Gd 2), mental data collection. Despite of our exhausting efforts, the crystal-
[Ln2(nap)6(bpy)2] (Eu 3 and Gd 4) [Ln2(nap)6(4,4'-dmbpy)2] (Eu 5 and linity degree of the compounds do not allow us determine crystal
Gd 6) and [Ln2(nap)6(phen)2] (Eu 7 and Gd 8), as have been proved by structure by PXRD.
X-ray diffraction analysis results. The carboxylate anion usually co-
ordinates to metal ions in different ways, with such differences being 3.3. Molecular structures of the EuIII complexes by RM1 model
reflected in the relative position of the asymmetric and symmetric
stretching vibrations of the COO- groups. The values of ∆ν = νasym(COO- In the impossibility of obtaining crystal structures, the optimized
) – νsym(COO-) in monodentate complexes are expected to be much ground state geometries of the four EuIII complexes were determined
larger than that in the sodium salt from carboxylic acid. In the other using the RM1 model [59]. It adds up to this, that knowledge of geo-
hand, in bidentate complexes, ∆ν will be significantly smaller than in metry is fundamental to predict spectroscopic properties as well as
the ionic precursor, whereas ∆ν of the bridging complexes is close to important energy transfer mechanism, posteriorly discussed in this
that of the sodium salt [55]. This is an important tool, since lanthanide work. It is worth mentioning that all characterization results discussed
carboxylates usually contain a variety of different coordination modes above are in excellent agreement with the optimized structures, which
as a result of their large coordination numbers. The FTIR spectrum of were further proved with the fitting of the theoretically calculated
the nap ligand in ionic form (Nanap) presents strong bands at 1583 and Judd-Ofelt parameters with those experimentally observed. The opti-
1391 cm-1, which are attributed to νasym(COO-) and νsym(COO-) mized ground state geometries of the [Eu2(nap)6(H2O)4] 1, [Eu2(nap)
stretching modes, respectively, providing ∆ν = 192 cm-1. For the all 6(bpy)2] 3, [Eu2(nap)6(4,4'-dmbpy)2] 5 and [Eu2(nap)6(phen)2] 7 com-
complexes, the existence of two νasym(COO-) and νsym(COO-) vibrations plexes and the general coordination environment of the EuIII ions are
led to the calculation of two values for ∆ν (see Table 1), indicating two depicted in Fig. 3a-d. All the EuIII naproxen complexes are homo-
different coordination modes of the carboxylate groups: chelate and bimetallic (formed by two lanthanide centers), with six monoanionic
bidentate bridging modes. A broad band centered at 3353 cm-1 (ob- nap ligands, what guarantee the charge-neutrality. Each EuIII ion is
served in the FTIR spectra of the [Ln2(nap)6(H2O)4] (Eu 1 and Gd 2) coordinated by seven oxygen atoms from six nap ligands, two oxygen
may be assigned to ν(O – H) stretching vibrations, suggesting the pre- atoms from two water molecules (in the compound 1) and two nitrogen
sence of water molecules in the structure of these compounds, not being atoms from one ancillary ligands (bpy, 4,4’-dmbpy and phen) in the 3, 5
observed for the other complexes. Additionally, the weak bands near and 7 compounds. The coordination geometry for the nine coordinated
3050 and 2960 cm-1 correspond to the ν(C–H) stretching modes of EuIII ions can be described as a distorted tricapped trigonal prism (Fig.
aromatic ring and aliphatic chain of the ligands, respectively. The co- S2 in Supplementary material), typical for this class of complexes.
ordination of the N,N-donors ligands to the lanthanide centers, through Additionally, all carboxylic groups are deprotonated in the complexes,
the pyridine nitrogen atom, can be generally confirmed by the analysis making the naproxenate anion, a powerful ligand in the complexation

108
E.M. Gomes, et al. Journal of Luminescence 210 (2019) 104–118

Fig. 1. FTIR spectra of the Nanap sodium salt and naproxen complexes: (a) [Ln2(nap)6(H2O)4] (Eu 1 and Gd 2), (b) [Ln2(nap)6(bpy)2] (Eu 3 and Gd 4), (c)
[Ln2(nap)6(4,4'-dmbpy)2] (Eu 5 and Gd 6) and (d) [Ln2(nap)6(phen)2] (Eu 7 and Gd 8).

of lanthanide ions, according with Pearson's rules for lanthanide com- one of the oxygen atoms is coordinated to another EuIII ion forming a
pounds. monoatomic bridge or μ-oxo bridge.
In the EuIII naproxen complexes, there are six nap ligands, adopting Some crystal structures of naproxen complexes containing d-block
three coordination modes, as can be seen in Fig. 4. In a coordination metals are reported, as [Cd(naproxen)2(CH3CH2OH)2] [61] where the
mode (a) naproxenate anion acts as simple bidentate chelating ligand CdII ion assumes distorted trigonal prismatic geometry, with two car-
toward one europiumIII center with two oxygen atoms from carboxylate boxylate oxygen atoms coming from the same naproxen ligand and one
group. The carboxylate group can also connect a pair of EuIII ions in a ethanolic oxygen atom occupying the corners of the trigonal face with
syn, syn-η1:η1:μ2 bidentate bridging fashion, as depicted in (b). Finally, the metal ion sitting on a two fold axis. In other example, Totta and
the carboxylate oxygen atoms can act as bidentate chelating, in (c), and, coauthors [62] synthesized and characterized several NiII naproxen

Table 1
IR wavenumbers (in cm-1) and tentative assignment of the most important bands for ligands (Nanap, bpy, phen and 4,4'-dmbpy) and 1–8 complexes.
Compounds νCH(aliph) νCH(arom) ρ(CH) νasym(CO2-) νsym(CO2-) ∆ν/cm-1

Nanap 2950 (w) 3058 (w) – 1583 (s) 1391 (m) 192
bpy – 3051 (m) 756 (vs) – – –
phen – 3058 (w) 735 (s) – – –
4,4'-dmbpy 2919 (w) 3045 (s) 822 (s) – – –
1 2968 (m); 2848 (m) 3054 (w) – 1603 (m); 1541 (m) 1409 (m); 1391 (vs) 194; 150
2 2969 (m); 2848 (m) 3054 (w) – 1603 (m); 1541 (m) 1409 (m); 1391 (vs) 194; 150
3 2967 (w); 2934 (w) 3058 (w) 758 (m) 1599 (vs); 1548 (s) 1406 (vs); 1392 (vs) 193; 156
4 2967 (w); 2934 (w) 3058 (w) 758 (m) 1599 (vs); 1548 (s) 1405 (vs); 1392 (vs) 194; 156
5 2962 (w); 2936 (w) 3058 (w) 823 (m) 1600 (vs); 1550 (s) 1406 (s); 1393 (s) 194; 157
6 2962 (w); 2936 (w) 3058 (w) 823 (m) 1600 (vs); 1550 (s) 1406 (s); 1393 (s) 194; 157
7 2965 (w); 2935 (w) 3057 (w) 733 (m) 1601 (vs); 1548 (s) 1407 (s); 1392 (s) 194; 156
8 2965 (w); 2935 (w) 3057 (w) 733 (m) 1601 (vs); 1548 (s) 1407 (s); 1392 (s) 194; 156

*(Abbreviations: (vs) = very strong; (s) = strong; (m) = medium; sym = symmetric; asym = asymmetric).

109
E.M. Gomes, et al. Journal of Luminescence 210 (2019) 104–118

Fig. 2. X-ray powder diffraction patterns (PXRD) of the complexes (a) [Ln2(nap)6(H2O)4] (Eu 1 and Gd 2), (b) [Ln2(nap)6(bpy)2] (Eu 3 and Gd 4), (c) [Ln2(nap)6(4,4'-
dmbpy)2] (Eu 5 and Gd 6) and (d) [Ln2(nap)6(phen)2] (Eu 7 and Gd 8).

complexes in the absence or presence of the ancillary ligands, as 2,2'- 3.4. Diffuse reflectance (DR) spectroscopy
bipyridine (bpy), 1,10-phenanthroline (phen), 2,2'-bipyridylamine (bi-
pyam), 2,2'-bipyridylketone oxime (Hpko) and pyridine (py), with such The possible ligand-to-metal charge transfer (LMCT) can be in-
complexes presenting activity to inhibit soybean lipoxygenase, being vestigated by comparison between EuIII and GdIII DR spectra, more
more active than the free NSAID naproxen. Despite these examples, precisely by the arithmetic subtraction between these DR spectra, as
works involving the synthesis and solid-state characterization of lan- published by Carlos and co-workers [69]. It is known which, once that
thanide naproxen complexes is still very scarce [63,64]. To the best of the GdIII excited levels lie much higher than the typical energy of the
our knowledge, only three crystalline structures are reported: [Ln2(μ2- ligand triplet states, these disabling any LMCT process [70]. The DR
L)4(L)2(phen)2] (where Ln = Dy, Gd and Er; L = naproxen ligand and spectra of the complexes and resulting arithmetic difference between
phen = 1,10-phenanthroline) [58]. In this work, the authors focused to them is showed in Fig. S3 (Supplementary material), with the spectra
the study of the magnetic properties of the complexes, in a brief recorded at room temperature. In the UV region (200–400 nm), a broad
structural characterization. absorption band is observed, and can be attributed to electronic tran-
As known, monocarboxylate ligands can react with lanthanide ions sitions from the ground-state level (π) S0 to the excited level (π*) S1 of
to form homobimetallic structures [65,66]. In this context, our group nap and nitrogen (bpy, 4,4'-dmbpy and phen) ligands, confirming your
has obtained several crystalline structures of homobimetallic com- presences in the complexes. The spectra also display the 7F0 → 5L6
plexes, such as [Ln2(cin)6(bpy)2] [67] and [Ln2(cin)6(phen)2] [11] (393 nm) and 7F0 → 5D2 (464 nm) peaks attributed to Laporte-for-
(where Ln = EuIII, GdIII and TbIII, cin = hydrocinnamate anion, bpy = bidden 4–4f transitions of the EuIII ions. It is clear the similarity of the
2,2'-bipyridine and phen = 1,10-phenanthroline). Tables S1–S4, in DR spectra, evidencing a similar electronic structure of the complexes,
Supplementary material, shows selected spherical atomic coordinates even in those where the N,N-donor ligands are present. The broad
for coordination polyhedra of the [Eu2(nap)6(H2O)4] 1, [Eu2(nap)6 LMCT bands are consistent with typical values of lanthanide charge-
(bpy)2] 3, [Eu2(nap)6(4,4'-dmbpy)2] 5 and [Eu2(nap)6(phen)2] 7 com- transfer (CT) bands, with widths in the range 5.000–15.000 cm-1, and
plexes, arising from the RM1 model. All the average Eu – O and Eu – N which may vary considerably depending on the electron-donating
predicted values obtained from RM1 model agreeing well with those ability of the ligands [71]. A detailed study of these LMCT states and
obtained from the crystallographic studies for the similar structures their infuence on the energy transfer process is outside the scope of this
reported in the literature [11,68]. article. However, it seems to us the LMCT bands have higher energy

110
E.M. Gomes, et al. Journal of Luminescence 210 (2019) 104–118

Fig. 3. Ground state geometry for the complexes [Eu2(nap)6(H2O)4] 1, [Eu2(nap)6(bpy)2] 3, [Eu2(nap)6(4,4'-dmbpy)2] 5 and [Eu2(nap)6(phen)2] 7, calculated by the
RM1 model and drawn using OLEX [60] program. Color codes: C, light-gray; H, light-blue; O, red; N, blue; and Eu, sparkling blue.

lowest triplet excited state energy of the system was determined from
the corresponding [Gd2(nap)6(H2O)4] 2, [Gd2(nap)6(bpy)2] 4,
[Gd2(nap)6(4,4'-dmbpy)2] 6 and [Gd2(nap)6(phen)2] 8 complexes. It is
important to emphasize that thermal energy facilitate the population of
the higher energy vibrational levels of the S1 state, difficulting the in-
tersystem crossing (S1 → T1), with these phosphorescence spectra being
preferably obtained in nitrogen liquid temperature (77 K). The emission
spectra of the GdIII complexes (Fig. 5a–d), do not present the in-
traconfigurational 4–4f transitions, due to: (i) the heavy atom effect and
the seven unpaired f electrons favor the mixing between singlet and
triplet states, increasing the rate of the intersystem crossing. This
mixing was initially assigned by the inhomogeneous magnetic field
caused by the paramagnetic Gd3+ ion. Later, it was shown that the
magnitude of such a magnetic perturbation is not enough to elucidate
the observed strong singlet-triplet mixing, but should interpreted the-
oretically as stronger singlet-triplet mixing resulting from an exchange
interaction between the ligand and metal electrons [72–74]. (ii) the
lowest excited level (6P7/2) of the GdIII ion has energy higher than the
lowest T1 levels of the organic ligands, which makes the energy transfer
Fig. 4. Coordination modes of the carboxylate groups in the structure of the ligands → GdIII unlikely to occur [72,73]. Still, the emission spectrum
EuIII naproxen complexes.
of the GdIII compounds should preferably be obtained adopting a delay
between excitation and detection (time-resolved spectrum). In this case,
than the triplet state, enabling the transfer of energy to EuIII ions. It is the fluorescence bands decrease very fast as the flash delay is increased
worth mentioning that the LMCT states can also work as deactivation and only the phosphorescence spectrum from the ligands is displayed.
channels if they have energy levels close or lower than the emission That way, in the time-resolved spectra (delay of 0.1 ms), the triplet
levels [69]. excited states (T1) were estimated fitting a tangent at the highest energy
band edge, called of 0-0 phonon transition, with the values of 502.3 nm
3.5. Photoluminescence studies (19,908.4 cm-1) for [Gd2(nap)6(H2O)4] 2, 494.3 nm (20,230.6 cm-1) for
[Gd2(nap)6(bpy)2] 4, 487.2 nm (20,525.4 cm-1) for [Gd2(nap)6(4,4'-
3.5.1. Phosphorescence of the GdIII compounds dmbpy)2] 6 and 494.7 nm (20,211.4 cm-1) for [Gd2(nap)6(phen)2] 8. As
The excited triplet states (T1, 3π,π*) of the ligands play a crucial role can be observed, all these triplet states have energy higher than the
in energy transfer process of the EuIII carboxylate compounds. Thus, the

111
E.M. Gomes, et al. Journal of Luminescence 210 (2019) 104–118

Fig. 5. Time-resolved phosphorescence spectra


of the GdIII complexes, recorded at 77 K, under
excitation at 334 nm for (a) [Gd2(nap)6(H2O)4]
2, 342 nm for (b) [Gd2(nap)6(bpy)2] 4, 337 nm
for (c) [Gd2(nap)6(4,4'-dmbpy)2] 5 and 341 nm
for (d) [Gd2(nap)6(phen)2] 7, using a tangent at
the highest energy band edge for 0-0 phonon
transition estimation.

main emitting state of EuIII (5D0), proving the possibility of in- continuously monitoring the 5D0 → 7F2 (λ = 614 nm) transition. These
tramolecular energy transfer in these complexes. spectra are dominated by broad bands assigned to S0 → S1 (π, π*)
transition of the naproxen and N,N-donors ligands (maximum at
3.5.2. Excitation and emission spectra of the homobimetallic EuIII naproxen 334 nm for 1, 342 nm for 3, 337 nm for 5 and 341 nm for 7). Also, a
complexes series of intra-4f 6 lines ascribed to transitions between the 7F0 level and
The excitation spectra (Fig. 6a-d) of the EuIII naproxen complexes, at the 5G3 (361 nm), 5H4 (376 nm), 5L7 (381 nm), 5L6 (394 nm), 5D2
nitrogen liquid temperature (77 K) and the solid state, were obtained by (464 nm) and 5D1 (524 nm) excited states are detected, remaining

Fig. 6. Excitation spectra of the EuIII complexes. All the spectra were obtained in the solid state, at 77 K, and monitoring the 5D0 → 7F2 (λ = 614 nm) transition.

112
E.M. Gomes, et al. Journal of Luminescence 210 (2019) 104–118

practically invariable in the four EuIII complexes. Analyzing each ex- complex [25]. Fig. 7a-d shows the 5D0 → 7FJ (J = 0, 1, 2, 3 and 4)
citation spectra separately, we can see that the 7F0 → 5L6 transition emission spectra of the EuIII naproxen complexes, at 77 K temperature.
(centered at 394 nm) is more intense in compounds [Eu2(nap)6(H2O)4] In each of these cases, the excitation is at the maximum on the S0 → S1
1 and [Eu2(nap)6(4,4’-dmbpy)2] 5 complexes, indicating that for these absorption band, corresponding to levels centered in the ligands. The
compounds the sensitization via EuIII ion excited states can is more emission profiles were typical of EuIII based luminescence: with a weak
efficient than the direct excitation in the ligand absorption bands. In the transition related to 5D0 → 7F0 transition together with the hy-
other hand, for the [Eu2(nap)6(bpy)2] 3 and [Eu2(nap)6(phen)2] 7 the persensitive transitions (5D0 → 7F2), that dominates the spectra, re-
broad bands attributable to the ligand levels is more intense, being an sulting in the characteristic red emission of EuIII compounds. As can be
indicative that indirect excitation can result in high emission intensity, observed, the emission spectra show different numbers of lines (Starks
for these two last compounds. It is worth mentioning that, comparisons components) for each transition 5D0 → 7FJ (J = 0–4). This can be ex-
of the energy transfer process efficiency involves not only an analysis of plained taking into account two factors: (i) the point symmetry [75]: we
the intensities of the transitions in these excitation spectra, but several expect that the microsymmetry of the EuIII ions in the compounds with
other factors, as triplet level position or presence of coordination water the ancillary ligands (compounds 3, 5 and 7) is slightly different than
molecules. that of the [Eu2(nap)6(H2O)4] 1 hydrated complex. This can be sug-
Another interesting observation in the excitation spectra is that the gested by the different spectral profile of compound 1 when compared
lower edge of the broad bands seem to go towards lower energies. This to nitrogen compounds, which are quite similar to each other. (ii) with
fact can be related to the presence of high energy LMCT states localized the crystallinity degree of the compounds: as shown by X-ray diffraction
between singlet and triplet levels, as discussed in the 3.4 Section results, the hydrated complex has a higher crystalline character, re-
(Diffuse reflectance spectroscopy). sulting in the well defined spectral shape.
The vibronic components can be observed between the 7F0 → 5L6 The relative intensity between the magnetic (5D0 → 7F1) and electric
and 5F0 → 5D2 transitions (observed more clearly in the a and c ex- (5D0 → 7F2) transitions depends on the polarizability of the ligands and
citation spectra). The differences between the bands of the vibronic also can be used as a criterion for the symmetry distortion of the EuIII
components with the 7F0 → 5D2 transition (zero phonon line, ZPL) is ion nearest surroundings. The ratios I(5D0 → 7F2)/I(5D0 → 7F1) are
equal to 1543 cm-1, and related to vibrational mode of νassym(COO−), about 4.85, 5.20, 4.51 and 5.17 for [Eu2(nap)6(H2O)4] 1,
which is consistent with the infrared spectra (Fig. 1). As we can see, the [Eu2(nap)6(bpy)2] 3, [Eu2(nap)6(4,4'-dmbpy)2] 5 and
absorption band related to the S0 →S1 is more intense in the EuIII na- [Eu2(nap)6(phen)2] 7, respectively, suggesting that the arrangement of
proxen complexes containing the N,N-donors ligands, having these li- the effective charges in the surroundings of EuIII ions in all the com-
gands an important role in the light absorption, and consequently in the plexes is noncentrosymmetric, with a low symmetry of the crystal field.
energy transfer process. Thus, is known that the presence of these an- Additionally, for all emission spectra, the 5D0 → 7F2 transition intensity
cillary ligands (as 2,2'-bipyridine and 1,10-phenanthroline) in the struc- is higher than 5D0 → 7F1 ones, indicating that the forced electric dipole
ture of the EuIII complexes, results in an 3–4-fold increase in the and the dynamic coupling mechanisms are dominant in relation to
quantum efficiency when compared with the EuIII complex without of magnetic dipole ones, once the EuIII ions are located in a low symmetry
these ligands. In fact, it was demonstrated by our group a substantial site.
increase in the luminescence intensity in the complexes The presence of the 5D0 → 7F0 transition, which appears only in low
[Eu2(ibf)6(LN)2] (ibf = ibuprofenate anion and LN = 2,2'-bipyridine and symmetries as in the point groups Cnv, Cn or Cs, indicates that all the
1,10-phenanthroline) when compared to EuIII-ibuprofen hydrated complexes present a low symmetry in the coordination geometry, as the

Fig. 7. Emission spectra of the EuIII complexes. All the spectra were obtained in the solid state, at 77 K. In each case, the excitation corresponds to singlet levels
centered in the ligands.

113
E.M. Gomes, et al. Journal of Luminescence 210 (2019) 104–118

Table 2 Table 3
Energy centroid of the 5D0 → 7F0 transition. Lifetime decay values (τ, in ms) for the EuIII naproxen complexes.
5
Complexes D0 →7F0 (nm) 5
D0 →7F0 (cm-1) Complexes lifetime (τ, ms)

[Eu2(nap)6(H2O)4] 1 583.2 17,146.8 [Eu2(nap)6(H2O)4] 1 0.588


[Eu2(nap)6(bpy)2] 3 581.8 17,188.0 [Eu2(nap)6(bpy)2] 3 1.58
[Eu2(nap)6(4,4'-dmbpy)2] 5 582.8 17,158.5 [Eu2(nap)6(4,4'-dmbpy)2] 5 1.82
[Eu2(nap)6(phen)2] 7 583.2 17,146.8 [Eu2(nap)6(phen)2] 7 2.22

selection rules for the electric dipole transition provides. Only one peak states of EuIII ions. In this case, the energy gap between the radiative
(without splitting) is observed for these transitions, indicating that and ground state manifolds of EuIII ion (energy gap between 5D0 and
there is a single europiumIII site present in each complex. This ob- 7
F6) is approximately 12,300 cm-1, matching well the vibrations of O –
servation still suggests that the EuIII ions experience the same crystal- H oscillators (νOH ~ 3300–3500 cm-1), with the quantum number ν = 3,
field strength and occupy sites of same symmetry in the complexes. As which leads to effective quenching of the 5D0 excited state of EuIII ion.
know, the energy of the 5D0 →7F0 transition can be correlated with the The complex [Eu2(nap)6(phen)2] 7 exhibit the highest lifetime value,
covalence degree of the EuIII-ligand bonds [76,77]. Thus, after careful being justified by presence of phen ligand, which promotes a greater
analysis in the 5D0 → 7F0 transition, between 580 and 585 nm, it was structural rigidity when compared to the complexes containing the bpy
concluded that these transitions consists of one single peak, with cen- and 4,4'-dmbpy ligands (complexes 3 and 5, respectively).
troid at 583.2 nm (~17,146.77 cm-1) for 1, 583.2 nm (~17,146.77 cm- A noteworthy fact, in complexes containing bpy and phen ligands is
1
) for 3, 583.2 nm (~17,146.77 cm-1) for 5 and 583.2 nm the dependence of 5D0 lifetime on the excitation wavelength, as re-
(~17,146.77 cm-1) for 7, as can be seen in the Table 2. ported by L. Carlos and coauthors [81]. Therefore, the 5D0 lifetime was
These results indicate the following energy order for the 5D0 → 7F0 also obtained, by the 7F0 → 5L6 (~ 394 nm) excitation for all EuIII
transition: [Eu2(nap)6(bpy)2] > [Eu2(nap)6(4,4'-dmbpy)2] > [Eu2(nap)6 compounds, and the obtained values are: 0.60 ms for 1, 1.66 ms for 3,
(phen)2] ~ [Eu2(nap)6(H2O)4]. Tanner showed that several factors can 2.22 ms for 5 and 1.88 ms for 7.
influence the energy changes of the 5D0 → 7F0 transition [78]. Here, the The luminescence spectroscopy technique, through the excitation/
red shift of this transition (nephelauxetic effect), shows that covalence emission spectra and emission decay curves, is widely used in the de-
degree can is related to the nature of N,N-donor ligand employed. Thus, termination and sensing of various molecules and drugs, as the non-
the higher covalent character is found in [Eu2(nap)6(phen)2] 7 and steroidal anti-inflammatory drugs (NSAID), including the naproxen
[Eu2(nap)6(H2O)4] 1, while that the smaller is attributed to the [17–24]. Thus, in order to understand such sensing process, becomes
[Eu2(nap)6(bpy)2] 3. It is interesting to note the smaller covalent necessary a complete elucidation of the energy transfer mechanism in
character of this last complex when compared to [Eu2(nap)6(4,4'- these EuIII naproxen complexes. Is recognized that the energy level
dmbpy)2] 5, resulting in a Δ(5D0 →7F0) = 29.5 cm-1. This fact can be match between lowest triplet state (T1) of the ligand system and the 5D0
explicated due to the fact of the methyl groups (in 4,4'-dmbpy ligand) to state of the EuIII ion is one the main factors that governs the lumines-
increase the electron density of the ring by inductive effect (+I). This cence efficiency in these complexes. In this context, Latva's empirical
increase in electron density is not, however, uniformly distributed over rule states shows that an optimal ligand → EuIII energy transfer process
all elements of the ring, but would place the largest concentration on requires ΔE(T1 – 5D0) = 2500–4000 cm-1 for the EuIII [82];. Based on
the nitrogen atoms, explaining the higher covalence degree for the Eu – the foregoing, it can be concluded that energy transfer in the EuIII
L bonds in [Eu2(nap)6(4,4'-dmbpy)2] 5. As commented previously, de- complexes is effective, with the values: 2761.6 cm-1 for
tailed studies on the photoluminescence properties for compounds of [Eu2(nap)6(H2O)4] 1; 3042.6 cm-1 for [Eu2(nap)6(bpy)2] 3; 3366.9 cm-1
EuIII with naproxen ligand are inexistent in the literature, making im- for [Eu2(nap)6(4,4'-dmbpy)2] 5 and 3064.6 cm-1 for [Eu2(nap)6(phen)2]
possible a comparison with other similar compounds. However, similar 7. A noteworthy fact is that the respective TbIII naproxen complexes
results were reported for [Eu(Et2NCS2)3bpy] and [Eu(Et2NCS2)3phen] were synthesized and characterized. However, such complexes did not
complexes [79] and for several other complexes in chloroform solutions show luminescence when irradiated with UV light, giving evidence of a
[80]. Recently, our group evidenced the influence of the methyl group non-efficient energy transfer process. Similar to that mentioned for the
position in the photoluminescence properties of some EuIII ibuprofen EuIII complexes, Latva's empirical rule states shows that an optimal li-
complexes, of general formula [Eu2(ibf)6(L)2] (where: L = 4,4'-dmbpy gand → TbIII energy transfer process requires ΔE(T1 –
5
= 4,4'-dimethyl-2,2'-bipyridine and 5,5'-dmbpy = 5,5'-dimethyl-2,2'-bi- D4) = 2500–4500 cm-1. For the TbIII complexes the ΔE(T1 – 5D4) values
pyridine) [26]. Another feature in photoluminescence data of the EuIII are: -455.7 cm-1 for [Tb2(nap)6(H2O)4], -148.5 cm-1 for
-1
complexes is that the emission spectra do not exhibit the broad phos- [Tb2(nap)6(bpy)2], 154.7 cm for [Tb2(nap)6(4,4'-dmbpy)2] and
phorescence band of the ligands, observed in the GdIII complexes (see -142.8 cm-1 for [Tb2(nap)6(phen)2], suggesting high back – energy
Fig. 5a–d), indicating an efficient energy transfer process, ligand → EuIII transfer rates for these complexes. In view of the aforementioned va-
ion. lues, the energy of the lowest triplet level (T1) is located below (ne-
gative values) or practically resonant to 5D4 emitting state. The back-
energy transfer process is one of the major mechanism of luminescence
3.5.3. Lifetime decays and energy transfer mechanism of the quenching in TbIII complexes, corroborating, as we suspected, with a
homobimetallic EuIII naproxen complexes non-efficient energy transfer process. The luminescence intensity in
The photoluminescence decay curves for the EuIII naproxen com- these TbIII complexes (in the room temperature and excitation at ligand
plexes (Fig. S4, in Supplementary material) were obtained in the solid singlet state) is very low, and an increase in this emission intensity can
state, at 297 K, under excitation at singlet states, with emission mon- be achieved by the acquisition of emission spectra at nitrogen liquid
itored at 5D0 → 7F2. The lifetimes values (τ) (Table 3) were obtained temperature (Fig. S5, in Supplementary material). In fact, most of the
using the equation I = I0 exp (-t/τ) and a curve fitting program, in a works involving the quantification of drugs using the luminescence of
monoexponential behavior. Collectively, these data imply the existence lanthanide naproxenates, is performed using the luminescence of EuIII
of a single chemical environment around EuIII ions, in all the com- complexes [17,18]. Here, one point becomes interesting: in our pre-
plexes. As expected, the lowest lifetime values are observed for the vious works, we observed an efficient energy transfer in the TbIII-ibu-
complex that have coordination water molecules (complex 1), once profen complexes [25,26]. Both molecules (ibuprofen and naproxen)
such molecules are typically vibrational deactivators of the excited

114
E.M. Gomes, et al. Journal of Luminescence 210 (2019) 104–118

[Eu2(nap)6(H2O)4] 1 and [Eu2(nap)6(4,4'-dmbpy)2] 5. These discussion


match well with the observations done for the emission spectra of these
EuIII naproxen complexes (see. I(5D0 → 7F2)/I(5D0 → 7F1) values). As
seen in Table 4, there is an excellent agreement between experimental
and calculated Judd-Ofelt intensity parameters, providing strong evi-
dences about the molecular structures of the EuIII homobimetallic na-
proxen complexes. In the [Eu2(nap)6(H2O)4] 1, the Anrad value is
greater than the Arad value, due mainly, the presence of the coordina-
tion water molecules in the compound. On the other hand, when such
water molecules are substituted by the bpy, 4,4'-dmbpy and phen ligands,
there is a decrease in the Anrad rates, increasing the quantum effi-
Fig. 8. Energy diagram showing the possibility (or impossibility) of energy ciencies (~ 2.7 times larger in [Eu2(nap)6(bpy)2] 3 and [Eu2(nap)6(4,4'-
transfer, triplet → EuIII or TbIII in complexes of ibuprofen, naproxen and ke- dmbpy)2] 5, and ~ 3.7 times larger in [Eu2(nap)6(phen)2] 7), when
toprofen. The energy of the triplet states were obtained from the respective GdIII compared with the hydrated complex (η = 25.41%). Another important
complexes: a – [Gd(ibf)3(H2O)2] (T1 = 25,974 cm-1) [25], b – [Gd2(ibf)6(4,4'- quantity to enable the energy transfer rates calculations is the energy
dmbpy)2] (T1 = 23,774 cm-1) [26], c – [Gd2(ibf)6(bpy)2] (T1 = 23,419 cm-1)
position of the triplet states. The corresponding calculated values of
[25], d – [Gd2(ibf)6(5,5'-dmbpy)2] (T1 = 23,201 cm-1) [26], e –
these excited state energies are: 18,883.20 cm-1 for [Eu2(nap)6(H2O)4]
[Gd2(ibf)6(phen)2] (T1 = 22,573 cm-1) [25], f, g h and i (compounds in study)
and j – [Gd(keto)3].1.5H2O ( T1 = 23,800 cm-1) [83]. The circles red and green
1; 20,628.90 cm-1 for [Eu2(nap)6(bpy)2] 3; 20,525.10 cm-1 for
indicates the sensitization of EuIII and TbIII ions, respectively. [Eu2(nap)6(4,4'-dmbpy)2] 5 and 20,630.70 cm-1 for [Eu2(nap)6(phen)2]
7 with these results reflecting excellent agreement between the theo-
retical and experimental data. Table 5 presents the intramolecular en-
belong to the class of non-steroidal anti-inflammatory drugs (NSAIDs),
ergy transfer (WET) and back-transfer (WBT) rates for the
however, only the ibuprofen is capable of promoting an efficient energy
[Eu2(nap)6(H2O)4] 1, [Eu2(nap)6(bpy)2] 3, [Eu2(nap)6(4,4'-dmbpy)4] 5
transfer to TbIII ion. Although the literature shows the application of the
and [Eu2(nap)6(phen)2] 7 systems.
luminescence of EuIII and TbIII ions for the identification/quantification
As we can observe, the T1 states are mainly responsible by the en-
of ibuprofen and naproxen drugs, to the best of our knowledge, there
ergy transfer in all EuIII naproxen complexes, especially through
are no detailed works explaining why only the luminescence of the EuIII
channels T1 → 5D1 and T1 → 5D0, considering WET only. The high va-
ion can be applied in the quantification of naproxen. Thus, the Fig. 8
lues of T1 → 5D1 energy transfer rates (see Table 5) can be justified
shows the energies of the main emitting levels of the EuIII and TbIII ions
considerating the components of the 5DJ multiplet, that are relatively
and the triplet states (obtained from the respective GdIII complexes),
close in energy, and therefore the energy transfer can occur on either
which can be very useful to predict the system ideal for identification/ 5
D1 or 5D0. Considering the above, the highest value of WET (T1 → 5D0)
quantification of these two important NSAIDs (ibuprofen and na-
compared to WET (T1 → 5D0) is due the best resonance condition be-
proxen). In this Figure is also inserted the triplet energy state of the [Gd
tween T1 and 5D0, which in turn also favors the high 5D1 → T1 back-
(keto)3].1.5H2O [83], where keto is the non-steroidal anti-inflammatory
transfer rates. Still, the smallest energy gap between the T1 and 5D1
drug ketoprofen, which is capable of transferring energy for both EuIII
levels, in [Eu2(nap)6(H2O)4] 1, is responsible for the highest value of
and TbIII ions.
WBT (5D1 → T1), when compared to the others homobimetallic EuIII
The Table 4 shows the experimental (obtained from the emission
complexes. This 5D1 → T1 channel, together the high Anrad value (due
spectra) and theoretical values, for the quantum efficiency (η), radiative
the presence of the coordination water molecules in the compound)
(Arad) and nonradiative (Anrad) rates of spontaneous emission and the
explain the low quantum efficiency value (η = 25.41%) for this com-
Judd-Ofelt intensity parameters (Ω2, Ω4 and Ω6), for the EuIII ion in the
pound. The Fig. 9a–d shows the proposed energy transfer diagram for
complexes.
the [Eu2(nap)6(H2O)4] 1, [Eu2(nap)6(bpy)2] 3, [Eu2(nap)6(4,4'-
The Judd-Ofelt parameters Ω2 and Ω4 are strongly correlated with
dmbpy)2] 5 and [Eu2(nap)6(phen)2] 7, respectively. Full lines concern
the point symmetry around the EuIII ion: a high value for the Ω2
the radiative transitions, whereas the dashed lines concern those asso-
parameter means a low point symmetry around EuIII ion. In this work,
ciated with non-radiative paths. The curved lines are related to the li-
the complexes [Eu2(nap)6(bpy)2] 3 and [Eu2(nap)6(phen)2] 7 presenting
gand → EuIII energy transfer or back-transfer. The RL values, defined as
the highest values for Ω2, suggesting that EuIII ions occupy the sites of
the distance difference between donor and acceptor states involved in
smaller symmetry in these complexes, when compared with the
the process of energy transfer, depends on the molecular orbital coef-
ficients of the corresponding atoms that contributes to the ligands states
(singlet or triplet), and the distance from the that atom to the
Table 4
Experimental and theoretical values of quantum efficiencies (%η), quantum
yield (%q), radiative (Arad) and nonradiative (Anrad) rates (in s-1) and intensity Table 5
(Ω2, Ω4 and Ω6) parameters (in 10-20 cm2), for the EuIII naproxen complexes in Calculated values of intramolecular energy transfer and back-transfer rates.
study.
Complexes WET (s-1) a
WBT (s-1) b

Complexes 1 3 5 7
[Eu2(nap)6(H2O)4] 1 S1→ 5D4 1.02 × 101 5
D4→ S1 3.19 × 10-19
Ω2(exp.)/Ω2(calc.) 8.62/8.63 9.09/9.09 7.93/7.93 9.13/9.12 T1→ 5D1 8.14 × 108 5
D1→T1 1.10 × 1010
Ω4(exp.)/Ω4(calc.) 8.26/8.26 7.45/7.46 6.82/6.82 8.16/8.17 T1→ 5D0 1.45 × 109 5
D0→T1 4.82 × 106
Ω6(calc.)a 0.0932 0.0540 0.0510 0.0630 [Eu2(nap)6(bpy)2] 3 S1→ 5D4 4.48 × 102 5
D4→ S1 2.94 × 10-13
Arad(exp.) /Arad(calc.) 432.19/ 432.93/ 387.10/ 442.05/ T1→ 5D1 9.89 × 108 5
D1→T1 4.56 × 105
433.60 435.23 390.71 446.91 T1→ 5D0 1.08 × 109 5
D0→T1 1.22 × 102
Anrad(exp.) /Anrad(calc.) 1268.49/ 199.18/ 162.35/ 8.40/3.54 [Eu2(nap)6(4,4'-dmbpy)2] 5 S1→ 5D4 1.50 × 102 5
D4→ S1 5.50 × 10-14
1267.08 197.68 158.74 T1→ 5D1 1.02 × 109 5
D1→T1 7.73 × 105
η(exp)/η(calc.) 25.41/25.50 68.40/ 70.45/ 98.14/99.21 T1→ 5D0 1.14 × 109 5
D0→T1 2.12 × 102
68.77 71.11 [Eu2(nap)6(phen)2] 7 S1→ 5D4 4.02 5
D4→ S1 1.16 × 10-21
q(calc.) 21.12 68.10 70.40 98.20 T1→ 5D1 9.98 × 108 5
D1→T1 4.56 × 105
T1→ 5D0 1.09 × 109 5
D0→T1 1.22 × 102
a
Ω6 parameter could not be estimated, once that the 5D0 → 7F6 not is ob-
served. * (a WET = Transfer rate; b
WBT = Back-transfer rate).

115
E.M. Gomes, et al. Journal of Luminescence 210 (2019) 104–118

Fig. 9. Schematic energy diagram for [Eu2(nap)6(H2O)4] 1 (a), [Eu2(nap)6(bpy)2] 3 (b), [Eu2(nap)6(4,4'-dmbpy)2] 5 (c) and [Eu2(nap)6(phen)2] 7 (d) homobimetallic
naproxen complexes, showing the most probable channels for the intramolecular energy transfer process.

lanthanide ion. The RL values obtained are: 7.966 Å for X = 0.691 and Y = 0.307 for [Eu2(nap)6(H2O)4] 1, X = 0.688 and
[Eu2(nap)6(H2O)4] 1, 7.309 Å for [Eu2(nap)6(bpy)2] 3, 7.294 Å for Y= 0.310 for [Eu2(nap)6(bpy)2] 3, X = 0.689 and Y = 0.308 for
[Eu2(nap)6(4,4'-dmbpy)2] 5 and 7.293 Å for [Eu2(nap)6(phen)2] 7. The [Eu2(nap)6(4,4'-dmbpy)2] 5 and X = 0.692 and Y = 0.306 for
RL values for 3, 5 and 7 are very similar and higher that the found for 1, [Eu2(nap)6(phen)2] 7. The compounds 3, 5 and 7 produce an intense
corroborating the higher values of quantum efficiency in the com- monochromatic emission, much larger than that observed for the hy-
pounds containing the N,N-donnors ligands. drated complex, as expected. Therefore, these complexes act as light
The EuIII naproxen complexes present similar (X,Y) color co- conversion molecular devices (LCMDs) producing intense red mono-
ordinates, in the red region of the CIE chromaticity diagram chromatic emission, fact very important in biological assays for mole-
(Commission Internationale l'Eclairage), as can be seen in Fig. 10: cular recognition of NSAIDs drugs [17–24].

Fig. 10. Chromaticity diagram from emission spectra of


the [Eu2(nap)6(H2O)4] 1; [Eu2(nap)6(bpy)2] 3,
[Eu2(nap)6(4,4'-dmbpy)2] 5 and [Eu2(nap)6(phen)2] 7 na-
proxen complexes. Together are inserted the photographs
of complexes 3, 5 and 7 (with a digital camera) displaying
the intense photoluminescence in red region, under UV
irradiation.

116
E.M. Gomes, et al. Journal of Luminescence 210 (2019) 104–118

4. Conclusion [10] W.T. Carnall, K.A. Gschneidner, L. Eyring (Eds.), Handbook on the Physics and
Chemistry of the Rare Earths, 3 North-Holland Publishing Company, 1979, pp.
171–208 (Chapter 24).
Four new important classes of highly luminescent homobimetallic [11] L.F. Marques, A. Cuin, G.S.G. Carvalho, M.V. dos Santos, S.J.L. Ribeiro,
lanthanide complexes, containing the non-steroidal anti-inflammatory F.C. Machado, Inorg. Chim. Acta 441 (2016) 67–77.
drug (NSAID) naproxen, were synthesized: [Ln2(nap)6(H2O)4] (Eu 1 [12] H.P. Santos, E.M. Gomes, M.V. dos Santos, K.A. D'Oliveira, A. Cuin, J.S. Martins,
W.G. Quirino, L.F. Marques, Inorg. Chim. Acta 448 (2018) 60–68.
and Gd 2), [Ln2(nap)6(bpy)2] (Eu 3 and Gd 4), [Ln2(nap)6(4,4'-dmbpy)2] [13] M.V. Marinho, D.O. Reis, W.X. Oliveira, L.F. Marques, H.O. Stumpf, M. Déniz,
(Eu 5 and Gd 6) and [Ln2(nap)6(phen)2] (Eu 7 and Gd 8), where: nap J. Pasán, C. Ruiz-Pérez, J. Cano, F. Lloret, M. Julve, Inorg. Chem. 56 (2017)
= naproxen ligand, bpy = 2,2'-bipyridine, 4,4'-dmbpy = 4,4'-dimethyl- 2108–2123.
[14] X. Wang, Y. Guo, Y. Li, E. Wang, C. Hu, N. Hu, Inorg. Chem. 42 (2003) 4135–4140.
2,2'-bipyridine and phen = 1,10-phenanthroline. All these compounds [15] L.B.L. Escobar, G.P. Guedes, S. Soriano, R.A.A. Cassaro, J. Marbey, S. Hill,
were fully characterized and their photophysical properties were stu- M.A. Novak, M. Andruh, M.G.F. Vaz, Inorg. Chem. 57 (1) (2018) 326–334.
died in details. According to the photoluminescence study, the phos- [16] A.S. Borges, F. Fulgêncio, J.G. da Silva, T.A. Ribeiro-Santos, R. Diniz,
D. Windmoller, W.F. Magalhães, M.H. Araújo, J. Lumin. 205 (2019) 72–81.
phorescence broad bands from ligands are not present in any spectra,
[17] M. Kaczmarek, J. Fluoresc. 21 (2011) 2201–2205.
which suggest that the intramolecular energy transfer, T1 → EuIII, is [18] Y. Gao, G. Yu, K. Liu, B. Wang, Sens. Actuators B 257 (2018) 931–935.
efficient, mainly in the compounds containing the N,N-donors ligands. [19] L.G. Samsonova, T.N. Kopylova, K.M. Degtyarenko, N.V. Ponarin, S.B. Meshkova,
The molecular structures of the four EuIII complexes were determined I.I. Zheltvai, Opt. i Spektrosk. 119 (2) (2015) 231–238.
[20] N. Selivanova, K. Vasilieva, Y. Galyametdinov, Luminescence 29 (2014) 202–210.
using the Sparkle/RM1 model, with the theoretically predicted values [21] S.M. Al-Kindy, F.E. Suliman, Luminescence 22 (4) (2007) 294–301.
of intramolecular energy-transfer rates being in excellent agreement [22] A.V. Egorova, A.V. Anelchyk, I.I. Leonenko, Y.V. Skripinets, V.P. Antonovich, J.
with those experimentally obtained, corroborating the structural pro- Anal. Chem. 70 (4) (2015) 440–449.
[23] S. Panadero, A. Gómez-Hens, D. Pérez-Bendito, Anal. Chim. Acta, vol. 329(1–2), pp.
posals. Such complexes are of great importance in chemical sensing, 135–141.
since that, through the sensitization of the LnIII ion, several NSAIDs as [24] E.S. Lianidou, P.C. Ioannou, C.K. Polydorou, C.E. Efstathiou, Anal. Chim. Acta, vol.
ibuprofen, naproxen, ketoprofen, indomethacin and diflusinal, can be 320(1) pp. 107–114.
[25] T.C. de Oliveira, H.P. Santos, M.G. Lahoud, D.F. Franco, R.O. Freire, J.D.L. Dutra,
detected efficiently. Despite this important application, the literature A. Cuin, J.F. de Lima, L.F. Marques, J. Lumin. 181 (2017) 196–210.
does not report nothing about the energy transfer mechanism in this [26] T.C. de Oliveira, J.F. de Lima, M.V. Colaço, L.T. Jesus, R.O. Freire, L.F. Marques, J.
compounds, with the works being focused in the development of lu- Lumin. 194 (2018) 747–759.
[27] H.A. Flaschka, EDTA Titrations, Pergamon Press, Oxford, 1964.
minescent methods for the determination of NSAIDs. Thus, to the best [28] W.T. Carnall, H.M. Crosswhite, Energy Structure and Transitions Probabilities in
of our knowledge, this is the first work that combines experimental and LnF3 of the Trivalent Lanthanides, Argonne National Laboratory Report, Argonne,
theoretical results of EuIII complexes containing the naproxen ligand, IL, 1997.
[29] R. Pavithran, N.S.S. Kumar, S. Biju, M.L.P. Reddy, S.A. Júnior, R.O. Freire, Inorg.
can be very useful in biosensors construction field. This fact becomes
Chem. 45 (2006) 2184–2192.
even more promising, due to the solubility and high quantum efficiency [30] J.D.L. Dutra, R.O. Freire, J. Photochem. Photobio. A 256 (2013) 29–35.
values presented by these complexes. [31] J.R. Diniz, J.R. Correa, D.D.A. Moreira, R.S. Fontenele, A.L. de Oliveira,
P.V. Abdelnur, J.D.L. Dutra, R.O. Freire, M.O. Rodrigues, B.A.D. Neto, Inorg. Chem.
52 (2013) 10199–10205.
Acknowledgments [32] M. Dolg, H. Stoll, H. Preuss, J. Chem. Phys. 90 (1989) 1730–1734.
[33] T.R. Cundari, W.J. Stevens, J. Chem. Phys. 98 (1993) 5555–5565.
The corresponding author declares that this work was developed [34] R.O. Freire, G.B. Rocha, A.M. Simas, Inorg. Chem. 44 (2005) 3299–3310.
[35] R.O. Freire, G.B. Rocha, A.M. Simas, Chem. Phys. Lett. 425 (2006) 138–141.
with the researcher's own resources, since that both the federal and [36] R.O. Freire, A.M. Simas, J. Chem. Theory Comput. 6 (2010) 2019–2023.
state governments have been increasingly reducing the resources for [37] J.D.L. Dutra, M.A.M. Filho, G.B. Rocha, R.O. Freire, A.M. Simas, J.J.P. Stewart, J.
research in Brazil, endangering the Brazilian science. Our research Chem. Theory Comput. 9 (2013) 3333–3341.
[38] M.A.M. Filho, J.D.L. Dutra, G.B. Rocha, R.O. Freire, A.M. Simas, RSC Adv. 3 (2013)
group was created 5 years ago, and never received any financial sup- 16747–16755.
port, despite numerous approved projects. The corresponding author [39] M.A. Filho, J.D.L. Dutra, H.L. Cavalcanti, G.B. Rocha, A.M. Simas, R.O. Freire, J.
also thanks to Prof. Dr. Sidney José Lima Ribeiro (IQ-UNESP) by the Chem. Theory Comput. 10 (2014) 3031–3037.
[40] R.O. Freire, G.B. Rocha, A.M. Simas, J. Mol. Model. 12 (2006) 373–389.
support with the photoluminescent facilities. The author R. O. F thanks
[41] D.A. Rodrigues, N.B. da Costa, R.O. Freire, J. Chem. Inf. Model 51 (2011) 45–51.
the financial support from the Brazilian funding agencies: CAPES, [42] J.E. Ridley, M.C. Zerner, Theor. Chim. Acta 42 (1976) 223–236.
CNPq, FACEPE (APQ - 0675-1.06/14) and FAPITEC-SE (FAPITEC/SE [43] M.C. Zerner, G.H. Loew, R.F. Kirchner, U.T. Muellerwesterhoff, J. Am. Chem. Soc.
102 (1980) 589–599.
/FUNTEC/CNPq No. 04/2011).
[44] F. Neese, Wires Comput. Mol. Sci. 2 (2012) 73–78.
[45] B.R. Judd, Phys. Rev. 127 (1962) 750–761.
Appendix A. Supplementary material [46] G.S. Ofelt, J. Chem. Phys. 37 (1962) 511–520.
[47] J.D.L. Dutra, T.D. Bispo, R.O. Freire, J. Comput. Chem. 35 (2014) 772–775.
[48] A.J. Freeman, J.P. Desclaux, J. Magn. Mater. 12 (1979) 11–21.
Supplementary data associated with this article can be found in the [49] O.L. Malta, S.J.L. Ribeiro, M. Faucher, P. Porcher, J. Phys. Chem. Solids 52 (1991)
online version at https://doi.org/10.1016/j.jlumin.2019.01.066. 587–593.
[50] O.L. Malta, Chem. Phys. Lett. 87 (1982) 27–29.
[51] O.L. Malta, Chem. Phys. Lett. 88 (1982) 353–356.
References [52] J.D.L. Dutra, N.B.D. Lima, R.O. Freire, A.M. Simas, Sci. Rep. 5 (2015) 13695.
[53] F.R.G.E. Silva, O.L. Malta, J. Alloy Compd. 250 (1997) 427–430.
[1] J. Vallejo, J. Cano, I. Castro, M. Julve, F. Lloret, O. Fabelo, L. Canadillas-Delgado, [54] W.T. Carnall, H. Crosswhite, H.M. Crosswhite, Argonne National Laboratory Report,
E. Pardo, Chem. Commun. 48 (2012) 7726–7728. 1977.
[2] K. Endo, T. Koike, T. Sawaki, O. Hayashida, H. Masuda, Y. Aoyama, J. Am. Chem. [55] G.B. Deacon, R.J. Phillips, Coord. Chem. Rev. 33 (1980) 227–250.
Soc. 119 (1997) 4117–4122. [56] L.F. Marques, M.V. Marinho, N.L. Speziali, L.C. Visentin, F.C. Machado, Inorg.
[3] S.V. Eliseeva, J.-C.G. Bunzli, New. J. Chem. 35 (2011) 1165–1176. Chim. Acta 365 (2011) 454–457.
[4] X. Zhu, N. Peyghambarian, Adv. OptoElectron (2010) 501956. [57] G.S. Papaefstathiou, A. Sofetis, C.P. Raptopoulou, A. Terzis, G.A. Spyroulias,
[5] M.C. Paul, S.W. Harun, N.A.D. Huri, A. Hamzah, S. Das, M. Pal, S.K. Bhadra, T.F. Zafiropoulos, J. Mol. Struct. 837 (2007) 5–14.
H. Ahmad, S. Yoo, M.P. Kalita, A.J. Boyland, J.K. Sahu, Opt. Lett. 35 (2010) [58] Y.-L. Li, Q.-Y. Liu, C.-M. Liu, Y.-L. Wang, L. Chen, Aust. J. Chem. 68 (3) (2014)
2882–2884. 488–492.
[6] L.F. Marques, H.P. Santos, K.A. D'Oliveira, N.P. Botezine, M.C.R. Freitas, [59] R.O. Freire, G.B. Rocha, A.M. Simas, Inorg. Chem. 44 (2005) 3299–3310.
R.O. Freire, J.D.L. Dutra, J.S. Martins, C. Legnani, W.G. Quirino, F.C. Machado, [60] O.V. Dolomanov, L.J. Bourhis, R.J. Gildea, J.A.K. Howard, H. Puschmann, OLEX2: a
Inorg. Chim. Acta 458 (2017) 28–38. complete structure solution, refinement and analysis program, J. Appl. Cryst. 42
[7] H.J.M.A.A. Zijlmans, J. Bonnet, J. Burton, K. Kardos, T. Vail, R.S. Niedbala, (2009) 339–341.
H.J. Tanke, Anal. Biochem. 267 (1999) 30–36. [61] C.S. Dendrinou, G. Tsotsou, L.V. Ekateriniadou, A.H. Kortsaris, C.P. Raptopoulou,
[8] S. Faulkner, S.J.A. Pope, B.P. Burton-Pye, Appl. Spec. Rev. 40 (2005) 1–31. A. Terzis, D.A. Kyriakidis, D.P. Kessissoglou, J. Inorg. Biochem. 71 (1998) 171–179.
[9] B.R. Judd, K.A. Gschneidner, L. Eyring (Eds.), Handbook on the Physics and [62] X. Totta, A.G. Hatzidimitriou, A.N. Papadopoulos, G. Psomas, New. J. Chem. 41
Chemistry of the Rare Earths, 11 North-Holland Publishing Company, 1988, pp. (2017) 4478–4492.
81–196 (Chapter 74). [63] D.A. Gálico, T.F.C. Fraga-Silva, J. Venturini, G. Bannach, Termochim. Acta 644

117
E.M. Gomes, et al. Journal of Luminescence 210 (2019) 104–118

(2016) 43–49. (1990) 742–745.


[64] Z.N. Chen, R.W. Deng, J.G. Wu, J. Inorg. Biochem. 47 (1992) 81–87. [75] C. Görller-Walrand, K. Binnemans, K. j. Gscheidner, L. Eyring (Eds.), Handbook on
[65] Q. Xiau, Z. Yanbin, L. Xia, J. Rare Earths 27 (5) (2009) 797–800. the Physics and Chemistry of Rare Earths, 23 Elsevier, Amsterdam, 1996, pp.
[66] R. Feng, F.-L. Jiang, M.-Y. Wu, L. Chen, C.-F. Yan, M.-C. Hong, Cryst. Growth Des. 121–283.
10 (2010) 2306–2310. [76] S.T. Frey, H.D. Horrocks, Inorg. Chim. Acta 229 (1995) 383–390.
[67] L.F. Marques, C.C. Correa, H.C. Garcia, T.M. Francisco, S.J.L. Ribeiro, J.D.L. Dutra, [77] L.D. Carlos, O.L. Malta, R.Q. Albuquerque, Chem. Phys. Lett. 415 (2005) 238–242.
R.O. Freire, F.C. Machado, J. Lumin. 148 (2014) 307–316. [78] P.A. Tanner, Chem. Soc. Rev. 42 (2013) 5090–5101.
[68] I.M.P.E. Silva, D.M. Profirio, R.E.F. de Paiva, M. Lancellotti, A.L.B. Formiga, [79] W.M. Faustino, O.L. Malta, E.E.S. Teotônio, H.F. Brito, A.M. Simas, G.F. de Sá, J.
P.P. Corbi, J. Mol. Struct. 1049 (2013) 1–6. Phys. Chem. A 110 (2006) 2510–2516.
[69] J.A. Fernandes, R.A.S. Ferreira, M. Pillinger, L.D. Carlos, I.S. Gonçalves, [80] C.Y. Su, M.Y. Tan, N. Tang, W. Liu, X. Wang, J. Coord. Chem. 38 (1996) 207–218.
P.J.A. Ribeiro-Claro, Eur. J. Inorg. Chem. (2004) 3913–3919. [81] R.A.S. Ferreira, M. Nolasco, A.C. Roma, R.L. Longo, O.L. Malta, L.D. Carlos, Chem.
[70] S. Shuvaev, M. Starck, D. Parker, Chem. Eur. J. 23 (42) (2017) 9974–9989. Eur. J. 18 (2012) 12130–12139.
[71] P.P. Lima, M.M. Nolasco, F.A.A. Paz, R.A.S. Ferreira, R.L. Longo, O.L. Malta, [82] M. Latva, H. Takalo, V.M. Mukkala, C. Matachescu, J.C. Rodriguez-Ubis,
L.D. Carlos, Chem. Mater. 25 (2013) 586–598. J. Kanakare, J. Lumin. 75 (1997) 149–169.
[72] S. Tobita, M. Arakawa, I. Tanaka, J. Phys. Chem. 89 (1985) 5649–5654. [83] D.A. Gálico, M.G. Lahoud, M.R. Davolos, R.C.G. Frem, T.F.C. Fraga-Silva,
[73] S. Tobita, M. Arakawa, I. Tanaka, J. Phys. Chem. 88 (1984) 2697–2702. J. Venturini, M.S.P. Arruda, G. Bannach, J. Inorg. Biochem. 140 (2014) 160–166.
[74] A.P. Suisalu, V.N. Zakharov, A.L. Kamyshnyi, L.A. Aslanov, Sov. Phys. JETP 71 (4)

118

You might also like