Next Article in Journal
Particle Size Related Effects of Multi-Component Flame-Retardant Systems in poly(butadiene terephthalate)
Next Article in Special Issue
Doxorubicin–Gelatin/Fe3O4–Alginate Dual-Layer Magnetic Nanoparticles as Targeted Anticancer Drug Delivery Vehicles
Previous Article in Journal
Polystyrene Magnetic Nanocomposites as Antibiotic Adsorbents
Previous Article in Special Issue
Synthesis, Physical, Mechanical and Antibacterial Properties of Nanocomposites Based on Poly(vinyl alcohol)/Graphene Oxide–Silver Nanoparticles
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Luminescent Properties of Lanthanoid-Poly(Sodium Acrylate) Composites: Insights on the Interaction Mechanism

by
Alan F. Y. Matsushita
1,
María José Tapia
2,
Alberto A. C. C. Pais
1 and
Artur J. M. Valente
1,*
1
CQC, Department of Chemistry, University of Coimbra, 3004-535 Coimbra, Portugal
2
Department of Chemistry, Universidad de Burgos, 09001 Burgos, Spain
*
Author to whom correspondence should be addressed.
Polymers 2020, 12(6), 1314; https://doi.org/10.3390/polym12061314
Submission received: 12 May 2020 / Revised: 3 June 2020 / Accepted: 4 June 2020 / Published: 9 June 2020
(This article belongs to the Special Issue Metal- and Metal Hybrid-Filled Polymer Nanocomposites)

Abstract

:
The interaction between polyelectrolytes and metal ions is governed by different types of interactions, leading to the formation of different phases, from liquid state to weak gels, through an appropriate choice of metal ion/polyelectrolyte molar ratio. We have found that lanthanide ions, europium(III) and terbium(III), are able to form polymer composites with poly(sodium acrylate). That interaction enhances the luminescent properties of europium(III) and terbium(III), showing that Eu3+/poly(sodium acrylate) (PSA) and Tb3+/PSA composites have a highly intense red and green emission, respectively. The effect of cations with different valences on the luminescent properties of the polymer composites is analyzed. The presence of metal ions tends to quench the composite emission intensity and the quenching process depends on the cation, with copper(II) being by far the most efficient quencher. The interaction mechanism between lanthanoid ions and PSA is also discussed. The composites and their interactions with a wide range of cations and anions are fully characterized through stationary and non-stationary fluorescence, high resolution scanning electronic microscopy and X-ray diffraction.

Graphical Abstract

1. Introduction

Metal-organic materials have received great attention in recent years, because of the metal-ligand interaction, which can form three-dimensional structures, such as metal-organic frameworks (MOF) and metal-organic gels (MOG); these materials have already presented promising applications for supercapacitors, sensing, catalysis and optoelectronics [1,2,3,4,5]. MOG preparation is simpler, involving hydrogen bonding interactions, π-π stacking, van der Waals forces and coordination bonds under mild conditions to form self-assembled supramolecular structures [6,7]. On the other hand, MOF usually display a highly crystalline structure, therefore requiring a more time-consuming preparation [8]. MOG, on the contrary, are characterized by extended structures, essentially driven by metal-ligand interactions, where the polymer (ligand) acts as a gelling agent [9]. These materials may show selective sensing to different stimuli including cations and anions [6,10]. Some advantages of these MOGs include high sensitivity, availability, fast response to a stimulus, low cost (small amounts of probe is required) and fluorescence, which is a technique frequently used in the sensing field [11]. In this regard, lanthanoid complexes have interesting spectroscopic properties—such as large Stokes shift, narrow emission bands, long emission lifetimes and high emission quantum yields [12]—being a good alternative as sensors for detection of cations, anions and biomolecules. In fact, the detection of metals, such as copper(II)—a relevant ion for biochemical and environmental issues [13,14,15]—can be done by different methods, including electrochemical methods, UV-vis spectroscopy, atomic absorption, inductively coupled plasma and atomic emission spectrometry (ICP-AES) [16,17,18,19]. These methods generally require costly equipment and the specific preparation of samples. Hence, the development of fluorescent sensors for metal ions has increased in recent years [20,21,22,23].
Following our previous work [24], we describe here the synthesis and characterization of two new polymer composites consisting of lanthanoid metal ions (europium and terbium) and a polyelectrolyte. The luminescent properties of these compounds were assessed, making it possible to have an insight into the mechanism of interactions between lanthanoid ion and the polyelectrolyte. Additionally, the effect of cations with different stable valences and good solubility were assessed to better understand the hypothetical competition mechanism with lanthanoid ions on the poly (sodium acrylate)-Ln3+ interaction; moreover, and once that salts with different monovalent anions were used, a series of monovalent anions were also studied. That made it possible to shed light on the possible application of these matrices for sensing, remediation or anti-counterfeiting.

2. Materials and Methods

2.1. Reagents

All chemicals were commercially available and used without further purification. The following salts: europium(III) chloride hexahydrate, terbium(III) chloride hexahydrate, aluminum hydroxide hydrate, calcium chloride dihydrate, cerium(III) chloride heptahydrate, chromium(III) chloride hexahydrate, copper(II) chloride dihydrate, mercury(II) thiocyanate, potassium chloride, sodium chloride, sodium acetate, sodium bromide, sodium cyanide, sodium fluoride, nickel nitrate hexahydrate, lead(II) nitrate, zinc chloride and poly(sodium acrylate) (PSA, M w ¯ = 2100 g mol−1) were purchased from Sigma-Aldrich (Steinheim, Germany). Sodium acetate, magnesium nitrate hexahydrate, were obtained from Fluka (Gillingham, UK) and anhydrous aluminum chloride, sodium nitrite and potassium thiocyanate were purchased from Merck (Darmstadt, Germany). Milli-Q water was used in all the experiments.

2.2. Eu3+/PSA and Tb3+/PSA Composite Solutions

The Eu3+/PSA and Tb3+/PSA composites were prepared by dissolving PSA (0.26 mol dm−3, in terms of polymer repetition units, whose repetitive unit molecular weight is 108.08 g mol−1) in milli-Q water, and the subsequent dropwise addition of europium or terbium chloride hexahydrates (0.026 mol dm−3), under continuous magnetic stirring. The same method was used elsewhere [24]. The pH of the solutions was adjusted at around 7.5, by adding either NaOH or HCl (0.1 mol dm−3), due the pKa value of PSA and the solubility of Eu3+ and Tb3+ [25]. It has been found that the addition of salts did not change the solution pH. Additionally, the stability of solutions was verified spectrophotometrically for at least 2 weeks after preparation [24].

2.3. Preparation of Metal-Organic Gels

Freshly prepared solutions of Eu3+/PSA and Tb3+/PSA were used for the luminescent studies. To investigate the selectivity of the composites towards different metal ions, aliquots of metallic ion stock solutions (Al3+, Ca2+, Ce3+, Cr3+, Cu2+, Hg2+, K+, Mg2+, Na+, Ni2+, Pb2+ and Zn2+) were added to Eu3+/PSA and/or Tb3+/PSA solution. The final concentration of metal ions in the solution was fixed at 3.33 mM. The solution was mixed using an ultrasonic bath for 30 min, before the emission spectra was recorded. All the spectra were registered at 25 °C.

2.4. Apparatus and Characterization Methods of Eu(III)/PSA and Tb(III)/PSA

The UV–vis spectra of solutions were recorded on a Shimadzu 2450 UV–vis spectrophotometer (Kyoto, Japan).
Emission spectra were recorded with a Fluoromax-4 spectrofluorometer (Kyoto, Japan), in a right-angle configuration, with excitation at 395 nm (Eu3+) and 273 nm (for Tb3+), and emission spectra scanned between 550 and 750 nm and 450 and 650 nm, respectively. Excitation and emission slits of 0.5 and 1.0 nm, respectively, were used.
Tb3+ and Eu3+ luminescence lifetimes were measured with the multi-channel scaling mode (MCS) counter module (TCC2) of the FLS980 spectrometer (Edinburgh Instruments, Livingston, UK), using a microsecond flash Xe-lamp (μF2) as light source. The decays were registered by exciting the composites at 395 nm and 270 nm and recording the emission at 613 nm and 543 nm for Eu3+/PSA and Tb3+/PSA, respectively. The decays were analyzed with the software of the equipment.
High Resolution Scanning Electronic Microscopy (HR-SEM) micrographs were obtained using a ZEISS Merlin scanning electron microscope (Oerzen, Germany), operating under low vacuum at 2kV. Elemental analysis on microscopic sections of composites was performed by Energy Dispersive Spectroscopy—Oxford Instruments, Oxon, UK. The samples used in these techniques were previously frozen at −20 °C and then lyophilized (Free Zone 4.5-Labconco), before being sputter-coated with a thin gold layer.
X-ray diffractograms were obtained with freeze-dried samples with the (X-ray Diffractometer Rigaku Ultima IV, Tokyo, Japan) with the following characteristics: Cu-kα radiation (1.54 A) in the range of 5° to 120° at a scan speed of 2θ/min.

3. Results

3.1. Lanthanoid-PSA Interactions

The Eu3+/PSA and Tb3+/PSA complexes, synthesized using a previous method [24], show an enhanced luminescence intensity in comparison to the free lanthanoides (see Figure 1). The metal-polymer interactions are likely due to the PSA carboxylic groups acting as chelators to the lanthanoid ions.
From analysis of Figure 1, it can be observed that upon excitation at 395 nm Eu3+/PSA exhibits strong emission peaks at 580, 593, 616, 652 and 694 nm. These peaks can be assigned to the characteristic 5D07FJ (J = 0–4) transitions of the Eu3+ ion. The 5D07F1 transition is a magnetic dipole transition, which corresponds to the weak emission band at 593 nm, less sensitive to the coordination environment. Therefore, when Eu3+ ions occupy the non-inversion centers sites, the electric dipole transition 5D07F2 is dominant promoting red-light emission. By comparing the transitions 5D07F2 and 5D07F1, we obtain an intensity ratio I(5D07F2)/I(5D07F1) of about 3.4, thus indicating that Eu3+ occupies non-inversion centers [26,27]. Similarly, upon excitation at 273 nm, Tb3+/PSA displays four clearly resolved peaks at 490, 545, 585 and 623 nm, which can be ascribed to 5D47FJ (J = 6–3) transitions of Tb3+ ions. 5D47F5 is an electric induced dipole transition, which is characterized by the intense emission band at 545 nm, more sensitive to the coordination environment [28,29]. Additionally, the lanthanoid(III)-polymer interaction should lead to a concomitant dehydration of metal ions [30]. To have an assessment on such hypothesis, non-stationary fluorescence experiments were carried out, to allow the computation of the number of water molecules coordinating the lanthanoid; this can be done by comparing the fluorescence lifetime of Tb3+ and Eu3+ in aqueous and D2O solutions [30,31,32]. The decays of Tb3+ and Eu3+ (2.7 × 10−2 mol dm−3) luminescence were registered in the absence and in the presence of PSA (0.23 mol dm−3, in terms of monomer), both in H2O and D2O solutions. Good monoexponencial decays were observed in all the cases, from which lifetimes (τ) and decay constants (k, reciprocal lifetime) were obtained (Table 1). From this, the number of bound water molecules (n) was computed. It has been found that Eu3+ and Tb3+, in aqueous solution, are coordinated by nine water molecules, which is in good agreement with the literature data [31]. However, in the presence of PSA, the number of bound water molecules decrease to approximately three. Such a decrease in the number of water coordinated molecules have also been observed for lanthanoids interacting with poly(vinyl sulfate) [33] or sodium dodecyl sulfate [34]. It can be concluded that PSA can replace up to six lanthanoid water coordinated molecules, indicating a strong interaction with Ln3+ ions, partially inhibiting the efficient radiationless lanthanoid emission deexcitation pathway via energy transfer to OH vibrational overtones [31,35]. This fact explains the significant increase of Eu3+ and Tb3+ emission intensity upon complexation with PSA (Figure 1).
A further insight into the interaction mechanism can also be gained from the analysis of lanthanoid/PSA solid composite. The XRD patterns of PSA, and Eu3+ and Tb3+ composites are shown in Figure 2. In the PSA diffraction pattern, we can see only a large shoulder diffuse peak, showing the amorphous structure of the polymer [36]. Unlike this, Eu3+/PSA and Tb3+/PSA diffractograms show a series of diffraction peaks at 2θ = 27.4, 31.5, 45.4, 56.5, 66.3, 75.4 and 83.9, and these peaks are assigned to the (111), (200), (220), (400), (420) and (422) plane of the NaCl-type structure [37]. This fact corroborates the occurrence of lanthanoid-PSA interaction, since the formation of NaCl is a consequence of the release and consequent interaction of counterions. It is also worth noting that no peak corresponding to the EuCl3 and TbCl3 crystal structure is observed, confirming that an interaction occurs, instead of a simple doping.
The SEM micrograph (Figure 2, right side) shows that the composite is covered with crystals, but the cross-section has a different and featureless morphology suggesting a different composition. Finally, the elemental mapping and EDX spectra (Figure S1) of Eu3+/PSA and Tb3+/PSA gel composites reveal a homogeneous distribution of lanthanoides through the polymer covered by a shell of NaCl.

3.2. Effect of Cations on the Luminescence Properties Lanthanoid/PSA Complexes

As it was discussed in the previous section, Eu3+/PSA and Tb3+/PSA complexes have carboxyl groups available to coordinate metal ions, affecting their spectroscopic properties and making these composites potentially useful as luminescent probes for different ions. In consequence, the ability of these composites to detect different cations, in aqueous solution, has been tested by recording the luminescence emission spectra of Eu3+/PSA and Tb3+/PSA in the presence of different cations: Al3+, Ca2+, Ce3+, Cr3+, Cu2+, Hg2+, K+, Mg2+, Na2+, Ni2+, Pb2+ and Zn2+, as aqueous solutions of NO3, Cl and SCN (Section 2.1). The results indicate that the luminescence intensity of the composites is strongly dependent on the metal ion species. Figure 3 shows the quenching efficiency of metal ions on the luminescence intensity of composite, defined as (I0I)/I0 × 100%, where I0 and I are the luminescence intensity without and with the addition of metal ions, respectively.
Remarkably, Cu2+ is the most effective quencher for both composites. It completely quenches the emission of Tb3+/PSA at 545 nm (changing Tb3+/PSA emission from green to almost colorless in the presence of Cu2+ (Figure 4) and more than 70% of Eu3+/PSA emission at 616 nm. The emission of the latter one changed from red to light red, while the other metals did not have such a pronounced effect on luminescence intensity for Eu3+/PSA (Figure 4). This indicates that Tb3+/PSA possesses a higher sensitivity to Cu2+, being a promising sensor for the detection of Cu2+ ions. This behavior will be discussed in terms of the suppression mechanism. In general, the transitional metal ions display a stronger effect in luminescence compared to alkaline metal ions and alkaline earth metals. The effect arises from unpaired d-electrons found in transition metal ions contrasting to the closed shell electron configuration of group I and II metal cations [28,38].

3.3. Effect of Anions on Lanthanoid/PSA-Based Complex Properties

In order to assess whether Eu3+/PSA and Tb3+/PSA composites can detect Cu2+ in the presence of other ions, the effects of several metal ions on the composite emission intensity were examined under the same experimental conditions. Firstly, the counter ion effect on the composite luminescence quenching was checked. For that, different sodium salts have been checked: Br, Cl, CN, F, NO2, NO3, OAc, OH and SCN. The results depicted in Figure 5 show no significant change in the luminescence emission intensity of Eu3+/PSA as the anion is varied (with variations around 5% at maximum, Figure 5 left) and, in consequence, it can be concluded that anions have no significant effect upon the detection of Cu2+ ions with Eu3+/PSA.
In contrast, the luminescence emission of Tb3+/PSA composite is effectively quenched (75%) by nitrite ion. This can be attributed to energy transfer from Tb3+ to NO2 [6] and, consequently NO2 might be considered as an interferent for the Cu2+ detection for the Tb3+/PSA composite. The energy transfer may occur from Tb3+ to NO2, due the 5D4 energy level of Tb3+ that matches with T1 energy level of NO2; the same does not happen with Eu3+ [6].
In order to gain an insight into the potential selective detection of Eu3+/PSA and Tb3+/PSA composite toward Cu2+ in aqueous solution, the selectivity detection and the potential interference with other metal ions was studied by using Eu3+/PSA and Tb3+/PSA, and Cu2+ (3.33 M) in the presence of other metal ions (at equimolar concentrations). In both cases, the composite emission quenching by Cu2+ was hardly affected, as shown in Figure S2. The results indicate that Cu2+ detection is weakly perturbed by coexisting cations in solutions, confirming that the composites could selectively detect Cu2+, even in the presence of the other competing metal ions, clearly suggesting the potential of these composites as Cu2+ probes. The high selectivity of these composites towards Cu2+ could take place via interactions between functional sites, such as the uncoordinated carboxylic oxygen atom of PSA with Cu2+ ions, which will be discussed as follows.

3.4. On the Interaction Mechanism Between Cu(II) and Lanthanoid/PSA-Based Complexes

Having Cu2+ ions showed high interaction with composites, the nature of the quenching process should be unveiled. For that, the lifetimes of Eu3+/PSA and Tb3+/PSA composites were measured in absence, and in the presence of different concentrations of Cu2+. The decay curves for Eu3+/PSA and Tb3+/PSA in the presence of variable concentrations of Cu2+ were also monoexponencial, which indicates that Cu2+ does not replace the lanthanoids in the PSA chain, which would give rise to biexponential decays, due to the simultaneous emission of the composite and the free lanthanoids.
The lifetimes fit properly the Stern-Volmer equation.
τ0/τ = 1 + KSV,D[Q]
where τ0 and τ are the composite lifetime before and after the incorporation of the metal cation respectively, KSV,D is the dynamic Stern-Volmer constant and [Q] = [Cu2+], as shown in Figure 6. The computed dynamic Stern-Volmer constants are equal to 44.0 (±0.5) and 583 (±7) M−1 for Eu3+/PSA and Tb3+/PSA, respectively, which are an order of magnitude lower than the ones obtained with the emission intensities Stern-Volmer plot (580 and 5655 M−1, for the same composites, respectively). These results confirm that both static and dynamic quenching are simultaneously taking place [11,26,39,40].
The Cu2+ paramagnetism contributes to the quenching of the fluorescence emission by increasing the intersystem crossing process [29,41,42,43,44]. The higher Tb3+/PSA sensitivity to Cu2+ with respect to that of Eu3+/PSA could be partially due to the overlap between the absorption spectra of both Tb3+/PSA and Cu2+ (spectra not shown), which is likely to significantly diminish Tb3+/PSA emission intensity [28,45,46]. In general, PSA plays an important role in the quenching process, due to its ability to coordinate Cu2+ ions. This interaction can enhance the proximity between the composites and Cu2+ promoting a dynamic quenching and, simultaneously, form a new non-luminescent composite. We hypothesize that the polymer may act as a surface, which promotes sensor luminescence quenching by adsorbing Cu2+ ions through the complexation with the polymer carboxylic groups, as will be confirmed by EDX mapping [47]. As stated before, such an interaction can also be used for the application of these composites on the selective removal of Cu2+ as an added value metal.

3.5. Effect of pH and Quenching Rate

The emission intensity of Eu3+/PSA/Cu2+ and Tb3+/PSA/Cu2+ composites, with three different Cu2+ concentrations for each system, were recorded at 616 and 545 nm at several delays after sample preparations (Figure S3), showing that, independently of Cu2+ concentrations, the emission intensity reaches an equilibrium almost instantaneously (around a minute), remaining constant for almost a week [48], showing the great stability of the composites. Additionally, the emission intensities of Eu3+/PSA/Cu2+ and Tb3+/PSA/Cu2+ composites, at 616 and 545 nm, respectively, recorded for solutions with different pH values (Figure S4) were approximately the same, proving that composites were stable over a wide pH range (6–11), and indicating their potential applications on a large pH range, including environmental and physiological conditions.

3.6. Sensitivity of Eu3+/PSA and Tb3+/PSA to Cu2+

Sensitivity is a key factor to evaluate the performance of a probe to determine traces of an analyte and also for remediation purposes. To further examine the Eu3+/PSA and Tb3+/PSA composite sensing sensitivity towards Cu2+, their luminescence emission quenching by CuCl2 (at concentrations ranging from 0.33 mM to 3.33 mM) were studied. The emission luminescence spectra of composites are shown in Supplementary Information (Figures S5–S8). Moreover, Figure 7a shows the emission luminescence intensity of Eu3+/PSA, at 616 nm which gradually decreases with Cu2+ concentration. When the concentration 3.33 mM Cu2+ is reached, the quenching efficiency is above 75%.
Figure 7c shows the ratio of emission intensities in absence (I0) and presence of Cu2+ (I) of 5D07F2 emission at 616 nm versus the concentration of Cu2+ for metal ion concentrations lower than 10−3 M. In this concentration range, Eu3+/PSA emission quenching follows the Stern-Volmer equation (Equation (1)). A good determination coefficient (R2 = 0.997) was obtained for the fitting of Equation (1) to experimental data, leading to a Stern-Volmer constant, KSV, equal to 580 (±1) M−1. The limit of detection (LOD) for ions Cu2+ has been calculated by using the equation: 3σ/k, where σ is the standard error and k the slope. The LOD has been calculated as 3.06 × 10−5 M (corresponding to 1.94 ppm). Therefore, Eu3+/PSA presents an excellent sensitivity towards Cu2+. Similar sensitivity tests were carried out with Tb3+/PSA composites, and results are shown in Figure 7b,d. The fluorescence intensity at 545 nm is drastically quenched by the gradual increase of the concentration of Cu2+ and for Cu2+ concentrations below 1.8 × 10−3 M, the Stern-Volmer equation is followed (Figure 7d). The calculated KSV value is 5655 (±68) M−1 and the limit of detection for Cu2+ is 3.56 × 106 M (0.226 ppm), indicating a higher sensitivity to Cu2+ when compared to Eu3+/PSA. KSV values obtained for Cu2+ in both composites are relatively high, and can be compared with the values obtained for other compounds previously used for Cu2+ detection (Table 2). This indicates that both composites have great potential for application in Cu2+ ion sensing, especially Tb3+/PSA, which shows greater sensitivity.

3.7. Characterization of Solid Lanthanoid/PSA Composites Containing Cu2+

SEM micrographs of the composites obtained with the objective of investigating the morphology and structure of the composites after contact with Cu2+ (Figure S9), indicate that the structure maintains its integrity and, therefore, the quenching effect is not a consequence of structure collapse. The EDX spectra of composite, obtained after immersion in a Cu2+ solution, show the presence of these ions, suggesting a strong interaction between Cu2+ and the composites. To support this hypotheses, elemental EDX maps have been made, and confirm the interaction of Cu2+ with the composites in the region where the composites are not coated with NaCl crystals. This also shows that most of the Cu2+ added is bound to the composites having a uniform distribution in the same region that Eu and Tb (Figure 8 and Figure S10, respectively). From the elemental maps, we can also conclude that in the Eu3+/PSA and Tb3+/PSA composites a simultaneous interaction of the polymer with both the lanthanoid and Cu2+ ions is produced; i.e., Cu2+ ions do not replace the lanthanoides in the PSA chain, in good agreement with the previously discussed lanthanoids lifetime results. The surface of the composites is characterized by a high content of oxygen, making available sites of interaction with Cu2+ and other metallic ions. We can therefore suggest that the main mechanism of the composite luminescent quenching induced by metal ions is the Cu2+ complexation, with the composite through the carboxylate groups of Ln3+/PSA.

4. Conclusions

Two highly luminescent water soluble stable metal-organic composites (Eu3+/PSA and Tb3+/PSA) were prepared, showing a significant emission quenching in the presence of Cu2+, when compared with a set of other metal ions: Al3+, Ca2+, Ce3+, Cr3+, Hg2+, K+, Mg2+, Na+, Ni2+, Pb2+ and Zn2+ that, in general, do not compete with copper ions for the composite. Polymer composite emission intensities and lifetimes in the presence of Cu2+ follow Stern-Volmer kinetics, indicating that both static and dynamic quenching processes take place simultaneously, with the former being one order of magnitude higher for both composites. This probes that complex formation is the main interaction mechanism between the lanthanoid/PSA composites and Cu2+, likely through the polymer carboxylate groups. The composite shows a fast response to the presence of Cu2+ (less than one minute) and linear for Cu2+ concentration below 1 mM (Eu3+/PSA) and 1.8 × 10−3 M (Tb3+/PSA), within an extended pH range (6 to 11), and with detection limits of 1.94 and 0.22 ppm, in the case of Eu3+/PSA and Tb3+/PSA, respectively. In our experimental conditions, no competition has been detected between the lanthanides and copper ions, indicating that the PSA has carboxylate groups available for coordination after the interaction with lanthanoid ions. Additionally, nitrite ions also promote high quenching efficiency for Tb3+/PSA composite through an energy transfer process. The experimental results reported have unraveled a new promising method for a simple and reliable monitoring or removal of Cu2+ from different media. Moreover, these composites can be used both as aqueous solution and in a gel state, enlarging the potential range of their practical applications.

Supplementary Materials

The following are available online at https://www.mdpi.com/2073-4360/12/6/1314/s1, Figure S1: EDX spectra and elemental maps of freeze-dried Eu3+/PSA (left) and Tb3+/PSA (right), Figure S2: Comparison of emission intensity of Eu3+/PSA at 616 nm (left) and Tb3+/PSA at 545 nm (right) of composites alone (I0) and interacting with different metal ions in aqueous solution under the same conditions (I), Figure S3: Emission intensities of (A) Eu3+/PSA at 616 nm (left) and (B) Tb3+/PSA at 545 nm (right) with different concentration of Cu2+ ions in aqueous solution at several delays after sample preparation, Figure S4: Effects of pH on the emission intensities of (A) Eu3+/PSA at 616 nm and (left) (B) Tb3+/PSA at 545 nm (right), without Cu2+ (black) and with 3.33 mM of Cu2+ (red), Figure S5: Emission spectra of Eu3+/PSA, 5D07F2 transition emission intensity at 616 nm (inset) of Eu3+/PSA with different concentrations of Cu2+, T S6: Emission spectra of Eu3+/PSA and Stern-Volmer plot of 5D07F2 transition at 616 nm (inset) for Eu3+/PSA with lower concentrations of Cu2+, Figure S7: Emission spectra of Tb3+/PSA and 5D47F5 transition intensities at 545 nm (inset) for Tb3+/PSA with different concentrations of Cu2+, Figure S8: Emission spectra of Tb3+/PSA and Stern-Volmer plot of 5D47F5 transition at 545 nm (inset) for Tb3+/PSA with lower concentrations of Cu2+, SEM micrographs of (a) Eu3+/PSA, (b)Tb3+/PSA before contact with Cu2+ and (c) Eu3+/PSA, (d) Tb3+/PSA after contact with Cu2+. Magnification ×500, Figure S10: (a) SEM micrography, (b) EDX spectra and elemental maps of Tb3+/PSA after contact with Cu2+.

Author Contributions

Conceptualization, A.F.Y.M., A.A.C.C.P. and A.J.M.V.; Data curation, A.F.Y.M., M.J.T. and A.J.M.V.; Formal analysis, A.F.Y.M., A.A.C.C.P., M.J.T. and A.J.M.V.; Investigation, A.F.Y.M.; Methodology, A.F.Y.M., M.J.T. and A.J.M.V.; Project administration, A.J.M.V., A.A.C.C.P.; Resources, A.A.C.C.P., M.J.T. and A.J.M.V.; Supervision, A.A.C.C.P., M.J.T. and A.J.M.V.; Writing—original draft, A.F.Y.M., M.J.T. and A.J.M.V.; Writing—review and editing, A.F.Y.M., A.A.C.C.P., M.J.T. and A.J.M.V. All the authors participated in the analysis of the results. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by the Coimbra Chemistry Centre which is supported by the Fundação para a Ciência e a Tecnologia (FCT) through the programmes UID/QUI/UI0313/2020 and COMPETE.

Acknowledgments

A.F.Y.M. acknowledges financial support from the Conselho Nacional de Desenvolvimento Científico e Tecnológico (CNPq—Brazil)—Science without borders, through the PhD grants (Grant number 249241/2013-7).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Xu, G.; Nie, P.; Dou, H.; Ding, B.; Li, L.; Zhang, X. Exploring metal organic frameworks for energy storage in batteries and supercapacitors. Mater. Today 2017, 20, 191–209. [Google Scholar] [CrossRef]
  2. Yao, C.-X.; Zhao, N.; Liu, J.-C.; Chen, L.-J.; Liu, J.-M.; Fang, G.; Wang, S. Recent Progress on Luminescent Metal-Organic Framework-Involved Hybrid Materials for Rapid Determination of Contaminants in Environment and Food. Polymers 2020, 12, 691. [Google Scholar] [CrossRef] [Green Version]
  3. Zhu, L.; Liu, X.-Q.; Jiang, H.-L.; Sun, L.-B. Metal–Organic Frameworks for Heterogeneous Basic Catalysis. Chem. Rev. 2017, 117, 8129–8176. [Google Scholar] [CrossRef] [PubMed]
  4. He, L.; Peng, Z.W.; Jiang, Z.W.; Tang, X.Q.; Huang, C.; Li, Y.F. Novel Iron(III)-Based Metal–Organic Gels with Superior Catalytic Performance toward Luminol Chemiluminescence. ACS Appl. Mater. Interfaces 2017, 9, 31834–31840. [Google Scholar] [CrossRef] [PubMed]
  5. Khajavi, H.; Gascon, J.; Schins, J.M.; Siebbeles, L.D.A.; Kapteijn, F.; Kapteijn, F. Unraveling the Optoelectronic and Photochemical Behavior of Zn4O-Based Metal Organic Frameworks. J. Phys. Chem. C 2011, 115, 12487–12493. [Google Scholar] [CrossRef]
  6. Yuan, D.; Zhang, Y.D.; Jiang, Z.W.; Peng, Z.W.; Huang, C.; Li, Y.F. Tb-containing metal-organic gel with high stability for visual sensing of nitrite. Mater. Lett. 2018, 211, 157–160. [Google Scholar] [CrossRef]
  7. Lloyd, G.O.; Steed, J.W. Anion-tuning of supramolecular gel properties. Nat. Chem. 2009, 1, 437–442. [Google Scholar] [CrossRef] [Green Version]
  8. Sutar, P.; Maji, T.K. Coordination polymer gels: Soft metal–organic supramolecular materials and versatile applications. Chem. Commun. 2016, 52, 8055–8074. [Google Scholar] [CrossRef]
  9. Zhang, J.; Su, C.-Y. Metal-organic gels: From discrete metallogelators to coordination polymers. Coord. Chem. Rev. 2013, 257, 1373–1408. [Google Scholar] [CrossRef]
  10. Tian, D.; Li, Y.; Chen, R.-Y.; Chang, Z.; Wang, G.; Bu, X.-H. A luminescent metal–organic framework demonstrating ideal detection ability for nitroaromatic explosives. J. Mater. Chem. A 2014, 2, 1465–1470. [Google Scholar] [CrossRef]
  11. Lakowicz, J.R. (Ed.) Principles of Fluorescence Spectroscopy; Springer US: Boston, MA, USA, 2006; ISBN 978-0-387-31278-1. [Google Scholar]
  12. Liu, J.; Zuo, W.; Zhang, W.; Liu, J.; Wang, Z.; Yang, Z.; Wang, B. Europium(III) complex-functionalized magnetic nanoparticle as a chemosensor for ultrasensitive detection and removal of copper(II) from aqueous solution. Nanoscale 2014, 6, 11473–11478. [Google Scholar] [CrossRef] [PubMed]
  13. Aksuner, N.; Henden, E.; Yilmaz, I.; Çukurovalı, A.; Yılmaz, I. A highly sensitive and selective fluorescent sensor for the determination of copper(II) based on a schiff base. Dye. Pigment. 2009, 83, 211–217. [Google Scholar] [CrossRef]
  14. Kluczka, J. Removal of Boron and Manganese Ions from Wet-Flue Gas Desulfurization Wastewater by Hybrid Chitosan-Zirconium Sorbent. Polymers 2020, 12, 635. [Google Scholar] [CrossRef] [Green Version]
  15. Vareda, J.P.; Valente, A.J.; Durães, L. Assessment of heavy metal pollution from anthropogenic activities and remediation strategies: A review. J. Environ. Manag. 2019, 246, 101–118. [Google Scholar] [CrossRef] [PubMed]
  16. Reddyprasad, P.; Imae, T. Selective detection of copper ion in water by tetradentate ligand sensor. J. Taiwan Inst. Chem. Eng. 2017, 72, 194–199. [Google Scholar] [CrossRef]
  17. Ghaedi, M.; Shokrollahi, A.; Kianfar, A.; Mirsadeghi, A.; Pourfarokhi, A.; Soylak, M. The determination of some heavy metals in food samples by flame atomic absorption spectrometry after their separation-preconcentration on bis salicyl aldehyde, 1,3 propan diimine (BSPDI) loaded on activated carbon. J. Hazard. Mater. 2008, 154, 128–134. [Google Scholar] [CrossRef] [PubMed]
  18. Ghaedi, M.; Tashkhourian, J.; Montazerozohori, M.; Biyareh, M.N.; Sadeghian, B. Highly selective and sensitive determination of copper ion by two novel optical sensors. Arab. J. Chem. 2017, 10, S2319–S2326. [Google Scholar] [CrossRef] [Green Version]
  19. Gómez, M.R.; Cerutti, S.; Sombra, L.L.; Silva, M.F.; Martínez, L.D. Determination of heavy metals for the quality control in argentinian herbal medicines by ETAAS and ICP-OES. Food Chem. Toxicol. 2007, 45, 1060–1064. [Google Scholar] [CrossRef]
  20. Wang, Q.; Chen, K.; Qu, Y.; Li, K.; Zhang, Y.; Fu, E. Hairy Fluorescent Nanospheres Based on Polyelectrolyte Brush for Highly Sensitive Determination of Cu(II). Polymers 2020, 12, 577. [Google Scholar] [CrossRef] [Green Version]
  21. Zeng, C.-H.; Meng, X.-T.; Xu, S.-S.; Han, L.-J.; Zhong, S.; Jia, M.-Y. A polymorphic lanthanide complex as selective Co2+ sensor and luminescent timer. Sens. Actuators B Chem. 2015, 221, 127–135. [Google Scholar] [CrossRef]
  22. Du, J.-L.; Zhang, X.-Y.; Li, C.-P.; Gao, J.-P.; Hou, J.-X.; Jing, X.; Mu, Y.-J.; Li, L.-J. A bi-functional luminescent Zn(II)-MOF for detection of nitroaromatic explosives and Fe3+ ions. Sens. Actuators B Chem. 2018, 257, 207–213. [Google Scholar] [CrossRef]
  23. Chaudhary, S.; Kumar, S.; Umar, A.; Singh, J.; Rawat, M.; Mehta, S. Europium-doped gadolinium oxide nanoparticles: A potential photoluminescencent probe for highly selective and sensitive detection of Fe3+ and Cr3+ ions. Sens. Actuators B Chem. 2017, 243, 579–588. [Google Scholar] [CrossRef]
  24. Matsushita, A.F.; Filho, C.M.; Piñeiro, M.; Pais, A.A.C.C.; Valente, A.J.M. Effect of Eu(III) and Tb(III) chloride on the gelification behavior of poly(sodium acrylate). J. Mol. Liq. 2018, 264, 205–214. [Google Scholar] [CrossRef]
  25. Qi, X.; Wang, Z.; Ma, S.; Wu, L.; Yang, S.; Xu, J. Complexation behavior of poly(acrylic acid) and lanthanide ions. Polymers 2014, 55, 1183–1189. [Google Scholar] [CrossRef]
  26. Chen, C.; Zhang, X.; Gao, P.; Hu, M. A water stable europium coordination polymer as fluorescent sensor for detecting Fe3+, CrO42−, and Cr2O72− ions. J. Solid State Chem. 2018, 258, 86–92. [Google Scholar] [CrossRef]
  27. Tang, K.; Ma, Q.; Zhan, Q.; Wang, Q. An intelligent copper(II) luminescent sensor using europium narrow emissions based on titania hybrid material. Opt. Mater. 2014, 36, 1520–1524. [Google Scholar] [CrossRef]
  28. Bogale, R.F.; Chen, Y.; Ye, J.; Yang, Y.; Rauf, A.; Duan, L.; Tian, P.; Ning, G. Highly selective and sensitive detection of 4-nitrophenol and Fe3+ ion based on a luminescent layered terbium (III) coordination polymer. Sens. Actuators B Chem. 2017, 245, 171–178. [Google Scholar] [CrossRef]
  29. Yang, W.; Feng, J.; Zhang, H. Facile and rapid fabrication of nanostructured lanthanide coordination polymers as selective luminescent probes in aqueous solution. J. Mater. Chem. 2012, 22, 6819. [Google Scholar] [CrossRef]
  30. Costa, D.; Burrows, H.D.; Miguel, M. Changes in Hydration of Lanthanide Ions on Binding to DNA in Aqueous Solution. Langmuir 2005, 21, 10492–10496. [Google Scholar] [CrossRef] [Green Version]
  31. Horrocks, W.D.; Sudnick, D.R. Lanthanide ion probes of structure in biology. Laser-induced luminescence decay constants provide a direct measure of the number of metal-coordinated water molecules. J. Am. Chem. Soc. 1979, 101, 334–340. [Google Scholar] [CrossRef]
  32. Horrocks, W.D.; Schmidt, G.F.; Sudnick, D.R.; Kittrell, C.; Bernheim, R.A. Laser-induced lanthanide ion luminescence lifetime measurements by direct excitation of metal ion levels. A new class of structural probe for calcium-binding proteins and nucleic acids. J. Am. Chem. Soc. 1977, 99, 2378–2380. [Google Scholar] [CrossRef] [PubMed]
  33. Tapia, M.J.; Burrows, H.D. Cation Polyelectrolyte Interactions in Aqueous Sodium Poly(vinyl sulfonate) as Seen by Ce3+ to Tb3+ Energy Transfer. Langmuir 2002, 18, 1872–1876. [Google Scholar] [CrossRef] [Green Version]
  34. Tapia, M.J.; Burrows, H.D.; Azenha, M.; Miguel, M.; Pais, A.A.C.C.; Sarraguca, J. Cation Association with Sodium Dodecyl Sulfate Micelles As Seen by Lanthanide Luminescence. J. Phys. Chem. B 2002, 106, 6966–6972. [Google Scholar] [CrossRef] [Green Version]
  35. Wang, Y.; Xin, X.; Li, W.; Jia, C.; Wang, L.; Shen, J.; Xu, G.-Y. Studies on the gel behavior and luminescence properties of biological surfactant sodium deoxycholate/rare-earth salts mixed systems. J. Colloid Interface Sci. 2014, 431, 82–89. [Google Scholar] [CrossRef] [PubMed]
  36. Dong, F.; Wang, J.; Wang, Y.; Ren, S. Synthesis and humidity controlling properties of halloysite/poly(sodium acrylate-acrylamide) composite. J. Mater. Chem. 2012, 22, 11093. [Google Scholar] [CrossRef]
  37. Adachi, T.; Shirotani, I.; Hayashi, J.; Shimomura, O. Phase transitions of lanthanide monophosphides with NaCl-type structure at high pressures. Phys. Lett. A 1998, 250, 389–393. [Google Scholar] [CrossRef]
  38. Jin, H.; Huang, Y.; Jian, J. Plate-like Cr2O3 for highly selective sensing of nitric oxide. Sens. Actuators B Chem. 2015, 206, 107–110. [Google Scholar] [CrossRef]
  39. Zhang, F.; Wang, Z.; Chu, T.; Li, W.; Yang, Y. A facile fabrication of electrodeposited luminescent MOF thin film for selective and recyclable sensing nitroaromatic explosives. Analyst 2016, 141, 4502–4510. [Google Scholar] [CrossRef]
  40. Aulsebrook, M.L.; Graham, A.P.B.; Grace, M.R.; Tuck, K.L. Lanthanide complexes for luminescence-based sensing of low molecular weight analytes. Coord. Chem. Rev. 2018, 375, 191–220. [Google Scholar] [CrossRef]
  41. Hao, Z.; Song, X.; Zhu, M.; Meng, X.; Zhao, S.; Su, S.; Yang, W.; Song, S.; Zhang, H. One-dimensional channel-structured Eu-MOF for sensing small organic molecules and Cu2+ ion. J. Mater. Chem. A 2013, 1, 11043. [Google Scholar] [CrossRef]
  42. Hao, Z.; Yang, G.; Song, X.; Zhu, M.; Meng, X.; Zhao, S.; Song, S.; Zhang, H. A europium(iii) based metal–organic framework: Bifunctional properties related to sensing and electronic conductivity. J. Mater. Chem. A 2014, 2, 237–244. [Google Scholar] [CrossRef]
  43. Xiao, Y.; Cui, Y.; Zheng, Q.; Xiang, S. A microporous luminescent metal–organic framework for highly selective and sensitive sensing of Cu2+ in aqueous solution. ChemComm 2010, 56, 5503–5505. [Google Scholar] [CrossRef] [PubMed]
  44. Su, R.; Gao, J.; Deng, S.; Zhang, R.; Zheng, Y. Dual-target optical sensors assembled by lanthanide complex incorporated sol–gel-derived polymeric films. J. Sol Gel Sci. Technol. 2016, 78, 606–612. [Google Scholar] [CrossRef]
  45. Bogale, R.F.; Ye, J.; Sun, Y.; Sun, T.; Zhang, S.; Rauf, A.; Hang, C.; Tian, P.; Ning, G. Highly selective and sensitive detection of metal ions and nitroaromatic compounds by an anionic europium(III) coordination polymer. Dalton Trans. 2016, 45, 11137–11144. [Google Scholar] [CrossRef]
  46. Bogale, R.F.; Chen, Y.; Ye, J.; Zhang, S.; Li, Y.; Liu, X.; Zheng, T.; Rauf, A.; Ning, G. A terbium(III)-based coordination polymer for selective and sensitive sensing of nitroaromatics and ferric ion: Synthesis, crystal structure and photoluminescence properties. New J. Chem. 2017, 41, 12713–12720. [Google Scholar] [CrossRef]
  47. Liu, W.; Wang, Y.; Song, L.; Silver, M.A.; Xie, J.; Zhang, L.; Chen, L.; Diwu, J.; Chai, Z.; Wang, S. Efficient and selective sensing of Cu2+ and UO22+ by a europium metal-organic framework. Talanta 2019, 196, 515–522. [Google Scholar] [CrossRef]
  48. Tan, H.; Zhang, Y.; Chen, Y. Detection of mercury ions (Hg2+) in urine using a terbium chelate fluorescent probe. Sens. Actuators B Chem. 2011, 156, 120–125. [Google Scholar] [CrossRef]
  49. Chen, B.; Wang, L.; Xiao, Y.; Fronczek, F.R.; Xue, M.; Cui, Y.; Qian, G. A Luminescent Metal-Organic Framework with Lewis Basic Pyridyl Sites for the Sensing of Metal Ions. Angew. Chem. Int. Ed. 2009, 48, 500–503. [Google Scholar] [CrossRef]
  50. Du, P.-Y.; Gu, W.; Liu, X. Multifunctional Three-Dimensional Europium Metal–Organic Framework for Luminescence Sensing of Benzaldehyde and Cu2+ and Selective Capture of Dye Molecules. Inorg. Chem. 2016, 55, 7826–7828. [Google Scholar] [CrossRef]
  51. Sun, Z.; Li, H.; Sun, G.; Guo, J.; Ma, Y.; Li, L. Design and construction of lanthanide metal-organic frameworks through mixed-ligand strategy: Sensing property of acetone and Cu2+. Inorg. Chim. Acta 2018, 469, 51–56. [Google Scholar] [CrossRef]
  52. Wang, Z.; Liu, H.; Wang, S.; Rao, Z.; Yang, Y. A luminescent Terbium-Succinate MOF thin film fabricated by electrodeposition for sensing of Cu2+ in aqueous environment. Sens. Actuators B Chem. 2015, 220, 779–787. [Google Scholar] [CrossRef]
Figure 1. Excitation spectra (left side) and emission spectra (right side) of (a) EuCl3 and Eu3+/PSA (λem = 616 nm and λex = 395 nm) and (b) TbCl3 and Tb3+/PSA (λem = 545 nm and λex = 273 nm) in aqueous solutions at pH 7.2.
Figure 1. Excitation spectra (left side) and emission spectra (right side) of (a) EuCl3 and Eu3+/PSA (λem = 616 nm and λex = 395 nm) and (b) TbCl3 and Tb3+/PSA (λem = 545 nm and λex = 273 nm) in aqueous solutions at pH 7.2.
Polymers 12 01314 g001
Figure 2. XRD patterns of PSA (black), Eu3+/PSA (red) and Tb3+/PSA (green), (left) and SEM micrograph of Eu3+/PSA freeze drying composite (right). Magnification 5000×.
Figure 2. XRD patterns of PSA (black), Eu3+/PSA (red) and Tb3+/PSA (green), (left) and SEM micrograph of Eu3+/PSA freeze drying composite (right). Magnification 5000×.
Polymers 12 01314 g002
Figure 3. Quenching efficiency of (a) Eu3+/PSA (λem = 616 nm) and (b) Tb3+/PSA (λem = 545 nm) with different metal ions in aqueous solution.
Figure 3. Quenching efficiency of (a) Eu3+/PSA (λem = 616 nm) and (b) Tb3+/PSA (λem = 545 nm) with different metal ions in aqueous solution.
Polymers 12 01314 g003
Figure 4. Fluorescence photographs of Eu3+/PSA (red) and Tb3+/PSA (green) alone (first to the left in each row), and in the presence of 3.33 mM of metal ions under UV light (365 nm) in aqueous solution.
Figure 4. Fluorescence photographs of Eu3+/PSA (red) and Tb3+/PSA (green) alone (first to the left in each row), and in the presence of 3.33 mM of metal ions under UV light (365 nm) in aqueous solution.
Polymers 12 01314 g004
Figure 5. Quenching efficiency of (a) Eu3+/PSA (λem = 616 nm) and (b) Tb3+/PSA (λem = 545 nm) with different Na+ salts in aqueous solution.
Figure 5. Quenching efficiency of (a) Eu3+/PSA (λem = 616 nm) and (b) Tb3+/PSA (λem = 545 nm) with different Na+ salts in aqueous solution.
Polymers 12 01314 g005
Figure 6. Stern-Volmer plot of Eu3+/PSA and Tb3+/PSA lifetimes with different Cu2+ concentrations.
Figure 6. Stern-Volmer plot of Eu3+/PSA and Tb3+/PSA lifetimes with different Cu2+ concentrations.
Polymers 12 01314 g006
Figure 7. Emission intensities of (a) Eu3+/PSA (5D07F2) transition, (b) Tb3+/PSA (5D47F5) transition and Stern-Volmer plots of (c) Eu3+/PSA and (d) Tb3+/PSA with lower concentrations of Cu2+.
Figure 7. Emission intensities of (a) Eu3+/PSA (5D07F2) transition, (b) Tb3+/PSA (5D47F5) transition and Stern-Volmer plots of (c) Eu3+/PSA and (d) Tb3+/PSA with lower concentrations of Cu2+.
Polymers 12 01314 g007
Figure 8. (a) SEM micrography, (b) EDX spectra and elemental maps of Eu3+/PSA after contact with Cu2+. The right-hand figures represent the combined elemental maps; for the sake of clarity, the elemental maps of counterions (Na and Cl) are represented in a separated figure.
Figure 8. (a) SEM micrography, (b) EDX spectra and elemental maps of Eu3+/PSA after contact with Cu2+. The right-hand figures represent the combined elemental maps; for the sake of clarity, the elemental maps of counterions (Na and Cl) are represented in a separated figure.
Polymers 12 01314 g008
Table 1. Lifetimes (τ), decay constant (k) and number of coordinated water molecules (n) of Eu3+ and Tb3+ in water and D2O and in poly(sodium acrylate) (PSA) aqueous and D2O solutions.
Table 1. Lifetimes (τ), decay constant (k) and number of coordinated water molecules (n) of Eu3+ and Tb3+ in water and D2O and in poly(sodium acrylate) (PSA) aqueous and D2O solutions.
SampleSolventτ (μs)k (ms−1)n
(H2O Coordinated)
EuCl3H2O1128.929
EuCl3D2O17750.56
Eu3+/PSAH2O2743.653
Eu3+/PSAD2O16020.62
TbCl3H2O4312.329
TbCl3D2O35650.28
Tb3+/PSAH2O8611.163
Tb3+/PSAD2O30020.33
Table 2. Stern-Volmer constant for the quenching of the emission of several lanthanoid-based sensors by Cu2+.
Table 2. Stern-Volmer constant for the quenching of the emission of several lanthanoid-based sensors by Cu2+.
Luminescent MaterialKsv (M−1)Reference
[Eu(pdc)1.5(dmf)]·(DMF)0.5(H2O)0.589[49]
Eu3+/PSA580This Work
{[Eu2(abtc)1.5(H2O)3(DMA)]·H2O·DMA}n529[50]
{[Eu(HL)(L)(H2O)2]_2H2O}n116[43]
{[Eu(L)(ox)0.5(H2O)2]_H2O}n2074[51]
Tb3+/PSA5655This work
Tb-SA6298[52]

Share and Cite

MDPI and ACS Style

Matsushita, A.F.Y.; Tapia, M.J.; Pais, A.A.C.C.; Valente, A.J.M. Luminescent Properties of Lanthanoid-Poly(Sodium Acrylate) Composites: Insights on the Interaction Mechanism. Polymers 2020, 12, 1314. https://doi.org/10.3390/polym12061314

AMA Style

Matsushita AFY, Tapia MJ, Pais AACC, Valente AJM. Luminescent Properties of Lanthanoid-Poly(Sodium Acrylate) Composites: Insights on the Interaction Mechanism. Polymers. 2020; 12(6):1314. https://doi.org/10.3390/polym12061314

Chicago/Turabian Style

Matsushita, Alan F. Y., María José Tapia, Alberto A. C. C. Pais, and Artur J. M. Valente. 2020. "Luminescent Properties of Lanthanoid-Poly(Sodium Acrylate) Composites: Insights on the Interaction Mechanism" Polymers 12, no. 6: 1314. https://doi.org/10.3390/polym12061314

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop