Next Article in Journal
The Non-Anhydrous, Minimally Basic Synthesis of the Dopamine D2 Agonist [18F]MCL-524
Previous Article in Journal
Electronic Currents and Anapolar Response Induced in Molecules by Monochromatic Light
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Yb to Tb Cooperative Upconversion in Supramolecularly Assembled Complexes in a Solution

1
Equipe de Synthèse Pour l’Analyse (SynPA), Institut Pluridisciplinaire Hubert Curien (IPHC), UMR 7178, CNRS/Université de Strasbourg, 25 rue Becquerel, 67087 Strasbourg, France
2
Laboratoire de Bioimagerie et Pathologies, CNRS UMR 7021, Faculté de pharmacie,Université de Strasbourg, 74 route du Rhin, 67401 Illkirch, France
*
Author to whom correspondence should be addressed.
Submission received: 30 July 2021 / Revised: 26 August 2021 / Accepted: 1 September 2021 / Published: 6 September 2021
(This article belongs to the Special Issue Molecular and Supramolecular Upconversion)

Abstract

:
The podand-type ligand L, based on a tertiary amine substituted by three pyridyl-6-phosphonic acid functions, forms hydrated complexes with Ln3+ cations. The luminescence properties of the YbL complex were studied in D2O as a function of the pD and temperature. In basic conditions, increases in the luminescence quantum yield and the excited state lifetime of the Yb centered emission associated with the 2F5/22F7/2 transition were observed and attributed to a change in the hydration number from two water molecules in the first coordination sphere of Yb at acidic pH to a single one in basic conditions. Upon the addition of TbCl3 salts to a solution containing the YbL complex in D2O, heteropolynuclear Yb/Tb species formed, and excitation of the Yb at 980 nm resulted in the observation of the typical visible emission of Tb as a result of a cooperative upconversion (UC) photosensitization process. The UC was further evidenced by the quadratic dependence of the UC emission as a function of the laser power density.

1. Introduction

The upconversion (UC) phenomenon is an anti-Stokes process in which low-energy photons are converted into higher energy ones. This generally results in the observation of visible-to-NIR emission upon excitation in the NIR domain [1,2]. This process is particularly useful for applications in biology, as the excitation and emission wavelengths match the biological window where the background is low. It thus provides a low autofluorescence background and a high penetration depth [3,4,5].
Although it is mostly observed in the solid state [6,7,8], examples at the molecular scale in solutions [9,10,11,12] are flourishing thanks to the wide use of lanthanides, with their unique luminescence properties such as long luminescent lifetimes and line-like emission spectra [13,14]. The molecular strategy opens the door to new perspectives in biological applications, considering that it might resolve the drawbacks due to the biotoxicity and hazardous reproducibility of nanomaterials [15].
However, in contrast to solid state devices and nanoparticles [7], the design of the molecular devices in solution is still challenging as a result of important non-radiative deactivations due to energy transfer to OH, NH and CH oscillators [16,17]. These oscillators favor the decay of the intermediate excited state and inhibit the absorption of a second photon or of an energy quantum that is mandatory to reach and emit at a higher energy. One can limit these deactivations by (1) using deuterated oscillators as in D2O which have less impact on quenching of the intermediate state [18]; (2) protecting the cation with a polydentate ligand, thereby limiting the presence of the solvent in the first coordination sphere [19]; and (3) introducing bulky substituents which will keep the solvent molecules away from the NIR emitters [20]. Thanks to these strategies, the number of reported molecular upconverting devices has increased in the last decade.
Among the different mechanisms used to develop upconverting devices [1,21], ETU is reported to be the most efficient for the observation of UC at the molecular scale [22]. Its efficiency results from the possibility of having at least one sensitizer transferring its energy to an emitting acceptor center. Thanks to its unique 2F5/22F7/2 electronic transition at 980 nm, Yb(III) is the most common sensitizer used, and it allows for reaching the excited state of numerous other lanthanides such as Eu, Tb, Ho, Er or Tm, all of which can emit in the visible spectrum [11,23,24,25]. Between those different pairs, the Yb/Tb one introduces a particular mechanism called cooperative sensitization [11,12,26,27]. This mechanism helps to overstep the absence of resonant electronic transition between Yb and Tb, but the efficiency is expected to be weak. In that case, the UC process is possible by exciting 2 Yb centers, which then transfer their energy to the excited state of the Tb (5D4), allowing its emission in the visible spectrum (Scheme 1).
Considering our previous findings on the cooperative sensitization of Tb by Yb complexes [11,12], we anticipated that a Yb complex formed by complexation with the heptadentate ligand L (Scheme 1) [28] might offer opportunities to observe UC in D2O solutions. The spectroscopic properties of the YbL complex were studied in detail together with the interactions of the complex with Tb cations and were shown to allow the formation of heteropolynuclear Yb/Tb complexes that exhibit UC in D2O.

2. Materials and Methods

The (YbL) complex was prepared by following a previously described procedure [28]. Spectroscopic measurements were performed with 10 × 10 mm² quartz suprasil certified cells (Helma Analytics). The steady state emission spectra were recorded on an Edinburgh Instrument FLP920 spectrometer working with a continuous 450-W Xe Lamp and a red-sensitive R928 photomultiplier from Hamamatsu in a Peltier housing for visible detection (230–900 nm) or a Hamamatsu R5 509-72 photomultiplier for the Vis-NIR part. 330-nm and 850-nm high-pass filters were used to eliminate the second-order artifacts when recording in the visible and NIR spectra, respectively. The steady state upconversion emission spectra were recorded on the FLP920 spectrometer using an MDL-3 III-980-nm 2-W laser (CNI Laser) with a 850-nm high-pass filter between the source and the sample to remove any potential visible excitation. The spectra were recorded using the Peltier-cooled R928 photomultiplier from Hamamatsu, using a fully opened emission slit width (20 nm) by accumulation of 10 spectra at 1 point/nm, with a dwell time of 0.1 s per nm. The pD value was adjusted with a solution of NaOD and a solution of DCl and was calculated with use of the equation of Mikkelsen (pD = pH + 0.4) [29].
The phosphorescence lifetimes were measured with the same instrument working in the multi-channel spectroscopy mode and using a Xenon flash lamp as the excitation source. Errors for the luminescence lifetimes were estimated to be ±10%.
The upconversion quantum yields were measured by following the published procedures [6,11] with a diluted solution (optical density < 0.05) using [TbGlu(H2O)]Na as a reference [30].
Low-temperature measurements were performed with a Mercury iTC thermocryostat equipped with a HiCUBE turbopump. The solution was prepared with 10% glycerol in a Tris buffer (10 mM, pH 7).
Power-dependent steady state and time-resolved UC luminescence measurements were performed in quartz cuvettes (Hellma QS Semi-Micro Cells). The continuous-wave laser beam (λ = 975 nm) of a single-mode fiber-coupled laser diode was collimated (Thorlabs, TC12APC-980, Ottawa, Canada) and focused in the cuvette by a +100-mm focal length lens. From the measured beam waist radius (w = 30.6 μm), we estimated the average excitation intensity in the beam waist to be 8.1 kW/cm2 at the maximum laser power. The emission of the sample was collected by a 10 × 0.25 numerical aperture objective and focused on an avalanche photodiode (Excelitas, SPCM-AQRH-16, Wiesbaden, germany). The scattered laser light was removed by a low-pass filter (Chroma, ET750-SP-2P, Olching, Germany) and a band-pass filter (Semrock, FF01-550/88-25, Vanves, France). For the steady state measurements (emitted intensity as a function of the excitation intensity), the output power of the laser was adjusted electronically (by varying the current intensity sent to the diode) and optically by using neutral density filters (Thorlabs) and then systematically measured by a power meter (Newport, 1917R and Ophir, 30A-BB-18). The signal of the photodiode was collected by a computer equipped with a multifunction input/output card (National Instruments multifunction board, PCIe 6711). During the time-resolved measurements, the laser output was modulated electronically at 32 Hz with a 12% duty cycle square waveform by a multifunction input/output card (National Instruments multifunction board, PCIe 6361). The single-photon events detected by the photodiode and the synchronization signal of the laser were recorded by a time-correlated single-photon counting board (Becker & Hickl, SPC-830, Berlin, Germany). SPCM 9.75 software (Becker & Hickl) was used to both record the data and build the photon distribution over time. Experimental luminescence decays were fitted by using the coupled differential rate equations model discussed in the main text in Matlab 2017b.

3. Results and Discussions

3.1. Spectroscopic Properties of the YbL Complex

The YbL complex was prepared as previously described [28]. Considering the peculiar hydration change observed for the parent Gd complex [31], the spectroscopic properties of the Yb analog were first studied as a function of the pD in the D2O solution. Figure 1 shows the evolution of the emission spectra and of the excited state lifetime of the YbL complex at various pD values. When moving from acidic conditions to basic conditions, an increase in the Yb intensity and lifetime was observed. This phenomenon can be explained by a change in the hydration state of the complex as a function of the acidity of the medium. At pD values below six, the complex contained two water molecules in the first coordination sphere, which was completed by the coordination of the heptadentate ligand, forming the [(YbL)(H2O)2] species. Above a pD of six, deprotonation of one of the phosphonate functions resulted in the hydrogen bridging of two phosphonate functions with a concomitant overall contraction of the ligand around the Ln atom, resulting in the repulsion of one of the two water molecules and the formation of the [(YbL)(H2O)] species. Due to the strong quenching effect of water [16,17], the displacement of one water molecule led to a large increase in the luminescence intensity and lifetime. In addition to the main transition observed at 975 nm at a low pD, a hypsochromic shift of the main band toward 970 nm was observed, as well as new bands appearing at ca. 1000 nm and 1040 nm when increasing the pD, together with a shoulder on the high energy tail at ca. 960 nm. While the former bands were ascribed to the 2F5/22F7/2 transitions of Yb with MJ splitting of the ground and excited states [32], the latter was ascribed to hot bands of Yb, as evidenced by the variable temperature measurements (see below) [33].

3.2. Determination of the Yb-Centered Quantum Yield

In order to obtain more insights into the photophysical properties of the Yb complex, the intrinsic luminescence quantum yield of YbL in D2O was determined using Equation (1). τobs corresponds to the observed excited state lifetime measured upon excitation of the ligand, and it amounts to 16.7 µs. τrad is the radiative lifetime of Yb in the complex, which could be determined using the modified Einstein Equation (2) [34,35,36]. In this second equation, NA is Avogadro’s number, c is the speed of light in a vacuum (in cm s−1), ῦ is the barycenter of the transition in cm−1, n is the refractive index of the medium ( n D 2 O = 1.33), ε(ῦ) is the absorption spectrum of the transition (in M−1 cm−1) and gl and gu are related to the degeneracies of the ground and excited states, respectively, and are equal to 2J + 1, where J = 7/2 for gl and J = 5/2 for gu. A value of 1.13 ms was calculated for τrad, which is generally between 0.5 and 1.3 ms in the solution [21,37]. Thus, an intrinsic Yb quantum yield of 1.5% could be calculated at a pD of 7.5:
Φ Y b Y b = k r a d k o b s = τ o b s τ r a d
1 τ r a d = 2303   8 π c n 2 υ ~ ² N A g l g u ε υ ~ d υ ~ With   υ ~ = υ ~ × ε υ ~ d υ ~ ε υ ~ d υ ~
The efficiency of sensitization ηsens was also determined using Equation (3) [36]. ϕ L Y b is the luminescent quantum yield upon ligand excitation, and it was equal to 1.2% [28]. Compared with other complexes [38,39], the value of 0.8 obtained for ηsens in YbL was higher and pointed to a remarkable transfer from the ligand triplet state to the Yb center:
η s e n s = ϕ L Y b ϕ Y b Y b

3.3. Low Temperature Measurements

To further characterize this complex, the Yb emission was monitored as a function of the temperature by excitation of the ligand to observe the evolution of the Yb signal in the 77–300 K range in D2O containing 12% glycerol. As shown in Figure 2, three sharp bands were observed, which could be ascribed to the 2F5/22F7/2 transitions of Yb with MJ splitting of the ground and excited states [32]. The lifetime was also measured (Table 1), showing a gradual decrease upon heating. The higher lifetime values at a low temperature can be explained by the minimization of the non-radiative transitions due to the solvent.

3.4. Upconversion Experiments

In order to check the possibility of a Yb-to-Tb UC process, titration was performed in D2O in which the Yb complex was titrated by additions of aliquots of a solution of TbCl3 salts. The emission spectrum was recorded in the visible domain upon excitation at 980 nm in the 2F5/22F7/2 absorption band of Yb (Figure 3). After the first additions, characteristic Tb-centered emission bands emerged at 485, 545, 580 and 620 nm, corresponding to the 5D47FJ (with J = 6, 5, 4 and 3, respectively) transitions of Tb. The signal then increased up to 1.5 eq and then formed a plateau right after. To confirm the UC process, a Log I vs. Log P plot was performed by monitoring the 5D47F5 transition as a function of the power of the laser at 980 nm (Figure 4) for a mixture containing a 1.4 equivalent of Tb added to the Yb complex. A value close to 2 (1.92 found) indicated the involvement of two photons in the process and confirmed the UC phenomenon. The UC quantum yield was determined using a protocol described in the literature [6,11]. A value of 2.6 × 10−8 was obtained (P = 10.8 W cm−2), which was of the same order of magnitude as those reported for similar compounds [11].
The data of the titration were analyzed by a nonlinear least-squares method using Specfit software [40,41]. The analysis pointed to the formation of three emitting species, which were ascribed to [(YbL)2Tbx] complexes with x = 1, 2 or 3. Their calculated distributions are presented in the inset of Figure 3. These compositions were consistent with previous spectroscopy titrations of the ligand by the addition of Ln(III) (Ln = Eu or Tb) [28]. All three species were able to generate UC, but the relative efficiencies, which were obtained thanks to the fitting procedure, pointed to the UC efficiencies of [(YbL)2Tb2] and [(YbL)2Tb3] being 1.84 and 5.6 times higher than that of [(YbL)2Tb], respectively.
To get deeper insight into the UC mechanism, time-resolved measurements were performed with a modulated 975-nm laser diode generating excitation pulses with a square temporal profile. As shown in Figure 5, 3–4 milliseconds were required for the rise of the UC emission before the steady state was reached. This result indicated a slow kinetic step in the UC process which was related to the energy transfer occurring throughout the cooperative energy transfer UC process [27] (Scheme 2). The fit of the rising and decaying signals was made by using the model described by Nonat et al. [11] by updating the fitting data with the initial parameters (τYb and εYb). In this model, the presence of two species was assumed as a result of the presence of the different complexes in the solution. In our case, the addition of the 1.4 equivalent of Tb per Yb complex pointed to the presence of [(YbL)2Tb2] and [(YbL)2Tb3] as the dominant species in the solution. The upconversion lifetimes (i.e., the Tb luminescence lifetimes upon excitation at 980 nm) were determined to be 0.38 ± 0.04 ms and 1.86 ± 0.2 ms. The shorter Tb lifetime may be attributed to a Tb cation more exposed to the solvent molecules (with a higher hydration number). The associated UC rate constants kUC were determined to be of 755 and 355 s−1. In comparison with the literature, the UC rate constants obtained were somehow smaller but consistent with those reported in the literature [11,27,42].
The energy transfer efficiency of the UC process ηUC was then calculated using Equation (4) [11]:
η U C = k U C k U C + k r + k n r
where kUC represents the UC energy transfer rate. The radiative rate constant kr of the Yb complex is the inverse of τr (1.13 ms) and is equal to 885 s−1. The non-radiative rate constant knr is obtained from the intrinsic quantum yield Φ Y b Y b by
Φ Y b Y b = k r   k r + k n r
Using Φ Y b Y b = 1.5%, a value of 58,100 s−1 was obtained for knr. The energy transfer efficiency of the UC process ηUC was then calculated to be equal to 1.8%.

4. Conclusions

Our in-depth study of the YbL complex in D2O showed a dependence on the hydration state as a function of the pD, resulting in the presence of two deuterated water molecules in the first coordination sphere of the complex at low pD values and a single one around the physiological pD. The dehydration was accompanied by a large increase in both the luminescence intensity and the excited state lifetime of the Yb emission. An increase in Yb emission was also observed at low temperatures as a result of decreased non-radiative transitions due to water molecules.
When the Yb complex was titrated by Tb3+ salts in a D2O solution, heteropolynuclear [(YbL)2Tbx] (x = 1–3) complexes formed, and direct excitation of the Yb atom at 980 nm in the heteronuclear complexes resulted in emission of the Tb atoms in the visible domain as a result of a cooperative sensitization UC mechanism. The energy transfer efficiency of the UC process was determined to be 1.8%, while the overall UC quantum yield value of 2.6 × 10−8 in D2O was obtained for an excitation power density of 10.8 W cm−2. Although weak, this value was in line with previous results in the literature [11], despite the presence of water molecules in the first coordination sphere of Yb. The improvement of the protection of Yb toward water de-excitation paves the way to even better UC devices for luminescence bioanalytical applications.

Author Contributions

Conceptualization, L.K.S. and C.C.; methodology, L.K.S., C.C. and L.J.C.; software, F.P.; validation, A.M.N. and L.J.C.; formal analysis, L.K.S. and F.P.; writing—original draft preparation, L.K.S. and L.J.C.; writing—review and editing, L.K.S., Y.M. and L.J.C.; supervision, L.J.C. All authors have read and agreed to the published version of the manuscript.

Funding

L.K.S. gratefully acknowledges the financial support from the French Ministère de l’enseignement supérieur et de la recherche and the France Canada Research Fund. C.C. and L.J.C. acknowledge the financial support of the French National Research Agency (project ANR LUCAS n° ANR-19-CE29-0014-01).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

On request to the authors.

Conflicts of Interest

The authors declare no competing financial interest.

References

  1. Auzel, F. Upconversion and Anti-Stokes Processes with f and d Ions in Solids. Chem. Rev. 2004, 104, 139–173. [Google Scholar] [CrossRef]
  2. Gamelin, D.R.; Güdel, H.U. Design of luminescent inorganic materials: New photophysical processes studied by optical spectroscopy. Acc. Chem. Res. 2000, 33, 235–242. [Google Scholar] [CrossRef]
  3. Idris, N.M.; Jayakumar, M.K.G.; Bansal, A.; Zhang, Y. Upconversion nanoparticles as versatile light nanotransducers for photoactivation applications. Chem. Soc. Rev. 2015, 44, 1449–1478. [Google Scholar] [CrossRef]
  4. Hemmer, E.; Benayas, A.; Légaré, F.; Vetrone, F. Exploiting the biological windows: Current perspectives on fluorescent bioprobes emitting above 1000 nm. Nanoscale Horiz. 2016, 1, 168–184. [Google Scholar] [CrossRef]
  5. Zhu, X.; Su, Q.; Feng, W.; Li, F. Anti-Stokes shift luminescent materials for bio-applications. Chem. Soc. Rev. 2017, 46, 1025–1039. [Google Scholar] [CrossRef] [PubMed]
  6. Chen, G.; Qiu, H.; Prasad, P.N.; Chen, X. Upconversion nanoparticles: Design, nanochemistry, and applications in Theranostics. Chem. Rev. 2014, 114, 5161–5214. [Google Scholar] [CrossRef]
  7. Haase, M.; Schäfer, H. Upconverting Nanoparticles. Angew. Chem. Int. Ed. 2011, 50, 5808–5829. [Google Scholar] [CrossRef] [PubMed]
  8. Suffren, Y.; Golesorkhi, B.; Zare, D.; Guénée, L.; Nozary, H.; Eliseeva, S.V.; Petoud, S.; Hauser, A.; Piguet, C. Taming Lanthanide-Centered Upconversion at the Molecular Level. Inorg. Chem. 2016, 55, 9964–9972. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  9. Nonat, A.; Chan, C.F.; Liu, T.; Platas-Iglesias, C.; Liu, Z.; Wong, W.-T.; Wong, W.-K.; Wong, K.-L.; Charbonnière, L.J. Room temperature molecular up conversion in solution. Nat. Commun. 2016, 7, 11978. [Google Scholar] [CrossRef] [Green Version]
  10. Souri, N.; Tian, P.; Platas-Iglesias, C.; Wong, K.-L.L.; Nonat, A.; Charbonnière, L.J. Upconverted Photosensitization of Tb Visible Emission by NIR Yb Excitation in Discrete Supramolecular Heteropolynuclear Complexes. J. Am. Chem. Soc. 2017, 139, 1456–1459. [Google Scholar] [CrossRef] [PubMed]
  11. Nonat, A.; Bahamyirou, S.; Lecointre, A.; Przybilla, F.; Mély, Y.; Platas-Iglesias, C.; Camerel, F.; Jeannin, O.; Charbonnière, L.J. Molecular Upconversion in Water in Heteropolynuclear Supramolecular Tb/Yb Assemblies. J. Am. Chem. Soc. 2019, 141, 1568–1576. [Google Scholar] [CrossRef]
  12. Knighton, R.C.; Soro, L.K.; Lecointre, A.; Pilet, G.; Fateeva, A.; Pontille, L.; Francés-Soriano, L.; Hildebrandt, N.; Charbonnière, L.J. Upconversion in molecular hetero-nonanuclear lanthanide complexes in solution. Chem. Commun. 2021, 57, 53–56. [Google Scholar] [CrossRef]
  13. Bünzli, J.C.G.; Piguet, C. Taking advantage of luminescent lanthanide ions. Chem. Soc. Rev. 2005, 34, 1048–1077. [Google Scholar] [CrossRef] [PubMed]
  14. Bünzli, J.C.G. On the design of highly luminescent lanthanide complexes. Coord. Chem. Rev. 2015, 293–294, 19–47. [Google Scholar] [CrossRef]
  15. Gnach, A.; Lipinski, T.; Bednarkiewicz, A.; Rybka, J.; Capobianco, J.A. Upconverting nanoparticles: Assessing the toxicity. Chem. Soc. Rev. 2015, 44, 1561–1584. [Google Scholar] [CrossRef] [PubMed]
  16. Horrocks, W.D., Jr.; 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]
  17. Beeby, A.; Clarkson, I.M.; Dickins, R.S.; Faulkner, S.; Parker, D.; Royle, L.; de Sousa, A.S.; Williams, J.A.G.; Woods, M. 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. J. Chem. Soc. Perkin Trans. 1999, 2, 493–504. [Google Scholar] [CrossRef]
  18. Bischof, C.; Wahsner, J.; Scholten, J.; Trosien, S.; Seitz, M. Quantification of C-H quenching in near-IR luminescent ytterbium and neodymium cryptates. J. Am. Chem. Soc. 2010, 132, 14334–14335. [Google Scholar] [CrossRef] [PubMed]
  19. Charbonnière, L.; Mameri, S.; Kadjane, P.; Platas-Iglesias, C.; Ziessel, R. Tuning the coordination sphere around highly luminescent lanthanide complexes. Inorg. Chem. 2008, 47, 3748–3762. [Google Scholar] [CrossRef]
  20. Zhang, T.; Zhu, X.; Cheng, C.C.W.; Kwok, W.M.; Tam, H.L.; Hao, J.; Kwong, D.W.J.; Wong, W.K.; Wong, K.L. Water-soluble mitochondria-specific ytterbium complex with impressive NIR emission. J. Am. Chem. Soc. 2011, 133, 20120–20122. [Google Scholar] [CrossRef]
  21. Nonat, A.M.; Charbonnière, L.J. Upconversion of light with molecular and supramolecular lanthanide complexes. Coord. Chem. Rev. 2020, 409, 213192. [Google Scholar] [CrossRef]
  22. Zare, D.; Suffren, Y.; Guénée, L.; Eliseeva, S.V.; Nozary, H.; Aboshyan-Sorgho, L.; Petoud, S.; Hauser, A.; Piguet, C. Smaller than a nanoparticle with the design of discrete polynuclear molecular complexes displaying near-infrared to visible upconversion. Dalton Trans. 2015, 44, 2529–2540. [Google Scholar] [CrossRef] [PubMed]
  23. Hemmer, E.; Venkatachalam, N.; Hyodo, H.; Hattori, A.; Ebina, Y.; Kishimoto, H.; Soga, K. Upconverting and NIR emitting rare earth based nanostructures for NIR-bioimaging. Nanoscale 2013, 5, 11339–11361. [Google Scholar] [CrossRef]
  24. Meijer, M.S.; Rojas-Gutierrez, P.A.; Busko, D.; Howard, I.A.; Frenzel, F.; Würth, C.; Resch-Genger, U.; Richards, B.S.; Turshatov, A.; Capobianco, J.A.; et al. Absolute upconversion quantum yields of blue-emitting LiYF4:Yb3+,Tm3+ upconverting nanoparticles. Phys. Chem. Chem. Phys. 2018, 20, 22556–22562. [Google Scholar] [CrossRef] [Green Version]
  25. Zhou, B.; Yan, L.; Tao, L.; Song, N.; Wu, M.; Wang, T.; Zhang, Q. Enabling Photon Upconversion and Precise Control of Donor–Acceptor Interaction through Interfacial Energy Transfer. Adv. Sci. 2018, 5, 1700667. [Google Scholar] [CrossRef] [Green Version]
  26. Salley, G.M.; Valiente, R.; Guedel, H.U. Luminescence upconversion mechanisms in Yb3+–Tb3+ systems. J. Lumin. 2001, 94–95, 305–309. [Google Scholar] [CrossRef]
  27. Salley, G.M.; Valiente, R.; Güdel, H.U. Phonon-assisted cooperative sensitization of Tb3+ in SrCl2:Yb, Tb. J. Phys. Condens. Matter 2002, 14, 301. [Google Scholar] [CrossRef] [Green Version]
  28. Charpentier, C.; Salaam, J.; Lecointre, A.; Jeannin, O.; Nonat, A.; Charbonnière, L.J. Phosphonated Podand Type Ligand for the Complexation of Lanthanide CationsPhosphonated Podand Type Ligand for the Complexation of Lanthanide Cations. Eur. J. Inorg. Chem. 2019, 2019, 2168–2174. [Google Scholar] [CrossRef]
  29. Mikkelsen, K.; Siguri, A.; Nielsen, O. Acidity measurements jyith the glass electrode is h,o-d,o mixtures. J. Phys. Chem. 1960, 64, 632–637. [Google Scholar] [CrossRef]
  30. Weibel, N.; Charbonnière, L.J.; Guardigli, M.; Roda, A.; Ziessel, R. Engineering of Highly Luminescent Lanthanide Tags Suitable for Protein Labeling and Time-Resolved Luminescence Imaging. J. Am. Chem. Soc. 2004, 126, 4888–4896. [Google Scholar] [CrossRef]
  31. Charpentier, C.; Salaam, J.; Nonat, A.; Carniato, F.; Jeannin, O.; Brandariz, I.; Esteban-Gomez, D.; Platas-Iglesias, C.; Charbonnière, L.J.; Botta, M. pH-Dependent Hydration Change in a Gd-Based MRI Contrast Agent with a Phosphonated Ligand. Chem.-A Eur. J. 2020, 26, 5407–5418. [Google Scholar] [CrossRef]
  32. Ziessel, R.F.; Ulrich, G.; Charbonnière, L.; Imbert, D.; Scopelliti, R.; Bünzli, J.C.G. NIR lanthanide luminescence by energy transfer from appended terpyridine-boradiazaindacene dyes. Chem.-A Eur. J. 2006, 12, 5060–5067. [Google Scholar] [CrossRef]
  33. Asano-Someda, M.; Kaizu, Y. Hot bands of (f, f*) emission from ytterbium(III) porphyrins in solution. J. Photochem. Photobiol. A Chem. 2001, 139, 161–165. [Google Scholar] [CrossRef]
  34. Shavaleev, N.M.; Scopelliti, R.; Gumy, F.; Bünzli, J.C.G. Surprisingly bright near-infrared luminescence and short radiative lifetimes of ytterbium in hetero-binuclear Yb-Na chelates. Inorg. Chem. 2009, 48, 7937–7946. [Google Scholar] [CrossRef]
  35. Werts, M.H.V.; Jukes, R.T.F.; Verhoeven, J.W. The emission spectrum and the radiative lifetime of Eu3+ in luminescent lanthanide complexes. Phys. Chem. Chem. Phys. 2002, 4, 1542–1548. [Google Scholar] [CrossRef]
  36. Eliseeva, S.V.; Bünzli, J.C.G. Lanthanide luminescence for functional materials and bio-sciences. Chem. Soc. Rev. 2010, 39, 189–227. [Google Scholar] [CrossRef]
  37. Bünzli, J.C.G. Lanthanide luminescence for biomedical analyses and imaging. Chem. Rev. 2010, 110, 2729–2755. [Google Scholar] [CrossRef]
  38. Nocton, G.; Nonat, A.; Gateau, C.; Mazzanti, M. Water stability and luminescence of lanthanide complexes of tripodal ligands derived from 1,4,7-triazacyclononane: Pyridinecarboxamide versus pyridinecarboxylate donors. Helv. Chim. Acta 2009, 92, 2257–2273. [Google Scholar] [CrossRef]
  39. Knighton, R.C.; Soro, L.K.; Troadec, T.; Mazan, V.; Nonat, A.M.; Elhabiri, M.; Saffon-Merceron, N.; Djenad, S.; Tripier, R.; Charbonnière, L.J. Formation of Heteropolynuclear Lanthanide Complexes Using Macrocyclic Phosphonated Cyclam-Based Ligands. Inorg. Chem. 2020, 59, 10311–10327. [Google Scholar] [CrossRef] [PubMed]
  40. Gampp, H.; Maeder, M.; Meyer, C.J.; Zuberbühler, A.D. Calculation of equilibrium constants from multiwavelength spectroscopic data-III. Model-free analysis of spectrophotometric and ESR titrations. Talanta 1985, 32, 1133–1139. [Google Scholar] [CrossRef] [PubMed]
  41. Maeder, M.; Zuberbühler, A.D. Nonlinear Least-Squares Fitting of Multivariate Absorption Data. Anal. Chem. 1990, 62, 2220–2224. [Google Scholar] [CrossRef]
  42. Suffren, Y.; Zare, D.; Eliseeva, S.V.; Guénée, L.; Nozary, H.; Lathion, T.; Aboshyan-Sorgho, L.; Petoud, S.; Hauser, A.; Piguet, C. Near-Infrared to Visible Light-Upconversion in Molecules: From Dream to Reality. J. Phys. Chem. C 2013, 117, 26957–26963. [Google Scholar] [CrossRef]
Scheme 1. (left) Principle of the cooperative sensitization between Yb and Tb [27]. (right) Representation of ligand L and its dihydrated Yb complex (H atoms are omitted for clarity).
Scheme 1. (left) Principle of the cooperative sensitization between Yb and Tb [27]. (right) Representation of ligand L and its dihydrated Yb complex (H atoms are omitted for clarity).
Chemistry 03 00074 sch001
Figure 1. (left) Emission spectra of the YbL complex with an increase in the pD value in the D2O. The arrow indicates the increase of the pD. (right) Evolution of the emission integral (red) and the lifetime (blue) (YbL = 9.5 × 10−5 M, D2O), where λexc = 270 nm.
Figure 1. (left) Emission spectra of the YbL complex with an increase in the pD value in the D2O. The arrow indicates the increase of the pD. (right) Evolution of the emission integral (red) and the lifetime (blue) (YbL = 9.5 × 10−5 M, D2O), where λexc = 270 nm.
Chemistry 03 00074 g001
Figure 2. Dependence of the Yb emission intensity as a function of the temperature for a 5.3 × 10−5 M solution of YbL in D2O containing 12% glycerol (λexc = 270 nm).
Figure 2. Dependence of the Yb emission intensity as a function of the temperature for a 5.3 × 10−5 M solution of YbL in D2O containing 12% glycerol (λexc = 270 nm).
Chemistry 03 00074 g002
Figure 3. Evolution of the visible terbium emission during the titration of YbL by TbCl3·6H2O (λex = 980 nm; YbL = 9.7 × 10−5 M; Tb = 7.7 × 10−4 M; solutions in D2O, pD = 7.7). The inset shows the evolution of the UC intensity (black dots) and evolution of the concentrations of the species formed during the titration ([(YbL)2Tb], blue; [(YbL)2Tb2], red and [(YbL)2Tb3] in green).
Figure 3. Evolution of the visible terbium emission during the titration of YbL by TbCl3·6H2O (λex = 980 nm; YbL = 9.7 × 10−5 M; Tb = 7.7 × 10−4 M; solutions in D2O, pD = 7.7). The inset shows the evolution of the UC intensity (black dots) and evolution of the concentrations of the species formed during the titration ([(YbL)2Tb], blue; [(YbL)2Tb2], red and [(YbL)2Tb3] in green).
Chemistry 03 00074 g003
Figure 4. Log–log plot of the Tb-emitted intensity of the 5D47F5 transition at 545 nm as a function of the excitation intensity of the 980 nm laser for a 5.3 × 10−5 M solution of YbL in D2O containing a 1.4 equivalent of TbCl3·6H2O.
Figure 4. Log–log plot of the Tb-emitted intensity of the 5D47F5 transition at 545 nm as a function of the excitation intensity of the 980 nm laser for a 5.3 × 10−5 M solution of YbL in D2O containing a 1.4 equivalent of TbCl3·6H2O.
Chemistry 03 00074 g004
Figure 5. Time-resolved rise and decay curves of the Tb-centered UC emission intensity at 550 nm (in blue) upon time-gated excitation at 974 nm (8.1 kW cm−2, orange square) for a 9.5 × 10−5 M solution of YbL in D2O containing a 1.4 equivalent of TbCl3·6H2O. The red line corresponds to the fit of the data (see text).
Figure 5. Time-resolved rise and decay curves of the Tb-centered UC emission intensity at 550 nm (in blue) upon time-gated excitation at 974 nm (8.1 kW cm−2, orange square) for a 9.5 × 10−5 M solution of YbL in D2O containing a 1.4 equivalent of TbCl3·6H2O. The red line corresponds to the fit of the data (see text).
Chemistry 03 00074 g005
Scheme 2. Energy level diagram of the cooperative UC sensitization process in the [(YbL)2Tbx] complex (x = 1–3). Symbols * and ** denote singly and doubly excited states respectively.
Scheme 2. Energy level diagram of the cooperative UC sensitization process in the [(YbL)2Tbx] complex (x = 1–3). Symbols * and ** denote singly and doubly excited states respectively.
Chemistry 03 00074 sch002
Table 1. Ytterbium lifetime in the range of 77–300 K (YbL = 5.3 × 10−5, D2O with 12% glycerol).
Table 1. Ytterbium lifetime in the range of 77–300 K (YbL = 5.3 × 10−5, D2O with 12% glycerol).
T (°K) 77 100 125 150 175 200 250 273 300
τYb (µs) 14.414.114.714.714.513.311.011.19.9
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Soro, L.K.; Charpentier, C.; Przybilla, F.; Mély, Y.; Nonat, A.M.; Charbonnière, L.J. Yb to Tb Cooperative Upconversion in Supramolecularly Assembled Complexes in a Solution. Chemistry 2021, 3, 1037-1046. https://0-doi-org.brum.beds.ac.uk/10.3390/chemistry3030074

AMA Style

Soro LK, Charpentier C, Przybilla F, Mély Y, Nonat AM, Charbonnière LJ. Yb to Tb Cooperative Upconversion in Supramolecularly Assembled Complexes in a Solution. Chemistry. 2021; 3(3):1037-1046. https://0-doi-org.brum.beds.ac.uk/10.3390/chemistry3030074

Chicago/Turabian Style

Soro, Lohona K., Cyrille Charpentier, Frédéric Przybilla, Yves Mély, Aline M. Nonat, and Loïc J. Charbonnière. 2021. "Yb to Tb Cooperative Upconversion in Supramolecularly Assembled Complexes in a Solution" Chemistry 3, no. 3: 1037-1046. https://0-doi-org.brum.beds.ac.uk/10.3390/chemistry3030074

Article Metrics

Back to TopTop