Open access peer-reviewed chapter

Luminescent Materials with Turn-on and Ratiometric Sensory Response Based on Coordination Compounds of Lanthanides

Written By

Claudio Pettinari, Andrei Drozdov and Yuriy Belousov

Submitted: 14 November 2022 Reviewed: 28 November 2022 Published: 03 January 2023

DOI: 10.5772/intechopen.109189

From the Edited Volume

Rare Earth Elements - Emerging Advances, Technology Utilization, and Resource Procurement

Edited by Michael T. Aide

Chapter metrics overview

123 Chapter Downloads

View Full Metrics

Abstract

Luminescent lanthanide complexes serve as a unique set of tools for creating sensory materials. The most significant types of sensory response in such materials are the turn-on/off response, when the analyte causes an increase or decrease in the emission intensity, respectively, as well as the ratiometric response, which manifests itself as a change in the ratio of luminescence intensities at different wavelengths. In this paper, we consider two of the most technologically advanced types of luminescent sensor materials based on lanthanide compounds—“turn on” and ratiometric sensors. The production of such materials is not only of importance per their possible application but is especially interesting from a fundamental point of view, since their design requires the implementation of non-trivial solutions.

Keywords

  • lanthanides
  • coordination compounds
  • luminescence
  • sensors
  • sensory response mechanism
  • turn-on sensors
  • ratiometric sensors

1. Introduction

The lanthanide (Ln) family, due to the features of the electron shell, forms a large number of complexes with a set of luminescent properties that is unique among all non-radioactive elements of the Periodic Table. The lanthanide compounds combine effective luminescence in the UV, visible, or near-IR ranges with narrow spectral lines and constant wavelengths. In addition, in the case of lanthanides, it is especially convenient to drive and control luminescence, which opens the way to the creation of new emerging technologies in a chemical sensorics. In the accompanying paper, we have described the principle of luminescence materials with turn-off sensory response, and here we complete our work by proponing both “turn on” and “ratiometric” sensors. This manuscript also concentrates on the role of elements of the Lanthanide Series in this emerging technology.

Advertisement

2. “Turn on” sensors

There are few publications on “turn on” materials with respect to “turn off” sensors, due to several factors: the rarity of processes accompanied by an increase in luminescence requires a rational design of materials, and at the same time, impurities in analytes can cause quenching processes that level the effect of luminescence enhancement. However, examples of materials with such a response have been growing in recent years, and the described mechanisms can serve as an inspiration for new researchers.

Like quenching, luminescence enhancement can be described by the Stern-Volmer equations, but the definition of the constants is given only in a small number of papers.

The most common “turn on” response mechanisms are:

  1. The analyte acts as an antenna, its coordination leads to an increase in the luminescence of lanthanide (Direct Analyte Antenna Function, DAAF).

  2. An analyte (anionic or neutral) does not directly act as an antenna, but displaces quencher molecules (usually H2O) from the lanthanide coordination sphere (solvent-quencher displacement, SQD).

  3. The analyte, without destroying the molecular structure of the complex, forms or destroys weak bonds (usually hydrogen), reducing the effectiveness of vibrational quenching control (VQC).

  4. The analyte leads to partial destruction of the material, removing the quenching fragment not directly bound to the Ln3+ ion (quenching fragment removal, QFR).

  5. The analyte suppresses luminescence quenching by PET, FRET, or DEE mechanisms (energy transfer quenching control, ETQC).

  6. The analyte affects the electronic structure of the sensor by changing the position of the triplet level (triplet level control, TLC), the effectiveness of sensitization (sensitization control, SC), changing the absorption and excitation spectra (excitation wavelength control, EWC).

An analysis of the literature shows that the SQD strategy (Figure 1) is the most often implemented. In this case, the organic solvent molecules or other ligands that do not have their own absorption in the same excitation region of the complex displace water molecules from the coordination sphere of lanthanide. Quenching of Ln3+-centered luminescence through interactions with OH, CH, and NH bonds is caused by the dissipation of the energy of the Ln3+ excited state into high-energy stretching vibrations of several neighboring molecules [1, 2]. The efficiency of vibrational quenching on these bonds depends on the energy of the excited state Ln3+ and the number of vibrational modes of the X-H bond that cover this energy (Figure 2).

Figure 1.

Substitution of the solvent quencher by the analyte (SQD: solvent-quencher displacement).

Figure 2.

Nonradiative deactivation of excited states of terbium and europium by multiphonon relaxation on O-H and O-D bonds.

For europium ions the required number of OH bonds vibration modes (4) is less than for terbium ions (5), which determines a much more efficient quenching of the luminescence of Eu3+. In the transition to O-D bonds, the number of modes for both ions increases (6 and 5, respectively). As a result, the observed lifetime of Eu3+(aq) is shorter than that of Tb3+(aq), and when the medium is replaced by D2O, the difference is noticeably leveled (see Part 1, Table 1). Even greater is the effect of vibrations of these bonds on the emission of IR-emitting ions. The difference in τobs of various REE ions in protic and deuterated water allows to estimate the number of water molecules in the near coordination sphere of lanthanide according to empirical formulas qLn=k×1τH2O1τD2Ob, where qLn is the number of water molecules in the coordination sphere of lanthanide, τH2O and τD2O are the observed lifetimes of excited REE states in water and D2O. k and b parameters are given in Table 1. This makes it possible to evaluate the difference in sensitivity for materials based on ions with lower slope values (Nd3+, Eu3+, and Yb3+) and larger ones (Tb3+, Sm3+, and Dy3+). Most of the papers are devoted specifically to europium derivatives, which is not accidental. Neodymium compounds have been studied as sensor materials with less SQD response, but appear to be very promising [7].

Ln3+Slope (k)Intercept (b)Ref.
Eu1.110.31[3, 4]
Eu1.050.44[5]
Tb5.00.06[3]
Tb4.030.87[5]
Sm2.540.37[5]
Dy2.110.60[5]
Nd0.3581.97[6]
Yb1.00.20[3]

Table 1.

Slope (k) and intercept (b) in the Ln3+ hydration number formulae.

An interesting example of this type of sensor is presented in [8], which shows the selectivity of the response with respect to methanol against the background of ethanol and propanol-1. The effect is due to the lock-and-key matching of the channel diameters in the MOF sensor structure with the sizes of the indicated alcohol molecules.

The QFR mechanism (Figure 3) has been often implemented using not-luminescent copper-lanthanide heterometallic complexes capable of “donating” copper ions under the action of various analytes, especially those containing sulfur [9, 10, 11, 12] and nitrogen [12, 13, 14] donor atoms with a high affinity for copper ions. This approach makes it possible to achieve greater selectivity with respect to background ions not strongly interacting with Cu2+ ions. In other cases, an analyte with strong oxidizing (ClO, [15]) or reductive (ascorbic acid, [16]) power removes the quencher fragment. Other analytes having the same redox properties are expected to show a similar effect.

Figure 3.

The quencher fragment removing mechanism (QFR).

The interaction of some analytes with the sensor material leads to a change in the electronic structure, which increases the efficiency of sensitization of lanthanide ions. The detailed mechanism of such enhancement is difficult to determine, but a suppression of the non-radiative relaxation of the singlet and triplet states of the ligands [17], an increase in the energy transfer constant from the ligand to the metal, and a change in the position of the triplet level [18], which affects the efficiency of the reverse transfer, can also make a contribution. These sensor materials showed a response with respect to s- and p-metal cations, as they generally contain a crown-containing fragment [17] or suitable coordination sites determining the material selectivity. The response to gases has been studied much less frequently than the response to analytes in solution; a paper [18] describing the turn-on sensor for NO2 is of particular interest, two luminescent Eu and Tb complexes being investigated: the Tb complex shows that a reversible sorption of NO2 leads to a “turn on” response, while for europium, a “turn off” response is observed. The interaction between the sensor with the analyte leads to an increase in the energy of the triplet level by ∼260 cm−1, which is favorable for energy transfer to a higher resonance level of terbium, but reduces the efficiency for the low-lying level of europium (Figure 4).

Figure 4.

Jabłoński diagram featuring the ground singlet (S0), first excited singlet (S1), and lowest triplet (T1) states of the ligand together with the relevant atomic levels of Gd3+,Tb3+, and Eu3+. Values for the energy levels are given in cm−1. Represented with a permission from Ref. [18].

A sensor material with a positive luminescent response to Cu2+ ions has been recently described [19]. Upon addition of up to 4 equivalents of 3d-metal salts (especially Cu2+) an intense band appears in the excitation spectrum associated with intraligand energy transfer. The appearance of this band makes the luminescence excitation more efficient when using the corresponding wavelength, which leads to a more than twofold increase in the quantum yield of Eu3+. This work is a unique example of a turn-on sensor for a d-metal cation.

Finally, the direct antenna function of the analyte (Figure 5) can take place if the analyte contains a suitable conjugate system [20, 21]. A not-high selectivity has been showed by such sensors, but the use of a well-defined excitation wavelength, coinciding with the absorption of the analyte, can increase it (Table 2).

Figure 5.

Direct antenna analyte function mechanism (DAAF).

MaterialAnalyteMediaLinearity rangeLODResponse timeMechanismRef.
Cu-Eu or Cu-Tb heterometallic MOFS2−H2O; HeLa cells0.01 μM8–12 minQFR (Cu2+)[9]
Cu-Tb heterometallic MOFS2−H2O0.13 μMQFR (Cu2+)[10]
Hybrid material Eu3+ complex-covered laponiteGlutathioneH2O0.5–30 μM0.162 μM10 minQFR (Cu2+)[11]
Hybrid material: Tb3+ covered Cu2+-BTC MOFamyloid β-peptideH2O1–550 nM
5–490 nM
0.3 nMQFR (Cu2+)[13]
Composite material: carbon quantum dots covered with a H2DPA. Eu3+ and Cu2+H2DPAH2O0–10 μM0.137 μMQFR (Cu2+)[12]
Tb@duoble hydroxid Ni-FeH2DPAH2O0–12 μM0.36 μMSQD+DAAF[22]
Eu@ duoble hydroxid Ni-FeH2DPAH2O0–12 μM1.03 μMSQD+DAAF
(Cu2+@Tb-MOFs)Uric acidH2O0–104 μM0.650μMQFR (Cu2+)[14]
Tb-MOFDMSOH2O1.0–100%1. 1.68%90sSQD[23]
Eu-MOFMeOHCH2Cl2<0.03 vol.%SQD[8]
Eu-complex nanoparticlesCipro-floxacinH2O1–40 μM780 nM15 minSQD[24]
Eu-MOFDiethyl phosphorochloridateCH3CN15–90 μM<2 minSQD[25]
Eu-MOFAMFH2O2 μMSQD[26]
nanoparticles [Tb(Cit)(H2O)]Guanosine-5-mono-phosphateH2O0.15–20 μM0.1 μMSQD[27]
Tb(HCOO)3l-kynurenineH2O + 5%DMF1–10 μM1 nMSQD[28]
Eu-MOFNH3CH3NH2Air(1) 5–350 ppm
(2) 0.7–15.5ppm
500–600 sSQD[29]
Eu-MOFNH3Air>28ppm100–250 sSQD[30]
Nd-MOFF-CH3CNVQC[7]
Eu(L)-(UIO-67)ClOH2O0.1–5 μM16nM5 sQFR[15]
Composite material: Eu3+ complex in a PMMA membrane reacted with a KMnO4Ascorbic acidH2O0–35 μM48 nM1 minQFR[16]
Eu-MOFZn2+MeOH2–8 mMVQC[31]
Eu-MOFCd2+H2O2 hVQC[32]
EuL
TbL
EuL
TbL
ML = crown-containing ligands
K+MeOH0–10eq
0–10eq
0–2eq
0–1.7eq
ETQC[17]
Tb-Zn heterometallic MOFMg2+DMF10−5–3·10−3 mM1.38·10−5 M
KSV = 1.81·103
n./a.
probably SC
[33]
Eu-MOFCd2+H2O0.2–2.2 μM1.1761×107n./a.
probably SC
[34]
Hybrid material: Eu3+ covered Zr-MOF Uio-66(Zr)–(BTEC-COOH)2Cd2+H2O0500 μM0.06 μM>2 minSC[35]
Tb-MOFPb2+EtOH10−8–10−3 MSC[36]
Eu-MOFPb2+H2O0–0.1 mM KSV=2970 M−18.22 μMn/a[37]
Tb-MOFNO21–5 ppm2 ppm1 minTLC[18]
Eu-MOFCu2+CH3CNEWC[19]
[Tb-MOFTryptophanH2O pH=42.5 10−5–2.5 10−41 μM2 hDAAF[20]
Nanoparticles of Na[Gd0.88Tb0.12F4]DopamineH2O0–10 μM47 nM5 minDAAF[21]
[Eu-MOFAl3+DMFn/a[38]
Tb3+-functionalized COFOchratoxin AH2O0–10 μM13.5 nM10 sDAAF+SQD[39]
Tb3+ MOFAcetyl-acetoneH2O0–1750 ppm0.129 ppmDAAF[40]
Aspartic acidH2O0–25 μM0.025 ppmDAAF[40]

Table 2.

Turn-on sensors.

To determine the correct mechanism of the observed “turn on” response, we propose the following algorithm, which is also relevant for the rational design of such materials:

  1. A chemical analysis of the “analyte + sensory material” system to answer the question “whether a new chemical compound is formed” during their interaction.

  2. A study in the case of a positive reply on the elimination of molecules or quenching ions (water, transition metals cations, etc.) occurs. In the case of a positive response, a mechanism such as SQD, QFR, and DAAF can be assumed. The usually observed increase in the τobs of lanthanide ion is evident in favor of such an assumption. In the case of vibrational quenching, one should keep in mind the significant temperature and isotope sensitivity of the phonon relaxation efficiency, so measurements of τobs in a deuterated solvent and/or upon cooling are the crucial experiment to confirm the SQD mechanism.

  3. An IR study, in the case of a preserved molecular structure of the sensor to verify the breaking or formation of weak bonds (hydrogen, etc.), which could support a VQC mechanism.

  4. A study confirming that a new ligand (usually an analyte) is coordinated to the sensor and is capable to perform an antenna function. The DAAF mechanism is, in fact, confirmed by the combination of analyte absorption and sensor excitation spectra in the presence and absence of the analyte, as well as measurements of τobs of Ln3+.

  5. A measure of the energy of the triplet level from the phosphorescence spectra of the Gd3+ complex in the presence of a chemical interaction between the sensor and the analyte. The change in the energy of the triplet level indicates the implementation of a relatively rare TLC mechanism. Similarly, the appearance of a new intense band in the excitation spectrum makes it possible to implement the EWC mechanism.

  6. A combined analysis of both absorption and excitation spectra of the sensor material, which can be used to prove the ETQC mechanism. For a more accurate determination of the latter, a calculation of the energy of the orbitals is required.

Advertisement

3. “Ratiometric” sensors

The third important sensor type is the “ratiometric” one. In materials showing these properties, the signal is estimated as the ratio between luminescence values at two different wavelengths, either affecting the effect of Ln3+ ions or affecting both the Ln3+ and the organic ligand.

Sensory ResonseS=I1I2=λ1minλ1maxIdλλ2minλ2maxIdλ,E1

where λ1,2(min/max) -are the initial and final coordinates of the bands in the spectrum involved in the integration.

“Ratiometric” sensors do not have the disadvantage to determine the presence of internal standards as in the case of “turn off” and “turn on” sensors. An unusual uniqueness of ratiometric sensory visibility materials is the naked eye color change of luminescent properties when they are in presence of an analyte.

The possible response mechanisms are similar to the “turn off” and “turn on” signals considered above. In addition, if two different Ln3+ ions are simultaneously present in a compound (Ln-Ln’ sensors), energy transfer between them is possible, the efficiency of which depends on the Ln-Ln distances and other factors [41]; the analyte can affect the transfer efficiency constant. This type of response is defined as Metal-to-Metal Energy Transfer (MMET).

Ratiometric materials generally use the Eu3+/Tb3+ bimetallic pair, since these ions have the most efficient luminescence. Other lanthanide-based systems were used only in a small number of cases, for example, Ce3+/Tb3+ [42], Dy3+/Eu3+ [43], and Eu3+/Yb3+ [44] systems. To achieve greater accuracy, as a rule, it is better to employ the most intense lines in the emission spectra, for example, those corresponding to the europium 5D0-7F2 (∼612 nm) and terbium 5D4-7F5 (∼544 nm) transitions, also if this is not entirely correct, because the europium transition 5D0-7F2 is partially superimposed on the low-intensity transition of terbium 5D4-7F5 (∼620 nm). The sensor response can be calculated using the europium 5D0-7F4 (∼700 nm) transition, which in many cases is also very intense and does not overlap with any terbium line, but lies in the region of reduced sensitivity of common photomultipliers [45]. The usual molar fractions ratio of lanthanides with a predominance of terbium ions is caused by the transfer of energy between lanthanide ions.

O2 in the triplet state can act as a luminescence quencher [46, 47]. The emission of Tb3+ ions is quenched by O2 molecules more efficiently than by Eu3+ due to the smaller difference in the energies of 3T level of O2 and the excited state of corresponding Ln3+ ions. This principle is the basis for a sensor for gaseous oxygen, which is a mechanical mixture of terbium and europium complexes immobilized on a quartz surface [48].

The possibility to determine, using “ratiometric sensors”, small H2O impurities against the background of organic solvents (for which Karl Fischer titration is usually used), which requires the use of toxic reagents and “capricious” equipment, is of great interest. In addition, “ratiometric” sensors can detect the mixture of light water in and D2O, which is impossible with Fischer titration and requires expensive mass spectrometers or precision IR spectrometers [49, 50]. Such materials are based on MOF structures containing intra-sphere water molecules. The sensor material is activated by heat treatment in vacuum, after which the obtained anhydrous powder is dispersed in the medium of the investigated solvent. This approach is not accidental: polymeric MOFs are insoluble in most organic solvents, which makes them easy to regenerate and reuse. OH quenching is the basis for the detection of methanol in ethanol and in the form of vapors in air [51], and a similar principle can be extended to CH oscillations in an elegant DMSO impurity sensor in deuterated DMSO-d6 [44]. Substitution of an OH group [45] or a water molecule [52] in the coordination sphere of lanthanide with F ions also suppresses Eu3+ quenching more effectively than Tb3+, which was used in the development of fluoride-sensitive sensors.

Mixed-metal MOFs with a statistical distribution of Ln3+ ions are commonly used as material for “ratiometric” sensors. However, Tscelykh et al proposed in [53] to use solutions of pentafluorobenzoates or even Ln3+ chlorides, since the sensitivity of the sensor is directly dependent on the number of water molecules in the environment of Eu3+. The disadvantages of this approach include the complexity of material regeneration and contamination of the analyzed media, which makes flow analysis impossible and increases its cost.

Moisture sensors are closely related to pH sensor materials. “Ratiometric” pH sensors based on carboxylate MOFs have been also considered [54, 55, 56]. In the first two cases, as the pH increases, the color of the luminescence changes from green to red; the relative intensity of Eu3+ emission increases. This is explained by the strong pH dependence of the excitation transfer rate constant from Tb3+ to Eu3+, which was confirmed by measurements of τobs in monometallic and bimetallic complexes [55] (in the analysis of kinetic data, certain caution is required due to the non-monoexponential nature of the Eu3+ decay curve in the presence of luminescence sensitization by Ln ions [41, 50]). It has been described [56] that as the pH increases, the luminescence color, on the contrary, changes from red to green, that is, the relative intensity of Tb3+ luminescence increases. It was shown by the DFT method that the energy of the 3T* triplet level increases from ∼24.400 to ∼26.400 cm−1 upon transition from the protonated to the deprotonated form of the ligand. The resonant level of europium is much lower than both of these values, so the sensitization of europium luminescence (17.200 cm−1) does not change its efficiency (τobs(Eu) also changes slightly with pH). The resonant level of terbium (20.500 cm−1) lies higher, and an increase in the triplet level energy weakens the back transfer of excitation from Tb3+ to the ligand. This leads to a significant increase in τobs(Tb) and in the intensity of Tb3+ emission.

The MMET mechanism can be further confirmed by measuring the response in a bimetallic complex and in a mechanical mixture of complexes of two REEs. The decrease or disappearance of the response in the second case indicates the partial or complete participation of the MMET mechanism. A similar approach was used [57] where the MMET mechanism is related to the response to hydrosulfide anions. Interestingly, the other two analyzed analytes (THF and Ag+ cations) exhibit different response mechanisms.

The highly selective sensor for a potassium ions, as in the case of the “turn-on” material described above [13], contains diaza-18-crown-6 in the structure, which makes it selective to the tested s-metal cations [58]. It has been shown that the capture of the potassium ion by the crown fragment leads to an increase in the triplet level energy from ∼22.400 cm−1 to ∼23.400 cm−1. This effect enhances the emission of Eu3+ to a greater extent than that of Tb3+. In addition, it was found that the efficiency of energy transfer from Tb3+ to a Eu3+ increases with increasing K+ concentration, which leads to an increase in the contribution of europium bands in the spectrum.

For Ln-Ln’ “ratiometric” sensors, in contrast to the “turn off” and “turn on” materials considered above, the response to conjugated organic analytes has been poorly studied. The sensory response of Eu-Tb bimetallic BTC-MOF film to a number of drugs [59] has been also investigated. The nature of the emission of this material changes significantly in the presence of a number of compounds, in particular, in the presence of coumarin and caffeine. The nature of the response to caffeine is not discussed, and for coumarin, an increase in kLnET is assumed, which is confirmed by kinetic measurements, but the corresponding quantum chemical studies are not provided.

In the presence of only one lanthanide luminescent center, the “ratiometric” response can still be realized if the ligand is fluorescent or phosphorescent (Ln-L sensors). This is possible if the antenna sensitization efficiency is low, which can be caused by too large or too small energy gap between 3T* and the REE resonant level, as well as by a large metal-ligand distance. Possible mechanisms for the occurrence of this type of sensory response include binding of the lanthanide ion caused by decomposition of the parent complex, which weakens the efficiency of antenna sensitization (lanthanide ion ejection, LIE), and direct antenna function of the analyte.

The detection of Hg2+ ions was possible by using a composite material containing a luminescent terbium coordination polymer impregnated with a coumarin solution [60]. The response is selective to Hg2+ ions with respect to a wide range of s-, p-, and d-metal ions and is associated with the displacement of Tb3+ ions from the adenosine monophosphate environment, leading to a decrease in the Tb3+ luminescence intensity and to a weakening of its sensitization by coumarin. The driving force behind the displacement reaction of the lanthanide ion into a complex with a lower luminescence intensity can be not only the strength of the “cationic analyte-ligand” bond, but also the formation of a stable “analyte-lanthanide cation” complex. A similar strategy was implemented for phosphate ions [61] and for alkaline phosphatase [62]. In some cases, the response occurs when the ligand is destroyed, for example, in sensors for the determination of HClO [63] or formaldehyde [64]. The vulnerability of these two strategies is the impossibility of a sensor regeneration.

Radiative relaxation by the PET or FRET process can also lead to a response if the analyte has the appropriate LUMO energy. This approach was implemented in a sensor material for the detection of nitrofuranose and furazolidone against the background of other antibiotics in a dysprosium-containing sensor [65]. Trinitrophenol (TNP) similarly blocks the sensitization of terbium luminescence in specifically designed Tb-MOF via PET and FRET mechanisms [66].

As found for Ln-L sensors, the response manifests itself as a drop in the relative efficiency of lanthanide luminescence. The opposite is also possible if the ligand displaces quencher molecules (SQD) and exhibits antenna properties itself (DAAF). An example of this approach is in a study [67, 68, 69] describing the response to dipicolinic acid, an important biological marker associated with antrax disease.

Bisphenol-A, an important industrial reagent in the production of plastics, can act as an effective antenna ligand for europium cations. This underlies the importance of the production of a highly sensitive sensor material based on a composite of carbon quantum dots and europium 5’-adenosine monophosphate [70].

Finally, a response can also be caused by an internal filter effect (IFE). This effect can be the main mechanism, as in the sensor for tetracyclines [71], or manifest itself simultaneously with quenching through ET mechanisms in sensors for Fe3+ [72] or for the aforementioned nitrofuranose and furazolidone (Table 3) [65].

MaterialAnalyteMediaResponse wavelengths λ1, λ2 (nm)Linearity rangeLODMechanismRef.
Ln-Ln’ materials
[Eu0.167Tb0.833 (L)(H2O)3] (1,4-dioxane) MOFH2OD2O
dioxane DMF
543Tb-616Eu10–120000 ppmSQD[49]
{[Eu0.1Tb0.9(L) (H2O)3] MOFH2OD2O615Eu-545Tb0.5–100%SQD[50]
dioxane0–100%.SQD
CH3CN0–60%.SQD
Tb0.9711Eu0.0289 (L)0.5 MOFH2OCH3CN543Tb-615Eu0–2.5 v%SQD[73]
{[(Eu0.02Dy0.98)2
(L)2]·2NH2 (CH3)2·H2O} MOF
H2OEtOH416 Dy-614Eu0–10%0.1%SQD[43]
[Tb0.3Eu0.7](L) MOFH2O (vapors)Air619Eu-546Tb20–100%4.3%SQD[74]
10Eu(C6F5COO)3+1Tb (C6F5COO)3 solutionH2OD2O612 Eu, 545 Tb0–50%SQD[53]
Eu0.022Tb0.978 MOFH2OEtOH544Tb-614Eu0–0.67%0.016%SQD[75]
Eu0.2Tb0.8MOFH2OEtOH546Tb-616Eu10–100%SQD[76]
Eu0.2Tb0.8MOFH2OEtOH546Tb-616Eu0–10%0.01%SQD
[Eu0.05Tb0.95(L) (H2O)Cl] MOFpH, H2ODMF547Tb-616EupH 3–11
0–0.8 v% H2O
SQD[54]
Eu0.034Tb0.966(L)2(C2O4)(H2O)4 MOFpHH2O618Eu-545TbpH 3–7SQD[55]
Eu0.205Tb0.795(L)(H2O)4·0.5H2O MOFpHH2O616Eu-543TbpH 4.0–7.5SQD[56]
Eu0.4Tb0.6 Ciprofloxacin complex + dextran aldehyde (DEX) + chitosan hydrogelpHH2O615Eu-545TbpH 5.5–8.0SQD[77]
{(Me2NH2)[Tb0.9
Eu0.1(L)2]·(L76
(H2O)1.5}n MOF
MeOHEtOH, air545Tb-618Eu0–1 v%SQD[51]
Eu0.1Yb0.9(dbm)3x(BPhen) solutionDMSODMSO-d6612Eu-975Yb0–50 v%SQD[44]
[Eu0.1Tb1.9(FDC)3(DMF)2]·2DMFCH3NHCODMF544Tb-614Eu0–100 v%[78]
[Eu0.5Tb1.5(FDC)3]ethylene glycol (EG)EG-dioxane mixture546Tb-616Eu0–100 v%SQD[79]
Hybrid material: mixture of solid Ln(L) complexes (Ln=Tb and Eu) immobilized on glass surfaceO2Air/N2546Tb-616Eu0–0.2 atmET quenching[48]
EuTbL complex with a crown-containing ligandK+MeOH545Tb-617Eu0.5–6 eq K+/LnTLC[58]
[Eu0.47Tb0.53(L) (H2O)3] MOFStyrene (vapors)Air545Tb-617EuPET+FRET quencing[80]
7:1 Tb:Eu complex with a 3,5-di-carboxy-benzene boronic acidCiprofloxacinH2O545Tb-615Eu0.3–24 μM90 nMDAAF[81]
[Eu0.5Tb0.5L(H2O)]Cl MOFF-H2O700Eu-545Tb0–2 eq F-/Ln17.7 nMSQD[45]
[Tb0.97Eu0.03(L) (H2O)] MOFF547Tb-617Eu0–1.9 ppm96 ppbSQD[52]
Hybrid material: luminescent complex [TbL] MOF enclosed in SiO2 nanoparticles, surface coated with luminescent europium complex Eu(L)HClOEtOH539Tb-607Eu0–40 μM0.27 μMLigand destruction[63]
Eu0.05Tb0.95BTC0.9(L)0.1 MOFCrO42−545Tb-618Eu1μMIFE+ET[82]
Tb0.6Eu0.4-MOFHg2+H2O617Eu-543Tb0–40 μM4.83 nMMMET[83]
Hybrid material: complex Eu4Tb6(acac) on the surface of NZL zeoliteEt3N, tBuNH2,
n-BuNH2, BnNH2, en
545Tb-618EuSC[84]
[La0.88Eu0.02Tb0.1 (L)(DMF)2]n ·H2O.0.5DMF MOFHS-DMF/ H2O544Tb-616Eu0–400 μMMMET[57]
THFDMF0–8%TLC
Ag+DMF/ H2OTLC
Tb0.833Eu0.167(L77)3 MOFH2DPAH2O549Tb-620Eu2–16 μM96 nMMMET+DAAF[85]
[Tb0.833Eu0.167BTC] .6H2O MOFH2DPAH2O545Tb-613Eu0–700 nM4.55 nMMMET+DAAF[86]
[Eu0.06Tb0.04Gd0.9 (L)1.5(H2O) (DMF)] MOFMnO4H2O613Eu-544Tb0–0.2 mM0.02 μMIFE+FRET[87]
Ce3+/Tb3+ bimetallic guanosine monophosphate complexalkaline phosphataseH2O552Tb-384Ce0.2–60 mU mL−10.12 mU/mLLigand destruction[42]
[Eu0.1Tb0.9BTC(H2O)2].3DMF.H2O MOFCoumarin613Eu-544Tb1–100 μM2.3 μMMMET[59]
Caffeine613Eu-544Tb1–100 μMn/a
Ln-L materials
Hybrid material: cerium/terbium adenosine monophosphate impregnated with coumarin solutionHg2+H2O445–548Tb0.08–1000 nM0.03 nMLIE[60]
Hybrid material: CsPbBr3@Eu-BTC MOFHg2+H2O520L-616Eu0–1 μM0.116 nMPET[88]
Hybrid material: Eu3+@ Ca-MOFHg2+H2O381L-590Eu0.02–200 μM2.6 nMTLC[89]
{[Eu(L)(H2O)1.35(DMF)0.65]·1.9DMF}n MOFPO43−H2O614Eu-368L0.1–15 μM52 nMLIE[61]
Solution: Eu3+ complex with H2DPA + luminolPO43−H2O615Eu—423L0.5–50 μM0.12 μMQFR (luminol)[90]
Eu-2′-amino- [1,1′:4′,1″-terphenyl]-3,3″,5,5″-tetracarboxylic acid MOFF-H2O615Eu-397L0−5120 μM11.26 μMTLC[91]
Composite material: Tb3+ guanosine 5’-disodium@Cu-nanoclustersalkaline phosphataseH2O545Tb-425L0.002–2 U mL−10.002 U mL−1LIE[62]
Eu0.7Gd0.3(L)1.5 (Phen).H2O complexFe3+H2O614Eu-415L0–60 μM91 nMIFE+ET[72]
Composite material Eu0.7Gd0.3(tfbdc)1.5 (Phen).H2O complex +FeCl3Ascorbic acid0–60 μM0.184 μMQFR (Fe3+)
Composite material Zn-MOF impregnated with a TbCl3 solutionH2DPAH2O544Tb-330L0–1 μM3.6 nMSQD+DAAF[67]
EuCl3 + sodium polycarylate + pyranine +H2DPAH2O615Eu-510L0–90 μM10.3 nMSQD+DAAF[68]
Eu3+@Zr-MOF UiO-66-(COOH)2-NH2H2DPAH2O621Eu-453L0–40 μM25 nMSQD+DAAF[69]
[Tb2(L2)1.5(NMP)2]n MOFTNPH2O546Tb-400L0–50 μM0.11 μMFRET and PET quenching[66]
[Tb0.02Gd1.98L1.5 (NMP)2]n
MOF
H2O0–80 μM0.41 μMFRET and PET quenching
Composite material: Eu MOF in a PMMA matrixForm-aldehydeH2O, Air453L-616Eu0.05–1%Ligand destruction[64]
[Dy(L9)(DMF)3]n MOFNitro-furanoseH2O385L-573Dy0–60 μM47.6 nMPET+IFE[65]
FurazolidoneH2O0–60 μM48.2 nMPET+IFE
[NMe2H2]2[Tb93-OH)82-OH)3 (H2O)3(L)6] ·11DMF·23H2O MOFTetra-cyclineH2O543Tb-345L0.14–23 μM0.18nMIFE[71]
Eu3+ complex with a luminol (LML) and guanosine 50-monophosphate (GMP) nanoparticlesTetra-cyclineH2O617Eu-430L0–60 μM23.2 nMIFE+ DAAF[92]
Eu-adenosine-monophosphate@CDBisphenol-AH2O429L-623Eu0–100 μM20nMDAAF[70]
Hybdrid material: Tb@Bi-based MOFSerotoninH2O545Tb-350L0–200 μM0.57 μMIFE[93]
Eu3+-functionalized hydrogen-bonded organic frameworkCH3NH2H2O425L-615Eu10−8–10−2 M0.87 ppmIFE[94]
[Eu(DMTP-DC)1.5(H2O)3]·DMF MOFarginineH2O441L-617Eu0–250 μM24.38 μMPET[95]
lysineH2O441L-617Eu0–250 μM9.31 μMPET
Eu3+@Al-based MOFβ-glucuro nidaseH2O450L-617Eu0.1–50 U L−10.03 U L−1IFE[96]

Table 3.

“Ratiometric” sensors.

As in the previous cases, the key stages in studying the mechanism of the “ratiometric” sensory response are the chemical analysis of the reaction product between the sensor and analyte; comparison of the excitation and emission spectra of the sensor and the absorption of the analyte; kinetic studies of excited state lifetimes and quantum chemical modeling. In general, the“ratiometric” response mechanisms usually coincide with those for“turn off” and“turn on” systems. An exception is the mechanism associated with changing the MMET constant (kLnET). This mechanism must be reliably established by kinetic measurements, as well as by studying the response in a mechanical mixture of complexes of two lanthanides.

Advertisement

4. Conclusions

The classification of sensory response mechanisms reported in this review, although perhaps not complete, allows us to successfully classify most of the cited works and make appropriate generalizations, moving from a descriptive style to a debatable one.

The reported examples show the significant progress in the field of lanthanide-based luminescent sensors achieved in recent years. A wide variety of analytes can be qualitatively or quantitatively determined using suitable lanthanide compounds, which requires rational system design. “Turn off” sensors can have a niche application in the analysis of nitroaromatic compounds, while more popular “Turn on” and “ratiometric” materials can be produced using fairly simple strategies: different quenching efficiency of various REE ions by bond vibrations, binding and removal of a quenching fragment by a suitable analyte (e.g., binding Cu2+ ions with sulfur-containing ligands), etc.

Progress has touched not only the field of materials design, but also a reliable determination of the sensory response mechanism, which requires several spectroscopic and kinetic measurements, and in some cases quantum chemical calculations of orbital energies. The number of papers containing these studies has been increasing in recent years, and we hope that just such a systematic approach will become the standard in the future works.

Advertisement

Acknowledgments

University of Camerino is gratefully acknowledged.

Advertisement

Conflict of interest

The authors declare no conflict of interest.

References

  1. 1. Bünzli JCG, Eliseeva SV. Basics of lanthanide photophysics. In: Hänninen P, Härmä H, editors. Lanthanide Luminescence. Springer Series on Fluorescence. Vol. 7. Berlin, Heidelberg: Springer; 2010. DOI: 10.1007/4243_2010_3
  2. 2. Belousov YA, Korshunov VM, Metlin MT, Metlina DA, Kiskin MA, Aminev DF, et al. Towards bright dysprosium emitters: Single and combined effects of environmental symmetry, deuteration, and gadolinium dilution. Dyes and Pigments. 2022;199:110078. DOI: 10.1016/j.dyepig.2021.110078
  3. 3. Beeby A, Clarkson IM, Dickins RS, Faulkner S, Parker D, Royle L, et al. Non-radiative deactivation of the excited states of europium, terbium and ytterbium complexes by proximate energy-matched OH, NH and CH oscillators: An improved luminescence method for establishing solution hydration states. Journal of the Chemical Society, Perkin Transactions. 1999;2(2):493-503. DOI: 10.1039/a808692c
  4. 4. Supkowski RM, Horrocks WDW. On the determination of the number of water molecules, q, coordinated to europium(III) ions in solution from luminescence decay lifetimes. Inorganica Chimica Acta. 2002;340:44-48. DOI: 10.1016/S0020-1693(02)01022-8
  5. 5. Kimura T, Kato Y. Luminescence study on hydration states of lanthanide(III)-polyaminopolycarboxylate complexes in aqueous solution. Journal of Alloys and Compounds. 1998;275–277:806-810. DOI: 10.1016/S0925-8388(98)00446-0
  6. 6. Kimura T, Kato Y. Luminescence study on determination of the inner-sphere hydration number of Am(III) and Nd(III). Journal of Alloys and Compounds. 1998;271–273:867-871. DOI: 10.1016/S0925-8388(98)00236-9
  7. 7. Chen W, Tang X, Dou W, Wang B, Guo L, Ju Z, et al. The construction of homochiral lanthanide quadruple-stranded helicates with multiresponsive sensing properties toward fluoride anions. Chemistry—A European Journal. 2017;23:9804-9811. DOI: 10.1002/chem.201700827
  8. 8. Wang J, Jiang M, Yan L, Peng R, Huangfu M, Guo X, et al. Multifunctional luminescent Eu(III)-based metal-organic framework for sensing methanol and detection and adsorption of Fe(III) ions in aqueous solution. Inorganic Chemistry. 2016;55:12660-12668. DOI: 10.1021/acs.inorgchem.6b01815
  9. 9. Yao Y, Delgado-Rivera L, Samareh Afsari H, Yin L, Thatcher GRJ, Moore TW, et al. Time-gated luminescence detection of enzymatically produced hydrogen sulfide: Design, synthesis, and application of a lanthanide-based probe. Inorganic Chemistry. 2018;57:681-688. DOI: 10.1021/acs.inorgchem.7b02533
  10. 10. Aulsebrook ML, Biswas S, Leaver FM, Grace MR, Graham B, Barrios AM, et al. A Luminogenic lanthanide-based probe for the highly selective detection of nanomolar sulfide levels in aqueous samples. Chemical Communications. 2017;53:4911-4914. DOI: 10.1039/c7cc01764b
  11. 11. Chen X, Wang Y, Chai R, Xu Y, Li H, Liu B. Luminescent lanthanide-based organic/inorganic hybrid materials for discrimination of glutathione in solution and within hydrogels. ACS Applied Materials & Interfaces. 2017;9:13554-13563. DOI: 10.1021/acsami.7b02679
  12. 12. Li X, Gao J, Rao S, Zheng Y. Development of a selective “on-off-on” nano-sensor based on lanthanide encapsulated carbon dots. Synthetic Metals. 2017;231:107-111. DOI: 10.1016/j.synthmet.2017.07.006
  13. 13. Liu B, Shen H, Hao Y, Zhu X, Li S, Huang Y, et al. Lanthanide functionalized metal–organic coordination polymer: Toward novel turn-on fluorescent sensing of amyloid β-peptide. Analytical Chemistry. 2018;90:12449-12455. DOI: 10.1021/acs.analchem.8b01546
  14. 14. Yang J, Che J, Jiang X, Fan Y, Gao D, Bi J, et al. A novel turn-on fluorescence probe based on Cu(II) functionalized metal–organic frameworks for visual detection of uric acid. Molecules. 2022;27:4803. DOI: 10.3390/molecules27154803
  15. 15. Zhou Z, Li X, Tang Y, Zhang CC, Fu H, Wu N, et al. Oxidative deoximation reaction induced recognition of hypochlorite based on a new fluorescent lanthanide-organic framework. Chemical Engineering Journal. 2018;351:364-370. DOI: 10.1016/j.cej.2018.06.123
  16. 16. Sun T, Hao S, Fan R, Zhang J, Chen W, Zhu K, et al. In situ self-assembled cationic lanthanide metal organic framework membrane sensor for effective MnO4− and ascorbic acid detection. Analytica Chimica Acta. 2021;1142:211-220. DOI: 10.1016/j.aca.2020.10.062
  17. 17. Junker AKR, Tropiano M, Faulkner S, Sørensen TJ. Kinetically inert lanthanide complexes as reporter groups for binding of potassium by 18-Crown-6. Inorganic Chemistry. 2016;55:12299-12308. DOI: 10.1021/acs.inorgchem.6b02063
  18. 18. Gamonal A, Sun C, Mariano AL, Fernandez-Bartolome E, Guerrero-Sanvicente E, Vlaisavljevich B, et al. Divergent adsorption-dependent luminescence of amino-functionalized lanthanide metal-organic frameworks for highly sensitive NO2Sensors. Journal of Physical Chemistry Letters. 2020;11:3362-3368. DOI: 10.1021/acs.jpclett.0c00457
  19. 19. Liu CL, Zhang RL, Lin CS, Zhou LP, Cai LX, Kong JT, et al. Intraligand charge transfer sensitization on self-assembled europium tetrahedral cage leads to dual-selective luminescent sensing toward anion and cation. Journal of the American Chemical Society. 2017;139:12474-12479. DOI: 10.1021/jacs.7b05157
  20. 20. Li SS, Ye ZN, Xu SS, Zhang YJ, Tao AR, Liu M, et al. Highly luminescent lanthanide CPs based on dinuclear cluster: Crystal structure and sensitive Trp sensor. RSC Advances. 2015;5:71961-71967. DOI: 10.1039/c5ra12149c
  21. 21. Ling X, Shi R, Zhang J, Liu D, Weng M, Zhang C, et al. Dual-signal luminescent detection of dopamine by a single type of lanthanide-doped nanoparticles. ACS Sensors. 2018;3:1683-1689. DOI: 10.1021/acssensors.8b00368
  22. 22. Niu X, Wang M, Cao R, Zhang M, Liu Z, Liu Z, et al. Ion exchange fabrication of lanthanide functionalized layered double hydroxides microcapsules for rapid and visual detection of anthrax biomarker. Spectrochimica Acta. 2022;281:121622. DOI: 10.1016/j.saa.2022.121622
  23. 23. Yang D, Lu L, Feng S, Zhu M. First Ln-MOF as a trifunctional luminescent probe for the efficient sensing of aspartic acid, Fe3+and DMSO. Dalton Transactions. 2020;49:7514-7524. DOI: 10.1039/d0dt00938e
  24. 24. Liu B, Huang Y, Shen Q, Zhu X, Hao Y, Qu P, et al. Turn-on fluorescence detection of ciprofloxacin in tablets based on lanthanide coordination polymer nanoparticles. RSC Advances. 2016;6:100743-100747. DOI: 10.1039/c6ra20357d
  25. 25. Gupta K, Patra AK. Luminescent Europium(III) “Turn-On” sensor for G-series chemical warfare simulants: A mechanistic investigation. ACS Sensors. 2020;5:1268-1272. DOI: 10.1021/acssensors.9b02552
  26. 26. Sahoo J, Arunachalam R, Subramanian PS, Suresh E, Valkonen A, Rissanen K, et al. Coordinatively unsaturated Lanthanide(III) helicates: Luminescence sensors for adenosine monophosphate in aqueous media. Angewandte Chemie. 2016;128:9777-9781. DOI: 10.1002/ange.201604093
  27. 27. Liu B, Shen H, Liu D, Hao Y, Zhu X, Shen Q, et al. Citrate/Tb lanthanide coordination polymer nanoparticles: Preparation and sensing of Guanosine-5-Monophosphate. Sensors and Actuators B: Chemical. 2019;300:126879. DOI: 10.1016/j.snb.2019.126879
  28. 28. Tang T, Liu M, Chen Z, Wang X, Lai C, Ding L, et al. Highly sensitive luminescent lanthanide metal–organic framework sensor for L-kynurenine. Journal of Rare Earths. 2021;40:415-420. DOI: 10.1016/j.jre.2021.02.008
  29. 29. Sergeev AA, Voznesenskiy SS, Petrochenkova NV, Shishov AS, Leonov AA, Emelina TB, et al. Luminescent chemosensors for amines and ammonia based on Eu(III) chelate complexes. In: Proc. SPIE 10176, Asia-Pacific Conference on Fundamental Problems of Opto- and Microelectronics. 14 Dec 2016. p. 1017610. DOI: 10.1117/12.2268131
  30. 30. Voznesenskiy SS, Sergeev AA, Mirochnik AG, Leonov AA, Petrochenkova NV, Shishov AS, et al. Specific features of europium tris-benzoylacetonate sensor response to gaseous ammonia. Sensors and Actuators B: Chemical. 2017;246:46-52. DOI: 10.1016/j.snb.2017.02.034
  31. 31. Feng X, Li R, Guo N, Sun Y, Ng SW, Liu X, et al. Two unique hydroxyl bridged lanthanide polymers incorporating mixed carboxylate ligands: Syntheses, structures, luminescence and magnetic property. Inorganica Chimica Acta. 2017;459:87-94. DOI: 10.1016/j.ica.2017.01.028
  32. 32. Zheng K, Liu Z, Jiang Y, Guo P, Li H, Zeng C, et al. Ultrahigh luminescence quantum yield lanthanide coordination polymer as a multifunctional sensor. Dalton Transactions. 2018;47:17432-17440. DOI: 10.1039/c8dt03832e
  33. 33. Feng X, Feng Y, Guo N, Sun Y, Zhang T, Ma L, et al. Series D-f heteronuclear metal-organic frameworks: Color tunability and luminescent probe with switchable properties. Inorganic Chemistry. 2017;56:1713-1721. DOI: 10.1021/acs.inorgchem.6b02851
  34. 34. Sun YL, Feng X, Guo N, Wang LY, Li RF, Bai RF. A novel europium coordination polymer based on mixed carboxylic acid ligands: Synthesis, structure and luminescence. Inorganic Chemistry Communication. 2016;67:90-94. DOI: 10.1016/j.inoche.2016.03.006
  35. 35. Hao JN, Yan B. A water-stable lanthanide-functionalized MOF as a highly selective and sensitive fluorescent probe for Cd2+. Chemical Communications. 2015;51:7737-7740. DOI: 10.1039/c5cc01430a
  36. 36. Wang KM, Du L, Ma YL, Zhao JS, Wang Q, Yan T, et al. Multifunctional chemical sensors and luminescent thermometers based on lanthanide metal-organic framework materials. CrystEngComm. 2016;18:2690-2700. DOI: 10.1039/c5ce02367j
  37. 37. Li L, Chen Q, Niu Z, Zhou X, Yang T, Huang W. Lanthanide metal-organic frameworks assembled from a fluorene-based ligand: Selective sensing of Pb2+ and Fe3+ ions. Journal of Materials Chemistry C. 2016;4:1900-1905. DOI: 10.1039/c5tc04320d
  38. 38. Liu X, Lin H, Xiao Z, Fan W, Huang A, Wang R, et al. Multifunctional lanthanide-organic frameworks for fluorescent sensing, gas separation and catalysis. Dalton Transactions. 2016;45:3743-3749. DOI: 10.1039/c5dt04339e
  39. 39. Yu Y, Li G. Design of Terbium(III)-functionalized covalent organic framework as a selective and sensitive turn-on fluorescent switch for ochratoxin A monitoring. Journal of Hazardous Materials. 2022;422:126927. DOI: 10.1016/j.jhazmat.2021.126927
  40. 40. Li YG, Hu JJ, Zhang JL, Liu SJ, Peng Y, Wen HR. Lanthanide-based metal-organic framework materials as bifunctional fluorescence sensors toward acetylacetone and aspartic acid. CrystEngComm. 2022;24:2464-2471. DOI: 10.1039/d2ce00174h
  41. 41. Gontcharenko VE, Kiskin MA, Dolzhenko VD, Korshunov VM, Taydakov IV, Belousov YA. Mono- and mixed metal complexes of Eu3+, Gd3+, and Tb3+ with a diketone, bearing pyrazole moiety and Chf2-group: Structure, color tuning, and kinetics of energy transfer between lanthanide ions. Molecules. 2021;26:1-16. DOI: 10.3390/molecules26092655
  42. 42. Li J, Gong CC, Li Z, Yao R, Huo P, Deng B, et al. A self-assembly lanthanide nanoparticle for ratiometric fluorescence determination of alkaline phosphatase activity. Journal of Photochemistry and Photobiology A: Chemistry. 2022;432:114054. DOI: 10.1016/j.jphotochem.2022.114054
  43. 43. Li H, Han W, Lv R, Zhai A, Li XL, Gu W, et al. Dual-function mixed-lanthanide metal-organic framework for ratiometric water detection in bioethanol and temperature sensing. Analytical Chemistry. 2019;91:2148-2154. DOI: 10.1021/acs.analchem.8b04690
  44. 44. Kornikov AI, Kozlov MI, Lepnev LS, Utochnikova VV. Europium-ytterbium bimetallic complex-based sensor for detecting DMSO impurities in DMSO-D6. Sensors and Actuators B: Chemical. 2022;370:132432. DOI: 10.1016/j.snb.2022.132432
  45. 45. Nonat A, Liu T, Jeannin O, Camerel F, Charbonnière LJ. Energy transfer in supramolecular heteronuclear lanthanide dimers and application to fluoride sensing in water. Chemistry—A European Journal. 2018;24:3784-3792. DOI: 10.1002/chem.201705532.+
  46. 46. Lakowicz JR. Principles of Fluorescence Spectroscopy. In: Lakowicz JR, editor. Boston, MA: Springer US; 2006
  47. 47. Nakai H, Seo J, Kitagawa K, Goto T, Matsumoto T, Ogo S. An oxygen-sensitive luminescent Dy(III) complex. Dalton Transactions. 2016;45:9492-9496. DOI: 10.1039/c6dt01057a
  48. 48. Lehr J, Tropiano M, Beer PD, Faulkner S, Davis JJ. Ratiometric oxygen sensing using lanthanide luminescent emitting interfaces. Chemical Communications. 2015;51:15944-15947. DOI: 10.1039/c5cc05738h
  49. 49. Dunning SG, Nuñez AJ, Moore MD, Steiner A, Lynch VM, Sessler JL, et al. A sensor for trace H2O detection in D2O. Chem. 2017;2:579-589. DOI: 10.1016/j.chempr.2017.02.010
  50. 50. Gontcharenko VE, Lunev AM, Taydakov IV, Korshunov VM, Drozdov AA, Belousov YA. Luminescent lanthanide-based sensor for H2O detection in aprotic solvents and D2O. IEEE Sensors Journal. 2019;19:7365-7372. DOI: 10.1109/JSEN.2019.2916498
  51. 51. Chen D, Sun C, Peng Y, Zhang N, Si H, Liu C, et al. Ratiometric fluorescence sensing and colorimetric decoding methanol by a bimetallic lanthanide-organic framework. Sensors and Actuators B: Chemical. 2018;265:104-109. DOI: 10.1016/j.snb.2018.03.028
  52. 52. Zeng X, Hu J, Zhang M, Wang F, Wu L, Hou X. Visual detection of fluoride anions using mixed lanthanide metal-organic frameworks with a smartphone. Analytical Chemistry. 2020;92:2097-2102. DOI: 10.1021/acs.analchem.9b04598
  53. 53. Tcelykh L, Kozhevnikova Khudoleeva V, Goloveshkin A, Lepnev L, Popelensky T, Utochnikova V. Sensing of H2O in D2O: Is there an easy way? Analyst. 2020;145:759-763. DOI: 10.1039/c9an02023c
  54. 54. Li H, Liu B, Xu L, Jiao H. A hetero- MOF-based bifunctional ratiometric fluorescence sensor for PH and water detection. Dalton Transactions. 2021;50:143-150. DOI: 10.1039/D0DT03626A
  55. 55. Xia T, Zhu F, Jiang K, Cui Y, Yang Y, Qian G. A luminescent ratiometric PH sensor based on a nanoscale and biocompatible Eu/Tb-mixed MOF. Dalton Transactions. 2017;46:7549-7555. DOI: 10.1039/c7dt01604b
  56. 56. Xia T, Cui Y, Yang Y, Qian G. Highly stable mixed-lanthanide metal-organic frameworks for self-referencing and colorimetric luminescent PH sensing. ChemNanoMat. 2017;3:51-57. DOI: 10.1002/cnma.201600331
  57. 57. Wang XY, Yao X, Huang Q, Li YX, An GH, Li GM. Triple-wavelength-region luminescence sensing based on a color-tunable emitting lanthanide metal organic framework. Analytical Chemistry. 2018;90:6675-6682. DOI: 10.1021/acs.analchem.8b00494
  58. 58. Yang D, Wang Y, Liu D, Li Z, Li H. Luminescence modulation: Via cation-π interaction in a lanthanide assembly: Implications for potassium detection. Journal of Materials Chemistry C. 2018;6:1944-1950. DOI: 10.1039/c7tc04580h
  59. 59. Gao Y, Yu G, Liu K, Wang B. Luminescent mixed-crystal Ln-MOF thin film for the recognition and detection of pharmaceuticals. Sensors and Actuators B: Chemical. 2018;257:931-935. DOI: 10.1016/j.snb.2017.10.180
  60. 60. Zhang Z, Wu Y, He S, Xu Y, Li G, Ye B. Ratiometric fluorescence sensing of mercuric ion based on dye-doped lanthanide coordination polymer particles. Analytica Chimica Acta. 2018;1014:85-90. DOI: 10.1016/j.aca.2018.01.065
  61. 61. Cheng Y, Zhang H, Yang B, Wu J, Wang Y, Ding B, et al. Highly efficient fluorescence sensing of phosphate by dual-emissive lanthanide MOFs. Dalton Transactions. 2018;47:12273-12283. DOI: 10.1039/C8DT01515E
  62. 62. Li X, Wang X, Guo W, Wang Y, Hua Q, Tang F, et al. Selective detection of alkaline phosphatase activity in environmental water samples by copper nanoclusters doped lanthanide coordination polymer nanocomposites as the ratiometric fluorescent probe. Biosensors. 2022;12. DOI: 10.3390/bios12060372
  63. 63. Ma H, Song B, Wang Y, Cong D, Jiang Y, Yuan J. Dual-emissive nanoarchitecture of lanthanide-complex-modified silica particles for in vivo ratiometric time-gated luminescence imaging of hypochlorous acid. Chemical Science. 2016;8:150-159. DOI: 10.1039/C6SC02243J
  64. 64. Wang Y, Zhang G, Zhang F, Chu T, Yang Y. A novel lanthanide MOF thin film: The highly performance self-calibrating luminescent sensor for detecting formaldehyde as an illegal preservative in aquatic product. Sensors and Actuators B: Chemical. 2017;251:667-673. DOI: 10.1016/j.snb.2017.05.063
  65. 65. Wu S, Zhu M, Zhang Y, Kosinova M, Fedin VP, Gao E. A water-stable lanthanide coordination polymer as multicenter platform for ratiometric luminescent sensing antibiotics. Chemistry—A European Journal. 2020;26:3137-3144. DOI: 10.1002/chem.201905027
  66. 66. Li L, Cheng J, Liu Z, Song L, You Y, Zhou X, et al. Ratiometric luminescent sensor of picric acid based on the dual-emission mixed-lanthanide coordination polymer. ACS Applied Materials & Interfaces. 2018;10:44109-44115. DOI: 10.1021/acsami.8b13719
  67. 67. Zhang D, Zhou Y, Cuan J, Gan N. A lanthanide functionalized MOF hybrid for ratiometric luminescence detection of an anthrax biomarker. CrystEngComm. 2018;20:1264-1270. DOI: 10.1039/c7ce01994g
  68. 68. Liu X, Li B, Xu Y, Li Z, Zhang Y, Ding ZJ, et al. A highly selective lanthanide-containing probe for ratiometric luminescence detection of an anthrax biomarker. Dalton Transactions. 2019;48:7714-7719. DOI: 10.1039/c9dt01477b
  69. 69. Huo P, Li Z, Yao R, Deng Y, Gong C, Zhang D, et al. Dual-ligand lanthanide metal–organic framework for ratiometric fluorescence detection of the anthrax biomarker dipicolinic acid. Spectrochimica Acta. 2022;282:121700. DOI: 10.1016/j.saa.2022.121700
  70. 70. Li Y, Min Q, Wang Y, Zhuang X, Hao X, Tian C, et al. A portable visual coffee ring based on carbon dot sensitized lanthanide complex coordination to detect Bisphenol A in water. RSC Advances. 2022;12:7306-7312. DOI: 10.1039/d2ra00039c
  71. 71. Li R, Wang W, El-Sayed ESM, Su K, He P, Yuan D. Ratiometric fluorescence detection of tetracycline antibiotic based on a polynuclear lanthanide metal–organic framework. Sensors and Actuators B: Chemical. 2021;330. DOI: 10.1016/j.snb.2020.129314
  72. 72. Yu H, Liu Q, Li J, Su ZM, Li X, Wang X, et al. A dual-emitting mixed-lanthanide MOF with high water-stability for ratiometric fluorescence sensing of Fe3+and ascorbic acid. Journal of Materials Chemistry C. 2021;9:562-568. DOI: 10.1039/d0tc04781c
  73. 73. Li B, Wang W, Hong Z, El-Sayed ESM, Yuan D. Ratiometric fluorescence detection of trace water in an organic solvent based on bimetallic lanthanide metal-organic frameworks. Chemical Communications. 2019;55:6926-6929. DOI: 10.1039/c9cc02324k
  74. 74. Xia D, Li J, Li W, Jiang L, Li G. Lanthanides-based multifunctional luminescent films for ratiometric humidity sensing, information storage, and colored coating. Journal of Luminescence. 2021;231:117784. DOI: 10.1016/j.jlumin.2020.117784
  75. 75. Zhai X, Feng P, Song N, Zhao G, Liu Q, Liu L, et al. Dual-functional ratiometric fluorescent sensor based on mixed-lanthanide metal-organic frameworks for the detection of trace water and temperature. Inorganic Chemistry Frontiers. 2022;9:1406-1415. DOI: 10.1039/d2qi00093h
  76. 76. Yu L, Zheng Q, Xiong L, Feng L, Xiao Y. Dual-lanthanide urea metal-organic framework based fluorescent traffic light microsensor for solvent decoding and visual trace water assay. Sensors and Actuators B: Chemical. 2022;356:131328. DOI: 10.1016/j.snb.2021.131328
  77. 77. Zhou Q, Dong X, Zhang B, Zhang X, Ou K, Wang Q, et al. Naked-eye sensing and target-guiding treatment of bacterial infection using PH-tunable multicolor luminescent lanthanide-based hydrogel. Journal of Colloid and Interface Science. 2022;610:731-740. DOI: 10.1016/j.jcis.2021.11.121
  78. 78. Qin SJ, Qu XL, Yan B. A self-calibrating bimetallic lanthanide metal-organic luminescent sensor integrated with logic gate operation for detecting N-methylformamide. Inorganic Chemistry Frontiers. 2018;5:2971-2977. DOI: 10.1039/c8qi00958a
  79. 79. Zhou J, Li H, Zhang H, Li H, Shi W, Cheng P. A bimetallic lanthanide metal-organic material as a self-calibrating color-gradient luminescent sensor. Advanced Materials. 2015;27:7072-7077. DOI: 10.1002/adma.201502760
  80. 80. Su Y, Yu J, Li Y, Phua SFZ, Liu G, Lim WQ, et al. Versatile bimetallic lanthanide metal-organic frameworks for tunable emission and efficient fluorescence sensing. Commun. Chem. 2018;1. DOI: 10.1038/s42004-018-0016-0
  81. 81. Sun Y, Dramou P, Song Z, Zheng L, Zhang X, Ni X, et al. Lanthanide metal doped organic gel as ratiometric fluorescence probe for selective monitoring of ciprofloxacin. Microchemical Journal. 2022;179:107476. DOI: 10.1016/j.microc.2022.107476
  82. 82. Wu RZ, Yang X, Zhang LW, Zhou PP. Luminescent lanthanide metal-organic frameworks for chemical sensing and toxic anion detection. Dalton Transactions. 2017;46:9859-9867. DOI: 10.1039/c7dt01790a
  83. 83. Wang X, Jiang Z, Yang C, Zhen S, Huang C, Li Y. Facile synthesis of binary two-dimensional lanthanide metal-organic framework nanosheets for ratiometric fluorescence detection of mercury ions. Journal of Hazardous Materials. 2022;423:126978. DOI: 10.1016/j.jhazmat.2021.126978
  84. 84. Li P, Li H. Amine vapor responsive lanthanide complex entrapment: Control of the ligand-to-metal and metal-to-metal energy transfer. Journal of Materials Chemistry C. 2016;4:2165-2169. DOI: 10.1039/c5tc04377h
  85. 85. Gao N, Zhang Y, Huang P, Xiang Z, Wu FY, Mao L. Perturbing Tandem energy transfer in luminescent heterobinuclear lanthanide coordination polymer nanoparticles enables real-time monitoring of release of the anthrax biomarker from bacterial spores. Analytical Chemistry. 2018;90:7004-7011. DOI: 10.1021/acs.analchem.8b01365
  86. 86. Zhang Y, Li B, Ma H, Zhang L, Jiang H, Song H, et al. A nanoscaled lanthanide metal-organic framework as a colorimetric fluorescence sensor for dipicolinic acid based on modulating energy transfer. Journal of Materials Chemistry C. 2016;4:7294-7301. DOI: 10.1039/c6tc01022a
  87. 87. Dong ZP, Zhao F, Zhang L, Liu ZL, Wang YQ. A white-light-emitting lanthanide metal-organic framework for luminescence turn-off sensing of MnO4-and turn-on sensing of folic acid and construction of a “Turn-on plus” system. New Journal of Chemistry. 2020;44:10239-10249. DOI: 10.1039/d0nj02145h
  88. 88. Shu Y, Ye Q, Dai T, Guan J, Ji Z, Xu Q, et al. Incorporation of perovskite nanocrystals into lanthanide metal-organic frameworks with enhanced stability for ratiometric and visual sensing of mercury in aqueous solution. Journal of Hazardous Materials. 2022;430:128360. DOI: 10.1016/j.jhazmat.2022.128360
  89. 89. Hao Guo NW, Peng L, Chen Y, Liu Y, Li C, Zhang H, et al. A novel ratiometric fluorescence sensor based on lanthanide-functionalized MOF for Hg2+ detection. Talanta. 2022;250:123710. DOI: 10.1016/j.talanta.2022.123710
  90. 90. Wu H, Ling Y, Ju S, Chen Y, Xu M, Tang Y. A smartphone-integrated light-up lanthanide fluorescent probe for the visual and ratiometric detection of total phosphorus in human urine and environmental water samples. Spectrochimica Acta. 2022;2022:279. DOI: 10.1016/j.saa.2022.121360
  91. 91. Yu K, Wang Q, Xiang W, Li Z, He Y, Zhao D. Amino-functionalized single-lanthanide metal-organic framework as a ratiometric fluorescent sensor for quantitative visual detection of fluoride ions. Inorganic Chemistry. 2022;61:13627-13636. DOI: 10.1021/acs.inorgchem.2c02533
  92. 92. Yin S, Tong C. Lanthanide coordination polymer nanoparticles as a ratiometric fluorescence sensor for real-time and visual detection of tetracycline by a smartphone and test paper based on the analyte-triggered antenna effect and inner filter effect. Analytica Chimica Acta. 2022;1206:339809. DOI: 10.1016/j.aca.2022.339809
  93. 93. Song L, Tian F, Liu Z. Lanthanide doped metal-organic frameworks as a ratiometric fluorescence biosensor for visual and ultrasensitive detection of serotonin. Journal of Solid State Chemistry. 2022;312:123231. DOI: 10.1016/j.jssc.2022.123231
  94. 94. Liu Y, Xu X, Lu H, Yan B. Dual-emission ratiometric fluorescent probe-based lanthanide-functionalized hydrogen-bonded organic framework for the visual detection of methylamine. Journal of Materials Chemistry C. 2022;10:1212-1219. DOI: 10.1039/d1tc04613f
  95. 95. Hu JJ, Li YG, Wen HR, Liu SJ, Peng Y, Liu CM. Stable lanthanide metal-organic frameworks with ratiometric fluorescence sensing for amino acids and tunable proton conduction and magnetic properties. Inorganic Chemistry. 2022;61:6819-6828. DOI: 10.1021/acs.inorgchem.2c00121
  96. 96. Li S, Yu L, Xiong L, Xiao Y. Ratiometric fluorescence and chromaticity dual-readout assay for β-glucuronidase activity based on luminescent lanthanide metal-organic framework. Sensors and Actuators B: Chemical. 2022;355:131282. DOI: 10.1016/j.snb.2021.131282

Written By

Claudio Pettinari, Andrei Drozdov and Yuriy Belousov

Submitted: 14 November 2022 Reviewed: 28 November 2022 Published: 03 January 2023