Next Article in Journal
Effect of Clove and Thyme Essential Oils on Candida Biofilm Formation and the Oil Distribution in Yeast Cells
Next Article in Special Issue
Special Issue “Raman Spectroscopy: A Spectroscopic ‘Swiss-Army Knife’”
Previous Article in Journal
Development of Indirect Competitive ELISA for Lithospermic Acid B of Salvia miltiorrhiza with Its Specific Antibodies Generated via Artificial Oil Bodies
Previous Article in Special Issue
Analysis of Cellulose and Lignocellulose Materials by Raman Spectroscopy: A Review of the Current Status
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

On the Hydration of Heavy Rare Earth Ions: Ho3+, Er3+, Tm3+, Yb3+ and Lu3+—A Raman Study

1
TU Dresden, Medizinische Fakultät, Institut für Virologie im MTZ, Fiedlerstr. 42, 01307 Dresden, Germany
2
Technische Universität Bergakademie Freiberg, Institut für Theoretische Physik, Leipziger Str. 23, 09596 Freiberg, Germany
*
Author to whom correspondence should be addressed.
Molecules 2019, 24(10), 1953; https://doi.org/10.3390/molecules24101953
Submission received: 3 April 2019 / Revised: 17 May 2019 / Accepted: 20 May 2019 / Published: 21 May 2019
(This article belongs to the Special Issue Raman Spectroscopy: A Spectroscopic 'Swiss-Army Knife')

Abstract

:
Raman spectra of aqueous Ho3+, Er3+, Tm3+, Yb3+, and Lu3+-perchlorate solutions were measured over a large wavenumber range from 50–4180 cm−1. In the low wavenumber range (terahertz region), strongly polarized Raman bands were detected at 387 cm−1, 389 cm−1, 391 cm−1, 394 cm−1, and 396 cm−1, respectively, which are fairly broad (full widths at half height at ~52 cm−1). These isotropic Raman bands were assigned to the breathing modes, ν1 Ln–O of the heavy rare earth (HRE) octaaqua ions, [Ln(H2O)8]3+. The strong polarization of these bands (depolarization degree ~0) reveals their totally symmetric character. The vibrational isotope effect was measured in Yb(ClO4)3 solutions in H2O and D2O and the shift of the ν1 mode in changing from H2O to D2O further supports the character of the band. The Ln–O bond distances of these HRE ions (Ho3+, Er3+, Tm3+, Yb3+, and Lu3+) follow the order of Ho–O > Er–O > Tm–O > Yb–O > Lu–O which correlates inversely with the band positions of the breathing modes of their corresponding octaaqua ions [Ln(OH2)8]3+. Furthermore, the force constants, kLn–O, were calculated for these symmetric stretching modes. Ytterbium perchlorate solutions were measured over a broad concentration range, from 0.240 mol·L−1 to 2.423 mol·L−1, and it was shown that with increasing solute concentration outer-sphere ion pairs and contact ion pairs were formed. At the dilute solution state (~0.3 mol·L−1), the fully hydrated ions [Yb(H2O)8]3+ exist, while at higher concentrations (CT > 2 mol·L−1), ion pairs are formed. The concentration behavior of Yb(ClO4)3 (aq) shows similar behavior to the one observed for La(ClO4)3(aq), Ce(ClO4)3(aq) and Lu(ClO4)3(aq) solutions. In ytterbium chloride solutions in water and heavy water, representative for the behavior of the other HRE ions, 1:1 chloro-complex formation was detected over the concentration range from 0.422–3.224 mol·L−1. The 1:1 chloro-complex in YbCl3(aq) is very weak, diminishing rapidly with dilution and vanishing at a concentration < 0.4 mol·L−1.

Graphical Abstract

1. Introduction

In aqueous solution, the HRE ions of holmium, erbium, thulium, ytterbium, and lutetium exist in the tervalent state [1] and, with their high charge to radius ratio, are strongly hydrated [2,3]. The HRE ions possess eight water molecules arranged in a square antiprismatic geometry (S8 symmetry) in their first coordination sphere. The hydration geometry of heavy rare earth ions in aqueous solution was determined by X-ray (XRD) and neutron diffraction (ND) [4,5,6] as well as extended X-ray absorption fine structure (EXAFS) [7,8,9,10] techniques. Computer simulations contributed to clarifying the details of the structure and dynamics of the waters in the first hydration sphere of the HRE ions [11,12,13] and their results confirmed the octahedral coordination.
While XRD and neutron diffraction measurements [4,5,6] are carried out in concentrated solutions, EXAFS spectra [7,8,9,10] are measured on dilute or moderately concentrated solutions. At high concentrations, however, it is known that the Ln3+(aq) ions form complex species/ion pairs with common ions (Cl, NO3 and SO42−). Perchlorate, however, acts as a weak complex forming anion and the anion is therefore suitable for hydration studies.
Raman spectroscopy has been applied frequently to characterize hydrated metal ions and related species in aqueous solution and is especially useful in characterizing the species formed at the molecular level such as hydrated ions, ion pairs between metal ions and anions or metal ion hydrolysis. Raman measurements on aqueous Ln3+(aq) ions should allow, in principle, the characterization of the solution structure in greater detail. Recently, light rare earth solutions (La3+, Ce3+, Pr3+, Nd3+ and Sm3+) as well as the solution of the HRE ion, Lu3+, have been measured using Raman spectroscopy [14,15,16,17] and it has been shown that the application of the so-called R-procedure is crucial in exploring the low frequency part of the Raman scattering.
The present study was undertaken to characterize the hydration and speciation in aqueous Ho3+, Er3+, Tm3+, and Yb3+ perchlorate solutions. Lu3+- solutions in water and heavy water have been measured and described in a recent publication [17] and these results are also included in this work. The Ho3+-, Er3+-, Tm3+- and Yb3+- perchlorate solutions were measured over broader concentration ranges and down to the terahertz frequency region at 22 °C. Furthermore, Yb(ClO4)3 solutions in water and heavy water were measured in order to characterize the vibrational isotope effect changing from [Yb(H2O)8]3+ to [Yb(D2O)8]3+ and the symmetric stretching modes of the two species were observed.
In chloride solutions, however, it was shown that light rare earth ions form chloro-complex species, [14,15,16,17]. In order to clarify whether chloro-complex species may also form in these HRE chloride solutions, YbCl3 solutions were measured at varying concentrations in water and heavy water. Recently, inner-sphere chloro- complexes were detected in aqueous solutions on a variety of rare earth chloride solutions using Raman spectroscopy [14,15,16,17].
In this study, we are specifically interested in the vibrational characterization of the Ln–O stretching modes of the octahydrates (fully hydrated ions, [Ln(OH2)8]3+) in dilute perchlorate solutions. The formation of ion pairs between Yb3+ ions and perchlorate has been studied on a concentration series of Yb(ClO4)3 solutions representative for other HRE ions. The influence of Cl on the fully hydrated Ln3+(aq) was studied exemplarily on YbCl3 solutions in water and heavy water. These results are discussed in connection with recently measured aqueous LuCl3 solutions [17].

2. Experimental Details and Data Analysis

Preparation of Solutions
The rare earth ion concentrations of all solutions were analysed by complexometric titration [18]. The solution densities were determined pycnometrically at 22 °C and the molar ratios of water per salt calculated (Rw -values). The solution pH values were measured with pH meter S220 using a pH electrode InLab Expert Pro-ISM (Mettler –Toledo GmbH, Deutschland, Giessen). For Raman spectroscopic measurements, the solutions were filtered through a fine sintered glass frit (1–1.6 µm pore size). The preparation of the lutetium perchlorate solutions at various concentrations were described earlier [17].
Ytterbium perchlorate solutions were prepared from Yb(ClO4)3∙6H2O (Alfa-Aesar (Thermo Fisher), 99.9%, Kandel, Deutschland) dissolved with ultrapure water (PureLab Plus, Ultra-pure Water Purification Systems). A concentrated Yb(ClO4)3 solution was prepared at 2.423 mol·L−1 (Rw = 16.86). This solution was acidified with a slight amount of HClO4 and a pH value at ~ 2.0 was obtained. From this stock solution, the following dilution series was prepared: 1.217 mol·L−1 (Rw = 39.30), 0.808 mol·L−1 (Rw = 62.13), 0.603 mol·L−1 (Rw = 85.16) and 0.240 mol·L−1 (Rw = 223.6). The solutions were analysed for dissolved chloride with a 5% AgNO3 solution and the absence of a white AgCl precipitate was proof that the stock solution was free of Cl.
Two Yb(ClO4)3 solutions in heavy water were prepared from a 2.526 mol·L−1 Yb(ClO4)3(D2O) stock solution with heavy water (99.9 atom% D; Sigma-Aldrich now Merck KGaA, Darmstadt, Germany) at 1.276 mol·L−1 and 0.779 mol·L−1. The deuteration degree for the solution was determined at ~97% D.
YbCl3 solutions were prepared from YbCl3∙6H2O (Sigma-Aldrich, 99.5%) and ultrapure water (PureLab Plus, Ultra-pure Water Purification Systems, now ELGA Labwater, Celle, Deutschland) by weight at 3.224 mol·L−1 (Rw = 15.64), 1.600 mol·L−1 (Rw = 33.06), 0.802 mol·L−1 (Rw = 67.55) and 0.400 mol·L−1 (Rw = 137).
A 3.300 mol·L−1 YbCl3 stock solution in D2O was used to prepare two dilute YbCl3 solutions with heavy water (99.9 atom% D; Sigma-Aldrich) at 0.844 mol·L−1 and 0.422 mol·L−1. The deuteration degree in the dilute solutions was determined at ~96.5% D.
A Tm(ClO4)3 solution was prepared from Tm2O3 (Sigma-Aldrich, 99.9%) which was dissolved with 6 mol·L−1 HClO4 solution (Fisher Scientific GmbH, Schwerte, Deutschland until a clear solution was obtained. The solute concentration was determined at 1.897 mol·L−1 and two dilute solutions were prepared from the stock solution and triply distilled water by weight at 0.980 mol·L−1 and 0.315 mol·L−1. These solutions contained a slight excess of HClO4.
An Er(ClO4)3 stock solution was a commercial product from Alfa-Aesar (Thermo Fisher) (Kandel, Deutschland) at 50 wt%, Reagent Grade (99.9%) at 2.245 moL−1. Two dilute solutions at 1.123 moL−1 and at 0.321 moL−1 were prepared by weight with ultrapure water. These solutions contained a slight excess of HClO4 (pH value ~1.5).
A Ho(ClO4)3 stock solution was prepared from Ho2O3 (Sigma-Aldrich, 99.9%) which was dissolved with 6 mol·L−1 HClO4 (Riedel-de Haen, 70 wt%) until a clear solution was obtained. The solute concentration was determined at 1.675 mol·L−1. Two dilute solutions at 0.838 mol·L−1 and 0.240 mol·L−1 were prepared by weight from the stock solution with ultrapure water. The solutions contained a slight excess of HClO4 (pH value ~1.75).
Raman spectroscopic measurements have been reported in detail elsewhere, so only a brief summary is given [19,20]. Raman spectra were measured in the macro chamber of the T 64000 Raman spectrometer from Jobin Yvon in a 90° scattering geometry at 22 °C. A quartz cuvette was used (Hellma Analytics, Müllheim, Germany) with a 10 mm path length and a volume at 1000 µL. The spectra were excited with the 487.987 nm or the 514.532 nm line of an Ar+ laser at a power level of 1100 mW at the sample. The Yb3+-perchlorate and -chloride solutions have no visible absorption bands and therefore, both excitation wavelengths may be used. Tm(ClO4)3 solutions were measured with the 514.532 nm Ar+ line. Er(ClO4)3 solutions were excited with the 487.987 nm Ar+ line and only the most dilute solution could be measured reliably because the concentrated ones were strongly absorbing at the absorption gap at ~488 nm of its UV-vis spectrum. Ho(ClO4)3 solutions were excited with the 514.532 nm Ar+ line and only the most dilute solution could be reliably measured.
After passing the spectrometer in subtractive mode, with gratings of 1800 grooves/mm, the scattered light was detected with a cooled CCD detector. The scattering geometries IVV = (X[ZZ]Y) and IVH = (X[ZX]Y) are defined as follows: the propagation (wave vector direction) of the exciting laser beam is in X direction and the propagation of the observed scattered light is in Y direction, the 90° geometry. The polarisation (electrical field vector) of the laser beam is fixed in Z direction (vertical) and the polarisation of the observed scattered light is observed in Z direction (vertical) for the IVV scattering geometry. For IVH the electric field vector of the exciting laser beam is in Z direction (vertical) and the observed scattering light is polarized in the X direction (horizontal). Thus, for the two scattering geometries it follows:
I VV = I ( X [ ZZ ] Y )   =   45 α ¯ 2 +   4 γ 2 ,
I VH =   I ( X [ ZX ] Y ) =   3 γ 2 ,
The symbols α ¯ ' and γ’ are the isotropic and the anisotropic invariant of the Raman polarizability tensor, respectively [19]. The isotropic spectrum (Iiso) was constructed according to Equation (3):
Iiso = IVV − 4/3IVH,
The polarization degree of the Raman bands, ρ (ρ = IVH/IVV) was determined using an analyzer and adjusted, if necessary, before each measuring cycle using CCl4 [19].
The calibration of the Raman spectra has been carried out using plasma lines [19]. The accuracy of the peak positions for the perchlorate deformation modes were not better than ±0.5 cm−1 and for the much narrower ν1(a1)ClO4 band ±0.2 cm−1. The peak positions of the bands were determined by fitting the baseline corrected bands with a Gauss-Lorenz product function (see ref. [20]). The accuracy of the weak and much broader ν1 Ln–OH2/OD2 bands was ±1 cm−1 using the perchlorate band at 461 cm−1 as an internal reference band.
In order to characterize the spectral features in the low wave number region, the Raman spectra in I-format were reduced and the R-spectra obtained. The R( ν ˜ ) spectra are independent of the excitation wavenumber ν ˜ L and the measured Stokes intensity is further corrected for the scattering factor (νL- ν ˜ )3. (The scattering factor must be to the power of 3 when applying counting methods [21].) The spectra were further corrected for the Bose-Einstein temperature factor, B = [1-exp(-h ν ˜ c/kT)] and the frequency factor, ν ˜ , to give the so-called reduced spectrum, R( ν ˜ ) (detailed in earlier publications [19,20]). The isotropic spectrum in R-format, Riso, is calculated according to equation 3 but substituting the spectra in I-format, IVV and IVH with RVV and RVH. In the low wavenumber region, the I( v ¯ ) and R( v ¯ ) spectra are significantly different and only the spectra in R-format are presented. An advantage in applying isotropic R-spectra is the almost flat baseline in the terahertz region allowing relatively unperturbed observation of any weak modes present.

3. Results and Discussion

3.1. Raman Spectra on Aqueous Ln(ClO4)3 Solutions (Ln = Lu3+, Yb3+, Tm3+, Er3+ and Ho3+)

The perchlorate spectrum, ClO4(aq): The Raman spectrum of the perchlorate ion in aqueous solution (NaClO4(aq)) has been well characterized and only a brief description shall be given [14,15,16]. The ClO4 ion possesses Td symmetry and has nine modes of internal vibrations spanning the representation Γvib(Td) = a1(Ra) + e(Ra) + 2f2(Ra, i.r.). All normal modes are Raman active, but in i.r. only the f2 modes are allowed. The ν1(a1) ClO4 band, centred at 931.5 cm−1, is totally polarized (ρ = 0.005) whereas ν3(f2) ClO4, centred at 1108 cm−1, is depolarized. The deformation modes ν4(f2) ClO4 at 629 cm−1 and ν2(e) ClO4 at 461 cm−1 [15] are also depolarized. In dilute aqueous NaClO4 solutions, the spectrum of ClO4 (aq) shows no sign of contact ion pairs or outer-sphere ion pairs; the ν1(a1) ClO4 band at 931.5 cm−1 is quite narrow with a full width at half height (fwhh) = 7.2 cm−1. However, the ν1(a1) ClO4 band shows an intrinsic low frequency shoulder at 923 cm−1 which is caused by Fermi resonance of the overtone of ν2(e) ClO4, 2x ν2(e) = 923 cm−1 with ν1(a1) ClO4 band [15]. The antisymmetric stretching mode, ν3(f2) ClO4, is much weaker in intensity than ν1(a1) and it appears at 1106 cm−1 with a fwhh = 65 cm−1. In Figure 1A the scattering profiles in R-format (RVV, RVH, and Riso) of a 3.801 mol·L−1 NaClO4(aq) are presented from 50 to 800 cm−1 and in addition the overview Raman scattering profiles (IVV, IVH and Iiso) from 50 to 1800 cm−1 are given in Figure S1 together with the band positions and assignments.
The Raman spectra of Lu(ClO4)3(aq), Yb(ClO4)3(aq), Tm(ClO4)3(aq), Er(ClO4)3(aq) and Ho(ClO4)3(aq): First, we present and discuss the spectroscopic results on Yb(ClO4)3 followed below by the spectra of aqueous Tm(ClO4)3, Er(ClO4)3 and Ho(ClO4)3 solutions. The Raman spectra of aqueous Lu(ClO4)3 solutions have been reported recently [17] with the ν1 LuO8 breathing mode at 396 cm−1.
Figure 1B,C show the scattering profiles of aqueous Yb(ClO4)3 solutions at 0.240 mol·L−1 (Rw = 226.6) and, in comparison, a solution at 2.423 mol·L−1 (RW = 16.86) respectively. The Raman scattering profiles of the 0.240 mol·L−1 aqueous Yb(ClO4)3 solution (Figure 1B) from 40–750 cm−1 shows two ClO4(aq) bands at 461 cm−1 and 629 cm−1 and a broad, weak polarized mode (isotropic scattering) at 394 cm−1 which does not occur in NaClO4(aq). Therefore, the band at 394 cm−1 has to be assigned to the ν1 YbO8 breathing mode of the [Yb(H2O)8]3+ species. In the at 2.423 mol·L−1 (RW = 16.86) Yb(ClO4)3 solution (Figure C), however, the ν1 YbO8 breathing mode is shifted by 4 cm−1 to lower wavenumbers compared to the dilute solution (Figure 1B). An overview Raman spectrum of the 2.423 mol·L−1 Yb(ClO4)3 solution is given in Figure S2 (top panel) from 80–1400 cm−1 which displays all four perchlorate bands and the weak isotropic mode at 390 cm−1 assigned to the ν1 YbO8 breathing mode. The vibrational bands in the anisotropic scattering could only be detected in the concentrated Yb(ClO4)3 solution at 2.423 mol·L−1 because of their very weak and broad nature. These anisotropic bands are even weaker than the already weak ν1 YbO8 breathing mode with an integrated band intensity at 3161. The band fit results for the anisotropic scattering are given in Table S1 and presented in Figure 2 Five bands appear at 88.5 cm−1 (fwhh = 119), 158.7 cm−1 (fwhh = 97.6.9 cm−1), 229.4 cm−1 (fwhh = 75.7 cm−1), 260.0 cm−1 (fwhh = 68cm−1) and 333.2 cm−1 (fwhh = 85.3) in the anisotropic scattering in the 2.423 mol·L−1 Yb(ClO4)3(aq) solution. These weak, broad bands stem from the YbO8 skeleton fundamentals of the [Yb(OH2)8]3+ species and break the symmetry of the YbO8 skeleton. Therefore, they appear only in the anisotropic scattering, but not in the isotropic profile. From group theoretical considerations we expect 7 Raman active modes for the YbO8 skeleton (ligated water molecules seen as point masses) and a brief group theoretical discussion shall be given. The YbO8 skeleton (D4d symmetry) with its 9 atoms leads to 21 normal modes and the irreducible representation follows as: Γv(D4d) = 2a1(Ra) + b1(i.a.) + 2b2(i.r.) + 3e1(i.r.) + 3e2(Ra) + 2e3(Ra). (The YbO8 skeleton possesses no symmetry centre but the mutual exclusion rule nevertheless applies.) Seven modes with the character a1, e2 and e3 are Raman allowed while six modes with the character b1, b2 and e1 are i.r. active. The totally symmetric Yb–O stretch, the breathing mode, is only Raman active and appears strongly polarized in the Raman spectrum as the strongest band of the YbO8 skeleton. Two additional depolarized Raman stretching modes are expected (character e2 and e3) as well as four other Raman deformation modes (character a1, e2 and e3). In infrared, two stretching modes (character b2 and e1) are expected and the remaining are deformations. In reality, however, we observe only six skeleton modes with one unaccounted mode (see also [17]). From our Raman spectroscopic results, it follows directly that the Yb3+–OH2 hydration shell cannot constitute a hexa-hydrate (Th symmetry) which has been, for instance, characterized for [Al(OH2)6]3+(aq) [22,23]. Group theoretical considerations expect only three skeleton modes in Raman for [Al(OH2)6]3+; one of which should be totally polarized (breathing mode for the AlO6 skeleton) and the remaining two depolarized. All of these bands were detected in the Raman spectrum of an Al(ClO4)3(aq) solution with the symmetric stretching mode of [Al(OH2)6]3+ at 525 cm−1 strongly polarized and two bands at 438 cm−1 and 332 cm−1 which are depolarized [22,23].
The concentration dependence of the band parameters (peak positions, full width at half height (fwhh) and integrated band areas) of the ν1 YbO8 breathing mode for Yb(ClO4)3 solutions allows the determination of the change of these band parameters as a function of concentration. The band profiles of the ν1 YbO8 breathing mode are given in Figure 3 at concentrations 0.240 mol·L−1 (RW = 268.74), 0.603 mol·L−1 (RW = 85.17), 0.808 mol·L−1 (RW = 62.13), 1.217 mol·L−1 (RW = 16.86) and finally at 2.423 mol·L−1 (RW = 16.86). The ν1Yb–O breathing mode appears at 394 cm−1 at the lowest concentrations and shifts ~4 cm−1 to lower wavenumbers at the highest concentration. Furthermore, the bandwidths also increase with increasing solute concentration from 52 cm−1 for the 0.240 mol·L−1 solution to 59 cm−1 for the 2.423 mol·L−1 solution (Figure 3). This slight change in band parameters of ν1 YbO8 breathing mode with increasing solute concentration may be due to ion pair formation in concentrated solutions (The ion pairing effect in perchlorate solutions is discussed in detail in [15,17]). The integrated band intensity of the ν1YbO8 breathing mode, A394, rises linearly with the solution concentration. The dependence of the integrated band intensity of the ν1 YbO8 breathing mode of [Yb(OH2)8]3+ as a function of the Yb(ClO4)3 concentration is given in Figure S3 and for the linear relationship follows: A394 = 1303.7·CYb(ClO4)3 (R2 = 99.9).
In addition to the ν1YbO8 of [Yb(OH2)8]3+ an extremely weak and broad band centered at 170 ± 10 cm−1 appears isotropic Raman scattering in of aqueous Yb(ClO4)3 solution (see Figure S2, top panel). This isotropic band is assigned to a restricted translational mode of the weakly H- bonded water molecules (O-H∙∙∙∙OClO3). The mode is strongly anion and concentration-dependent [14,15,16]. The influence of the ClO4 on the water spectrum has been discussed in recent studies on aqueous Ln(ClO4)3 solutions [14,15,16,17]. The Raman scattering profiles, IVV, IVH and Iiso of the O-H stretching region of H2O and its bending mode of a Yb(ClO4)3 solution at 2.423 mol·L−1 and their peak positions are given in Figure S2, bottom panel, and for details see [14,15,16].
The triflate ion (trifluoromethanesulfonate) in aqueous solution acts as an even weaker complex forming anion and is suited, therefore, for studying metal ion hydration. In aqueous solution, however, the weak ν1 band of [Yb(OH2)8]3+ at ~394 cm−1 is overlapped by a strongly polarized triflate band at 319 cm−1 and so band fit analysis was applied. The isotropic Raman spectrum of Yb(CF3SO3)3(aq) at 1.25 mol·L−1 is shown in Figure S4 and the band fit analysis gave two bands with the first band component at 319 cm−1 and the second band at 394 cm−1 (fwhh = 50 cm−1). The first band, a polarized band, stems from CF3SO3 (aq) but the second band is the ν1 YbO8 breathing mode of [Yb(H2O)8]3+. Band parameters and assignments of CF3SO3(aq) modes are given in [16].
The effect of deuteration on the YbO8 skeleton modes of [Yb(OD2)8]3+ was studied in Yb(ClO4)3− D2O solution and resulted in a shift of the Yb–OD2 mode down to 374 cm−1. The shift of ν1 on deuteration is given as ν 1’ = ν 1[m(H2O)/m(D2O)]1/2 = ( 394 cm 1 ) 18.02 / 20.03 = 0.9485 × 394 cm−1 = 373.7 cm−1. (The water and heavy water molecules are taken as point masses.) A Raman spectra of Yb(ClO4)3 solutions in D2O at 0.779 mol·L−1 is presented in Figure 4 and for a concentration at 1.276 mol·L−1 in Figure S5. This isotope shift of the symmetrical stretch of YbO8 in changing from [Yb(OH2)8]3+(H2O) to [Yb(OD2)8]3+(D2O) and the totally symmetric character of the mode, that is, showing a polarization degree ~0, are indeed proof for the character of this mode.
In the Raman spectra of the Tm(ClO4)3, Er(ClO4)3 and Ho(ClO4)3 solutions appear also as strongly polarized bands and were observed at 391, 389 cm−1, and 387 cm−1 respectively. These isotropic bands are unique in these HRE ion solutions and cannot be found in the hydrated ClO4 (aq) spectrum.
The representative Raman spectra of Tm(ClO4)3, Er(ClO4)3 and Ho(ClO4)3 solutions are given in Figures S6–S8, respectively. Several concentrations of the Tm(ClO4)3(aq) were measured but from the coloured Er(ClO4)3 and Ho(ClO4)3 solutions only the dilute solutions could be reliably measured. (Concentrated Er(ClO4)3 and Ho(ClO4)3 solutions absorb the laser light markedly.) The peak positions for the ν1 LnO8 breathing modes for [Lu(OH2)8]3+ (taken from [17]), [Yb(OH2)8]3+, [Tm(OH2)8]3+, [Er(OH2)8]3+ and [Ho(OH2)8]3+ are given in Table 1. Force constant calculations for the ν1 LnO8 breathing modes of this species, applying a simple model, have been carried out according to equation (4):
k M O = 4 π 2 c 2 ν ˜ i 2 N 1 A L ,
with c, the velocity of light, ν ˜ i the wavenumber of the mode i, N the Avogadro constant and AL the molecular weight of the ligand, in our case water. The force constants, kLn–O, calculated for the measured ν1 breathing modes are given in Table 1 together with the corresponding Ln3+- O bond distances [7]. The force constants increase from Lu3+, Yb3+, Tm3+, Er3+ and Ho3+ in the same order as the corresponding Ln–O bond distances decrease, namely Lu–O < Yb–O < Tm–O < Er–O < Ho–O (Figure S9).
Relative scattering intensities, Sh, for the ν1 Ln–O breathing modes are also given in Table 1 and for the definitions of the Sh see ref. [20]. The small scattering intensity values at 0.0156 to 0.0165 for the ν1 Ln–O modes of the HRE octahydrates reflect the fact that the Ln–OH2 bonds possess low polarizability and are hard cations [25]. The accuracy of the scattering coefficient is not better than ± 0.0004 due to the low scattering intensity, the broadness of the modes and the uncertainties in subtracting the baseline. (Note that the Sh value for the totally symmetric stretching mode, ν1Lu–O is 0.0156 ± 0.0004 and the value reported in [17] is too small.)
From ab initio quantum mechanical charge field molecular dynamics studies, the mean Ln–O bond distances (Ln = Ho3+, Er3+, Tm3+, Yb3+ and Lu3+) of the octahydrates, [Ln(OH2)8]3+, average coordination numbers, vibrational frequencies and the corresponding force constants were presented [13]. The authors claimed an “excellent agreement with experimental results” [13] of the computed frequencies with the measured ones in the glassy state [26]. The theoretical force constants for the ν1 breathing modes in [13] deviate considerably from our data in Table 1and do not follow the expected trend given in Figure S9 in going from holmium to lutetium. This trend reflects the steady increase of the force constants of the Ln–O breathing modes with decreasing Ln–O bond distances in going from holmium to lutetium (Table 1; Figure S9). The force constant for ν1 Er–OH2 breathing mode in [13] was given 360 cm−1 equal to the one for the ν1 La–OH2 breathing mode of [LaOH2)9]3+. However, our recently published datum for the ν1 La–OH2 breathing mode [14,15] is with 343 cm−1 much smaller. The calculations in [13] are based on the simplified model of a heteronuclear diatomic species but such an assumption may not be correct. The character of the symmetrical normal mode ν1 of the LnO8 skeleton of the corresponding [Ln(OH2)8]3+ species reveals that the central cations remain stationary and only the water molecules are involved in the breathing motion without disturbing the symmetry and therefore these normal modes are totally polarized.
It is known from kinetic studies [24,27,28] that the water exchange reactions of the [Ln(OH2)8]3+ species for the octahydrates are very fast and these ions are known to be labile. From the rate constants, kex at 25 °C, given in Table 1, follow the water residence times, the time the water molecules reside at these cations. The water residence times are in the range of several nanoseconds (see Table 1) which shows that these ions are indeed quite labile. From the vibration periods of the ν1 Lu–O modes which are at 0.086 to 0.084 ps in going from holmium(III) to lutetium(III) it follows that these species vibrate several hundred thousand times [17] before one water exchange occurs. Although the HRE ions are labile structures, Raman spectroscopy probes the actual structure of these octahydrate species. (It is worth mentioning that the intramolecular bond exchange rate is only a few picoseconds, much faster than the water exchange reaction, therefore for such labile structures for instance [Cu(OH2)5]2+ [27], Raman observes an average structure and, so, a single broad mode appears as a result and at higher peak positions than for comparable divalent metal ions.)

3.2. YbCl3 Solutions

From thermodynamic measurements, it is known that Lu3+ and Yb3+ form weak chloro-complexes/ion-pairs [29,30,31,32]. As a model system for the HRE ion hydrates, aqueous YbCl3 solutions were investigated. The polarized, depolarized and isotropic Raman scattering profiles of a 3.224 mol·L−1 YbCl3 solution compared to a solution at 0.802 mol·L−1 are presented in Figure 5. Furthermore, the isotropic scattering profiles of three YbCl3(aq) solutions at 3.224 mol·L−1 (Rw = 15.64), 1.600 mol·L−1 (Rw = 33.06) and 0.802 mol·L−1 (Rw = 67.55) from 55–700 cm−1 are given in Figure 6. Two YbCl3 solutions in heavy water were also investigated at 0.422 mol·L−1 and at 0.844 mol·L−1 and the overview Raman scattering spectra of a 0.422 mol·L−1 YbCl3 solution in D2O is given in Figure S10. In dilute YbCl3(aq), the Yb3+ ion is fully hydrated indicated by the ν1Yb–OH2 mode at 394 cm−1 for the [Yb(OH2)8]3+ species while in dilute YbCl3 solution in heavy water at 0.422 mol·L−1 (see Figure S10), the ν1Yb–OD2 mode of [Yb(OD2)8]3+ appears at 376 cm−1 due to the vibrational isotope effect (see discussion further above; H2O/D2O are considered point masses).
The ν1 Yb–OH2 stretching mode in the 3.224 mol·L−1 YbCl3(aq) solution, with a mole ratio of solute to water at 1 to 15.64 appears at 389 cm−1 and shifts with dilution to higher frequencies (see Figure 5). In a 0.400 mol·L−1 (Rw = 136.98) YbCl3(aq) solution the ν1 Yb–OH2 breathing mode appears at 394 cm−1 with a fwhh at 52 cm−1 and these band parameters are comparable to the ones in a dilute Yb(ClO4)3(aq) solution in which the fully hydrated [Yb(OH2)8]3+ exists.
A broad isotropic component at 206 cm−1 and a broad feature at 256 cm−1 are also observed. The band at 256 cm−1 is due to the partially hydrated water molecules of the [Yb(OH2)7Cl]2+ species. A Yb-Cl stretching mode should appear at much higher frequencies, namely at ~500 cm−1, but may be very broad and weak and could not be observed (see spectroscopic and DFT results on ZnCl2(aq) [33]). The isotropic component at 206 cm−1 is assigned to the restricted translation band of water of its O-H∙∙∙O/Cl units. These findings are evidence that Cl substitutes water from the first hydration shell of Yb3+ and a partially hydrated Yb3+- chloro-complex formulated as [Yb(OH2)7Cl]2+ is formed.
The integrated band intensities of ν1YbO8 band of the fully hydrated species, [Yb(OH2)8]3+, as a function of concentration was determined from quantitative Raman analysis and it turned out that the integrated band intensity, A394, does not increase linearly with the total YbCl3 concentration (CT). However, a linear increase in band intensity would be expected if the Yb3+ - octahydrate is the only stable species in YbCl3 solution and such a linear relationship was observed in Yb(ClO4)3(aq) solutions (see Figure S3). The measured integrated band intensity of the ν1 YbO8 band in YbCl3 (aq), A394, follows a linear relationship between A394 and CT up to ~0.4 mol·L−1 but then levels off noticeably at higher YbCl3 concentrations (Figure S11). Obviously, above ~0.4 mol·L−1 YbCl3 fractions of the fully hydrated Yb3+ (aq) are converted to a 1:1 Yb3+ chloro-complex species. The existence of higher chloro complexes than 1:1 can be convincingly ruled out taking into account the results of earlier anion exchange studies on aqueous rare earth chloride systems [32]. The mole fractions of both species are plotted in Figure 7. The fraction of the chloro-complex at 29%, in the most concentrated solution, is rather small and the fully hydrated species at 71% is still dominant. With dilution, the fraction of the chloro-complex species diminishes quickly and at ~0.4 mol·L−1 it is zero.
Shown (Figure 5) is the ν1YbO8 symmetric stretching mode at 394 cm−1 in a 0.802 mol·L−1 solution compared to the one at 389 cm−1 in a 3.224 mol·L−1 solution. An additional isotropic band appears at 256 cm−1 in the 3.224 mol·L−1 solution which is due to the stretching mode of the chloro-complex species, [Yb(OH2)7Cl]2+ (details in Riso scattering in the terahertz region see Figure 6). The remaining bands in both panels are due to the water being strongly influenced by the solute at the most concentrated solution. First, in the terahertz region (RVV scattering), weak, broad bands appear at 186 cm−1 in the 0.802 mol·L−1 solution and at 202 cm−1 in the 3.224 mol·L−1 solution assigned to the restricted translational band of water of the O-H∙∙∙O/Cl units. Second, very broad bands (RVV scattering) with peak maxima which appear at 712 cm−1 (0.802 mol·L−1) and 684 cm−1 (3.224 mol·L−1) are due to the librational bands of water. Third, the band at 1272 (0.802 mol·L−1) and 1204 cm−1 (3.224 mol·L−1) are due to overtones of water librations. Finally, the bands at 1645 and 1647 cm−1 respectively are due to the deformation mode of water, ν2 H2O.
The formation of a 1:1 complex with Cl at higher YbCl3 concentrations may be written as:
[ Yb ( OH 2 ) 8 ] 3 +   +   Cl     [ Yb ( OH 2 ) 7 Cl ] 2 +   +   H 2 O
The formation constant for the 1:1 Yb3+-chloro-complex, K1, may be formulated according to Equation (6):
K 1 = [ YbCl 2 + ] [ Yb 3 + ] [ Cl ] f YbCl 2 2 + f Yb 3 + f Cl
with K1′ the “concentration quotient” we get: K 1 = K 1 f YbCl 2 2 + f Yb 3 + f Cl
The concentration quotient can be measured by Raman spectroscopy according to equation (7):
K 1   = ( C T [ Yb 3 + ] ) [ Yb 3 + ] [ Cl ] ,
where CT is the total YbCl3 concentration and the concentrations in brackets denote the equilibrium concentrations of the fully hydrated Yb3+ and Cl. The equilibrium concentration of Yb3+ determined by Raman spectroscopy allows us to calculate K 1 .
The estimated K1 value for chloro complex formation in YbCl3(aq) from K 1 (see ref [17] for details) equal to ca. 0.06 ± 0.015 and a logK1 value at ca. −1.22 follows at 22 °C. (Quantitative Raman spectroscopy applied to these solution spectra with weak and broad low frequency bands is not very precise and therefore a higher uncertainty results.) Data from thermodynamic and spectroscopic studies on YbCl3(aq) solutions confirm the weak nature of the complex species [29,30,31,32]. The results on aqueous LuCl3 solutions and similar rare earth systems [17,29,30,31,32] confirm our findings on YbCl3 solutions. The chloride ion substitutes a water molecule from the flexible first hydration shell of Lu3+ and Yb3+. With dilution, the weak chloro-complex species dissociates and fully hydrated Yb3+(aq) ions detected. This is in contrast to AlCl3(aq) solutions, even in concentrated AlCl3(aq), Cl does not substitute water in the first hydration shell of Al3+ and it is known that the hydration shell of [Al(OH2)6]3+ is quite inert [22,23].
The results of an extensive EXAFS study by Allen and co-workers [34] on 0.1 and 0.01 mol·L−1 Lu3+-, Yb3+ and Tm3+- solutions in 0.20 mol·L−1 HCl and with 14 mol·L−1 LiCl are worthwhile to consider. It could be shown that in solutions with low chloride concentrations, the Ln–O bond distance for Yb3+ is consistent with the fully hydrated Yb3+. In solutions with an excess of LiCl, it was demonstrated that inner sphere chloro-complexation takes place together with a loss of water [34]. Furthermore, a current study in the terahertz frequency range of YbCl3 solutions using FT-IR spectroscopy [35] confirmed weak contact ion pairs as do our recent Raman results on LuCl3(aq) [17]. Choppin and Unrein [36] claimed that only outer-sphere ions pairs exist in lanthanide chloride solutions, but such a view has been questioned [17,29,30,31,32].
To summarize, the [Yb(OH2)7Cl]2+ modes in chloride solutions could be detected and formation of weak chloro-complexes with Yb3+ verified. In dilute solutions (CT < 0.4 mol·L−1) the chloro-complex species disappeared upon dilution and [Yb(OH2)8]3+ and Cl (aq) formed. The chloro-complex formation may be one reason for the data scatter of the recently published Yb–O bond distances and coordination numbers presented for Yb3+(aq) and other rare earth chloride systems [35,36]. In recent experimental structural studies, it was observed that inner-sphere chloro-complex species are formed in aqueous LnCl3 solution (Ln = Lu and Yb) with high chloride concentrations while in dilute solutions, fully hydrated ions exist [17,34,35].

4. Conclusions

Raman measurements on dilute aqueous Lu(ClO4)3, Yb(ClO4)3, Tm(ClO4)3, Er(ClO4)3 and [Ho(H2O)8]3+ solutions have been carried out. In these solutions, strongly polarized modes at 396 cm−1, 394 cm−1, 391 cm−1, 389 cm−1 and 387 cm−1 were detected and assigned to the breathing modes, ν1 Ln–O of the octahydrates [Lu(H2O)8]3+, [Yb(H2O)8]3+, [Tm(H2O)8]3+, [Er(H2O)8]3+ and [Ho(H2O)8]3+,respectively. The force constants of these Ln–O breathing modes were calculated from these vibrational bands. In dilute perchlorate solutions, these species represent the fully hydrated [Ln(OH2)8]3+ ions. The calculated force constants of the octahydrates of [Lu(H2O)8]3+, [Yb(H2O)8]3+, [Tm(H2O)8]3+, [Er(H2O)8]3+ and, [Ho(H2O)8]3+ strengthen with the corresponding decrease of their Ln–O bond distances. The Ln–O bond of the five octahydrates is not very polarizable and therefore the scattering intensities of the ν1 Ln–O bands are small. In Yb(ClO4)3 in heavy water, the Yb–O breathing mode shifts to 375 cm−1 for the deuterated species [Yb(OD2)8]3+ as the result of the vibrational isotope effect changing from H2O to D2O.
As a representative example for the lanthanide perchlorate solutions, higher concentrated aqueous Yb(ClO4)3 solutions were studied and in solutions > 2 mol·L−1 a small fraction of contact ion pairs between Yb3+ and ClO4 were detected. This is supported by the parameters of ClO4-bands (peak position, half width, band shapes) as well as changes of the band parameters of the ν1 Yb–O breathing mode.
In YbCl3 solutions, Cl penetrates into the first hydration sphere of Yb3+(aq) by pushing out a water molecule, and a weak 1:1 chloro-complex species forms. However, the fraction of the chloro-complex diminishes rapidly upon dilution and at a concentration < 0.4 mol·L−1, the chloro-complex species vanished. Our Raman spectroscopic findings were substantiated by recently published EXAFS and terahertz FT-IR and results [34,35].

Supplementary Materials

The following are available online. Table S1. Band fit data on of the anisotropic spectrum of an aqueous 2.423 mol·L−1 Yb(ClO4)3 solution in the terahertz region. Five YbO8 skeleton modes of [Yb(H2O)8]3+(aq) were detected. In addition to isotropic band at 390 cm−1 (fwhh = 59 cm−1) was observed with an integrated band intensity = 3160. Figure S1. Raman scattering profiles (IVV (black), IVH (blue) and Iiso (red)) of a 3.800 mol·L−1 NaClO4(aq) solution. The depolarized modes, ν2(e) at 461 cm−1 and ν2(f2) at 629 cm−1 are the deformation modes of perchlorate, ClO4(aq). The strongly polarized mode, ν1(a1) at 933 cm−1 is the Cl-O symmetric stretching mode and the depolarized band, ν3(f2) at 1110 cm−1 is due to the antisymmetric stretching mode of ClO4 (aq). The broad mode at 1634 cm−1 is due to the deformation mode of H2O. Figure S2. Top panel: Raman scattering profiles (RVV, RVH and Riso) of a 2.423 mol·L−1 Yb(ClO4)3 solutions in water. The left inset shows the symmetric YbO8 stretching mode of [Yb(H2O)8]3+ at 390 cm−1. Note the perchlorate bands at 461 cm−1 and at 629 cm−1 as well as the band at 1112 cm−1. The inset at the right shows the ν1ClO4 mode at its full scale. Shown are the isotropic scattering profile (red), the anisotropic one (blue) and the polarized scattering (black). Bottom panel: Raman spectrum of the 2.423 mol·L−1 Yb(ClO4)3 solution in the wavenumber region from 1400 cm−1 to 4160 cm−1 (spectra in I-format: from top to bottom: polarized (black), depolarized (blue) and isotropic (red) scattering profiles). The deformation mode of water appears at 1623 cm−1 and in the O-H stretching region the water band is modified by the typical ClO4 H2O mode of weakly hydrogen bonded O-H oscillators. The band at 3538 ± 3 cm−1 is due to the stretching bands of the weakly bonded O-H····OClO3 units of H2O/ClO4-. Figure S3. Integrated band intensity, A394, of the symmetric stretching mode ν1YbO8 in arbitrary units as a function of the Yb(ClO4)3 solution concentration. Figure S4. Isotropic Raman scattering profile (R-format) of a 1.25 mol·L−1 Yb(CF3SO3)3 solution. In addition to the strong triflate band at 319.5 cm−1 a very weak and relatively broad band appears at 394 cm−1 which is assigned to the totally symmetric stretching mode of YbO8. Figure S5. Raman scattering profiles of a 1.276 mol·L−1 Yb(ClO4)3 solution in heavy water in order to observe the isotope effect on the symmetric stretching mode of the hydrate respectively deuterated species by changing from [Yb(H2O)8]3+(H2O) to [Yb(D2O)8]3+(D2O). Shown are the isotropic scattering profile (red), the anisotropic one (blue) and the polarized scattering (black). Figure S6. Tm(ClO4)3 solution at 0.315 mol·L−1 (Rw = 226.6) in R-format (spectra from top to bottom: RVV, RVH and Riso). The inset shows the Riso spectrum in greater detail. Note the broad and weak ν1TmO8 stretching mode at 391 cm−1 (fwhh = 53 cm−1) of the [Tm(OH2)8]3+ species. The much larger, depolarized bands at 461 cm−1 and 629 cm−1 are the deformation modes of perchlorate, ClO4(aq). Shown are the isotropic scattering profile (red), the anisotropic one (blue) and the polarized scattering (black). Figure S7. Polarized Raman scattering profile of an aqueous Er(ClO4)3 solution at 0.321 mol·L−1. The weak and strongly polarized mode at 389 cm−1 is assigned to the Er-O symmetric stretching mode of [Er(OH2)8]3+. Note the strong deformation mode, ν2(e) ClO4(aq) at 461 cm−1. Figure S8. Polarized Raman scattering profile of an aqueous Ho(ClO4)3 solution at 0.240 mol·L−1. The weak and strongly polarized mode at 387 cm−1 is assigned to the Ho-O symmetric stretching mode of [Ho(OH2)8]3+. Note the strong deformation mode, ν2(e) ClO4 (aq) at 461 cm−1. Figure S9. The heavy REE ions at the X-axis from left to right: Ho3+, Er3+, Tm3+, Yb3+, and Lu3+. At the left Y-axis the peak positions of the corresponding breathing modes, ν1 Ln-O are given and at the right Y-axis the Ln-O bond distances of the octahydrates, [Ln(OH2)8]3+ [7]. Figure S10. Raman scattering profiles in R-format (RVV (black), RVH (blue) and Riso (red) scattering) of a 0.422 mol·L−1 YbCl3 solutions in heavy water. The YbO8 symmetric stretch of [Yb(D2O)8]3+(D2O) (D2O taken as point mass) is shifted to 376 cm−1 while the same mode in light water is observed at 394 cm−1 demonstrating the vibrational isotope effect on the symmetric stretching mode by changing from [Yb(H2O)8]3+(aq) to [Yb(D2O)8]3+(D2O). Note furthermore the restricted translation band of the O-D···D unit at 182 cm−1, the librational band at 547 cm−1 and the deformation mode ν2 D-O-D at 1205 cm−1. The mode at 1445 cm−1 stems from the small amount of HDO and the band at 1542 cm−1 is a combination band of D2O. Figure S11. The linear dependence of the integrated band intensity, A394 as a function of the Yb(ClO4)3 concentration with A394 = 1303.7 × CT (see Figure S4). The lower curve shows the integrated band intensity of the band at 394 cm−1 in YbCl3 solutions (black squares). The solute concentration is denoted as CT.

Author Contributions

Conceptualization, W.R. and G.I.; methodology, W.R.; software, G.I.; validation, W.R., G.I.; formal analysis, W.R.; investigation, W.R.; resources, W.R.; data curation, W.R.; writing—original draft preparation, W.R.; writing—review and editing, W.R.; visualization, W.R.; supervision, W.R.; project administration, W.R.; funding acquisition, W.R.

Funding

This research received no external funding.

Acknowledgments

We thank J. Kortus of the Institut für Theoretische Physik, TU Bergakademie Freiberg, Germany for his hospitality and permission for the use of the Raman spectrometer. We are grateful to the colleagues of the Institut for their hospitality and stimulating discussions. We particularly wish to acknowledge Frau B. Ostermay for her help in measuring the samples.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Huang, C.H.; Bian, Z. Introduction in Rare Earth Coordination Chemistry: Fundamentals and Applications; Huang, C.-H., Ed.; Wiley: New York, NY, USA, 2010. [Google Scholar]
  2. Marcus, Y. A simple empirical model describing the thermodynamics of hydration of ions of widely varying charges, sizes, and shapes. Biophys. Chem. 1994, 51, 111–127. [Google Scholar] [CrossRef]
  3. Burgess, J. Ions in Solution: Basic Principles of Chemical Interactions; Ellis Horwood, Chichester/Halsted Press: New York, NY, USA, 1988. [Google Scholar]
  4. Habenschuss, A.; Spedding, F.H. The coordination (hydration) of rare earth ions in aqueous chloride solutions from x ray diffraction. I. TbCl3, DyCl3, ErCl3, TmCl3, and LuCl3. J. Chem. Phys. 1979, 70, 2797–2806. [Google Scholar] [CrossRef]
  5. Cossy, C.; Barnes, A.C.; Enderby, J.E.; Merbach, A.E. The hydration of dysprosium(3+) and ytterbium(3+) in aqueous solution: A neutron scattering first order difference study. J. Chem. Phys. 1989, 90, 3254–3260. [Google Scholar] [CrossRef]
  6. Cossy, C.; Helm, L.; Powell, D.H.; Merbach, A.E. A change in coordination number from nine to eight along the lanthanide(lll) aqua ion series in solution: A neutron diffraction study. New J. Chem. 1995, 19, 27–35. [Google Scholar]
  7. Persson, I.; D’Angelo, P.; Panifilis, S.D.; Sandström, M.; Eriksson, L. Hydration of Lanthanoid(III) Ions in Aqueous Solution and Crystalline Hydrates Studied by EXAFS Spectroscopy and Crystallography: The Myth of the “Gadolinium Break”. Chem.-Eur. J. 2008, 14, 3056–3066. [Google Scholar] [CrossRef] [PubMed]
  8. Spezia, R.; Duvail, M.; Vitorge, P.; D’Angelo, P. Molecular dynamics to rationalize EXAFS experiments: A dynamical model explaining hydration behaviour across the lanthanoid(III) series. J. Phys. Conf. Ser. 2009, 190, 012056. [Google Scholar] [CrossRef]
  9. Sessa, F.; Spezia, R.; D’Angelo, P. Lutetium(iii) aqua ion: On the dynamical structure of the heaviest lanthanoid hydration complex. J. Chem. Phys. 2016, 144, 204505. [Google Scholar] [CrossRef]
  10. D’Angelo, P.; Spezia, R. Hydration of Lanthanoids(III) and Actinoids(III): An Experimental/Theoretical Saga. Chem. Eur. J. 2012, 18, 11162–11178. [Google Scholar] [CrossRef] [PubMed]
  11. Duvail, M.; D’Angelo, P.; Gaigeot, M.P.; Vitorge, P.; Spezia, R. What first principles molecular dynamics can tell us about EXAFS spectroscopy of radioactive heavy metal cations in water. Radiochim. Acta 2009, 97, 339–346. [Google Scholar] [CrossRef]
  12. Duvail, M.; Spezia, R.; Vitorge, P.A. Dynamic Model to Explain Hydration Behaviour along the Lanthanide Series. Chem. Phys. Chem. 2008, 9, 693–696. [Google Scholar] [CrossRef] [PubMed]
  13. Passler, P.P.; Rode, B.M. Thulium(III) and ytterbium(III) in aqueous solution ab initio quantum mechanical charge field molecular dynamics studies. Chem. Phys. Lett. 2015, 638, 128–132. [Google Scholar] [CrossRef]
  14. Rudolph, W.W.; Irmer, G. Hydration and ion pair formation in common aqueous La(III) salt solutions–A Raman scattering and DFT study. Dalton Trans 2015, 44, 295–305. [Google Scholar] [CrossRef] [PubMed]
  15. Rudolph, W.W.; Irmer, G. Raman spectroscopic characterization of light rare earth ions: La3+, Ce3+, Pr3+, Nd3+ and Sm3+–hydration and ion pair formation. Dalton Trans. 2017, 46, 4235–4244. [Google Scholar] [CrossRef]
  16. Rudolph, W.W.; Irmer, G. Hydration and ion pair formation in aqueous Y3+–salt solutions. Dalton Trans. 2015, 44, 18492–18505. [Google Scholar] [CrossRef] [PubMed]
  17. Rudolph, W.W.; Irmer, G. Hydration and ion pair formation in aqueous Lu3+ - solution. Molecules 2018, 23, 3237. [Google Scholar] [CrossRef]
  18. Schwarzenbach, G.; Flaschka, H.A. Die Komplexometrische Titration; F. Enke Verlag: Stuttgart, Gremany, 1965. [Google Scholar]
  19. Rudolph, W.W.; Irmer, G. Raman and Infrared Spectroscopic Investigations on Aqueous Alkali Metal Phosphate Solutions and Density Functional Theory Calculations of Phosphate—Water Clusters. Appl. Spectrosc. 2007, 61, 1312–1327. [Google Scholar] [CrossRef]
  20. Rudolph, W.W.; Brooker, M.H.; Pye, C.C. Hydration of Lithium Ion in Aqueous Solutions. J. Phys. Chem. 1995, 99, 3793–3797. [Google Scholar] [CrossRef]
  21. LaPlant, F.; Laurence, G.; Ben-Amotz, D. Theoretical and Experimental Uncertainty in Temperature Measurement of Materials by Raman Spectroscopy. Appl. Spectrosc. 1996, 50, 1034–1038. [Google Scholar] [CrossRef]
  22. Rudolph, W.W.; Mason, R.; Pye, C.C. Aluminium(III) hydration in aqueous solution. A Raman spectroscopic investigation and an ab initio molecular orbital study of aluminium(III) water clusters. Phys. Chem. Chem. Phys. 2000, 2, 5030–5040. [Google Scholar] [CrossRef]
  23. Rudolph, W.; Schönherr, S. Raman-und Infrarotspektroskopische Untersuchungen an konzentrierten Aluminiumsalzlösungen. Z. Phys. Chem. 1989, 270, 1121–1134. [Google Scholar]
  24. Cossy, C.; Helm, L.; Merbach, A.E. Oxygen-17 Nuclear Magnetic Resonance Kinetic Study of Water Exchange on the Lanthanide(II1) Aqua Ions. Inorg. Chem. 1988, 27, 1973–1979. [Google Scholar] [CrossRef]
  25. Pearson, R.G. Hard and Soft Acids and Bases. J. Am. Chem. Soc. 1963, 85, 3533–3539. [Google Scholar] [CrossRef]
  26. Kanno, H. Hydrations of metal ions in aqueous electrolyte solutions: A Raman study. J. Phys. Chem. 1988, 92, 4232–4236. [Google Scholar] [CrossRef]
  27. Helm, L.; Merbach, A.E. Applications of advanced experimental techniques: High pressure NMR and computer simulations. J. Chem. Soc. Dalton Trans. 2002, 5, 633–641. [Google Scholar] [CrossRef]
  28. Purdie, N.; Farrow, M.M. The application of ultrasound-absorption to reaction kinetics. Coord. Chem. Rev. 1973, 11, 189–226. [Google Scholar] [CrossRef]
  29. Wood, S.A. The aqueous geochemistry of the rare-earth elements and yttrium. 1. Review of available low-temperature data for inorganic complexes and the inorganic REE speciation of natural waters. Chem. Geol. 1990, 82, 159–186. [Google Scholar] [CrossRef]
  30. Fernández-Ramírez, E.; Jiménez-Reyes, M.; Solache-Ríos, M.J. Effects of Ionic Strength and Charge Density on the Stability of Chloride Complexes of Trivalent Lanthanides. J. Chem. Eng. Data 2008, 53, 1756–1761. [Google Scholar] [CrossRef]
  31. Fukasawa, T.; Kawasuji, I.; Mitsugahiro, T.; Sat, A.; Suzuki, S. Investigation on the Complex Formation of Some Lanthanoids(III) and Actinoids(III) with chloride and Bromide. Bull. Chem. Soc. Jpn. 1982, 55, 726–729. [Google Scholar] [CrossRef]
  32. Marcus, Y. Anion exchange of metal complexes—XV: Anion exchange and amine extraction of lanthanides and trivalent actinides from chloride solutions. J. Inorg. Nucl. Chem. 1966, 28, 209–219. [Google Scholar] [CrossRef]
  33. Pye, C.C.; Corbeil, C.R.; Rudolph, W.W. An ab initio investigation of zinc chloro complexes. Phys. Chem. Chem. Phys. 2006, 8, 5428–5436. [Google Scholar] [CrossRef] [PubMed]
  34. Allen, P.G.; Bucher, J.J.; Shuh, D.K.; Edelstein, N.M.; Craig, I. Coordination chemistry of trivalent lanthanide and actinide ions in dilute and concentrated chloride solutions. Inorg. Chem. 2000, 39, 595–601. [Google Scholar] [CrossRef] [PubMed]
  35. Klinkhammer, C.; Böhm, F.; Sharma, V.; Schwaab, G.; Seitz, M.; Havenith, M. Anion dependent ion pairing in concentrated ytterbium halide solutions. J. Chem. Phys. 2018, 148, 222802. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Choppin, G.R.; Unrein, P.J. Halide complexes of the lanthanide elements. J. Inorg. Nucl. Chem. 1963, 25, 387–393. [Google Scholar] [CrossRef]
Sample Availability: Samples of the Ytterbium salts are available from the authors.
Figure 1. (A). Raman scattering profiles in R-format (from top to bottom: RVV (black), RVH (blue) and Riso(red)) of a 3.801 mol·L−1 NaClO4(aq) solution (RW = 12.05). Note the broad weak mode at 166 cm−1 is due to the restricted translation of the O-H···H unit of the water molecules. The much larger, depolarized modes at 461 cm−1 and 629 cm−1 are the deformation modes of perchlorate, ClO4(aq). (B). A Yb(ClO4)3 solution spectrum at 0.240 mol·L−1 (Rw = 226.6) in R-format (spectra from top to bottom: RVV (black), RVH (blue) and Riso (red)). The inset shows the Riso spectrum in greater detail. Note the broad and weak ν1YbO8 stretching mode at 394 cm−1 (fwhh = 52 cm−1) of the [Yb(OH2)8]3+ species. The much larger, depolarized bands at 461 cm−1 and 629 cm−1 are the deformation modes of perchlorate, ClO4(aq). (C). Raman scattering profiles of a Yb(ClO4)3 solution at 2.423 mol·L−1 in R-format (spectra from top to bottom: RVV (black), RVH (blue) and Riso (red)). The inset shows the Riso spectrum in greater detail. Note the broad and weak ν1 YbO8 stretching mode at 390 cm−1 of the [Yb(OH2)8]3+ species. The much larger, depolarized bands at 461 cm−1 and 629 cm−1 are the deformation modes of perchlorate, ClO4(aq).
Figure 1. (A). Raman scattering profiles in R-format (from top to bottom: RVV (black), RVH (blue) and Riso(red)) of a 3.801 mol·L−1 NaClO4(aq) solution (RW = 12.05). Note the broad weak mode at 166 cm−1 is due to the restricted translation of the O-H···H unit of the water molecules. The much larger, depolarized modes at 461 cm−1 and 629 cm−1 are the deformation modes of perchlorate, ClO4(aq). (B). A Yb(ClO4)3 solution spectrum at 0.240 mol·L−1 (Rw = 226.6) in R-format (spectra from top to bottom: RVV (black), RVH (blue) and Riso (red)). The inset shows the Riso spectrum in greater detail. Note the broad and weak ν1YbO8 stretching mode at 394 cm−1 (fwhh = 52 cm−1) of the [Yb(OH2)8]3+ species. The much larger, depolarized bands at 461 cm−1 and 629 cm−1 are the deformation modes of perchlorate, ClO4(aq). (C). Raman scattering profiles of a Yb(ClO4)3 solution at 2.423 mol·L−1 in R-format (spectra from top to bottom: RVV (black), RVH (blue) and Riso (red)). The inset shows the Riso spectrum in greater detail. Note the broad and weak ν1 YbO8 stretching mode at 390 cm−1 of the [Yb(OH2)8]3+ species. The much larger, depolarized bands at 461 cm−1 and 629 cm−1 are the deformation modes of perchlorate, ClO4(aq).
Molecules 24 01953 g001aMolecules 24 01953 g001b
Figure 2. Result of a band fit of an anisotropic Raman scattering profile of a Yb(ClO4)3 solution at 2.423 mol·L−1 in R-format. Shown are the measured spectrum, baseline corrected, the sum curve of the band fit and the band components. Underneath is the residue curve, which is the difference of the measured spectrum and the sum curve.
Figure 2. Result of a band fit of an anisotropic Raman scattering profile of a Yb(ClO4)3 solution at 2.423 mol·L−1 in R-format. Shown are the measured spectrum, baseline corrected, the sum curve of the band fit and the band components. Underneath is the residue curve, which is the difference of the measured spectrum and the sum curve.
Molecules 24 01953 g002
Figure 3. Stack plot of five isotropic Raman profiles in R-format of aqueous Yb(ClO4)3 solution. From bottom to top: at 0.240 mol·L−1, 0.603 mol·L−1, 0.808 mol·L−1, 1.217 mol·L−1 and 2.423 mol·L−1.
Figure 3. Stack plot of five isotropic Raman profiles in R-format of aqueous Yb(ClO4)3 solution. From bottom to top: at 0.240 mol·L−1, 0.603 mol·L−1, 0.808 mol·L−1, 1.217 mol·L−1 and 2.423 mol·L−1.
Molecules 24 01953 g003
Figure 4. Raman scattering profiles in R-format (spectra from top to bottom: RVV, RVH and Riso) of a 0.779 mol·L−1 Yb(ClO4)3 solution in D2O. The inset shows the low frequency region in larger detail. The weak band at 375 cm−1 is assigned to the Yb–OD2 mode of the YbO8 skeleton which is shifted due to the isotope effect by changing from H2O to D2O (see also Figure 1B). Note the band at 1204 cm−1 which is due to the deformation mode of D2O.
Figure 4. Raman scattering profiles in R-format (spectra from top to bottom: RVV, RVH and Riso) of a 0.779 mol·L−1 Yb(ClO4)3 solution in D2O. The inset shows the low frequency region in larger detail. The weak band at 375 cm−1 is assigned to the Yb–OD2 mode of the YbO8 skeleton which is shifted due to the isotope effect by changing from H2O to D2O (see also Figure 1B). Note the band at 1204 cm−1 which is due to the deformation mode of D2O.
Molecules 24 01953 g004
Figure 5. Raman scattering profiles (RVV (black), RVH (blue) and Riso (red)) of two YbCl3(aq) solutions from 55–1910 cm−1. Top panel: 0.802 mol·L−1; Bottom panel: 3.224 mol·L−1.
Figure 5. Raman scattering profiles (RVV (black), RVH (blue) and Riso (red)) of two YbCl3(aq) solutions from 55–1910 cm−1. Top panel: 0.802 mol·L−1; Bottom panel: 3.224 mol·L−1.
Molecules 24 01953 g005
Figure 6. Isotropic Raman scattering profiles of YbCl3(aq) solutions from 55–900 cm−1. From bottom to top: 3.224 mol·L−1, 1.600 mol·L−1 and 0.802 mol·L−1. The symmetric YbO8 stretching mode appears at 389 cm−1 in a solution at 3.224 mol·L−1 and is shifted to 393.5 cm−1 in a 0.802 mol·L−1 solution. The band at 256 cm−1 is due to the stretching mode of the chloro-complex species, [Yb(OH2)7Cl]2+. The restricted translation band of the O-H···O/Cl band of water shifts from 206 cm−1 in the 3.224 mol·L−1 solution and appears at 188 cm−1 in the 0.802 mol·L−1 solution. The broad librational band of water shifts from 738 cm−1 to 789 cm−1 in going from a 3.224 mol·L−1 solution to the one at 0.802 mol·L−1.
Figure 6. Isotropic Raman scattering profiles of YbCl3(aq) solutions from 55–900 cm−1. From bottom to top: 3.224 mol·L−1, 1.600 mol·L−1 and 0.802 mol·L−1. The symmetric YbO8 stretching mode appears at 389 cm−1 in a solution at 3.224 mol·L−1 and is shifted to 393.5 cm−1 in a 0.802 mol·L−1 solution. The band at 256 cm−1 is due to the stretching mode of the chloro-complex species, [Yb(OH2)7Cl]2+. The restricted translation band of the O-H···O/Cl band of water shifts from 206 cm−1 in the 3.224 mol·L−1 solution and appears at 188 cm−1 in the 0.802 mol·L−1 solution. The broad librational band of water shifts from 738 cm−1 to 789 cm−1 in going from a 3.224 mol·L−1 solution to the one at 0.802 mol·L−1.
Molecules 24 01953 g006
Figure 7. Fraction of species detected by quantitative Raman spectroscopy. The filled circles denote the [Yb(OH2)8]3+, the fully hydrated Yb3+ and the filled squares the mono chloro-complex species, [Yb(OH2)7Cl]2+. Note the error bar in the right corner of the graph.
Figure 7. Fraction of species detected by quantitative Raman spectroscopy. The filled circles denote the [Yb(OH2)8]3+, the fully hydrated Yb3+ and the filled squares the mono chloro-complex species, [Yb(OH2)7Cl]2+. Note the error bar in the right corner of the graph.
Molecules 24 01953 g007
Table 1. Band parameters such as peak positions, ν1, force constants, kLn–O, the full width at half heights, and the scattering intensities for the Ln–O breathing mode of Ho3+, Er3+, Tm3+, Yb3+ and Lu3+ for the [Ln(OH2)8]3+ species at 22 °C. Furthermore, the literature data for the water exchange [24] and the Ln–O bond distances [7] for these [Ln(OH2)8]3+ species are given.
Table 1. Band parameters such as peak positions, ν1, force constants, kLn–O, the full width at half heights, and the scattering intensities for the Ln–O breathing mode of Ho3+, Er3+, Tm3+, Yb3+ and Lu3+ for the [Ln(OH2)8]3+ species at 22 °C. Furthermore, the literature data for the water exchange [24] and the Ln–O bond distances [7] for these [Ln(OH2)8]3+ species are given.
[Ln(OH2)8]3+ν1 Ln–O /cm−1kLn–O/Nm−1Fwhh/cm−1 Sh k e x 298 / 10 8   s 1   a ) τw/nsLn–O/Å b)
Ho3+387158.97550.01701.915.242.379
Er3+389160.62540.01681.188.482.364
Tm3+391162.27530.01650.8112.342.350
Yb3+394164.77520.01600.4124.392.324
Lu3+396166.45520.0156--2.316
a) Ref. [24]; data at 25 °C. b) Ref. [7]; EXAFS data from L3 edge (one shell fit) on aqueous trifluoromethansulfonate solutions.

Share and Cite

MDPI and ACS Style

Rudolph, W.; Irmer, G. On the Hydration of Heavy Rare Earth Ions: Ho3+, Er3+, Tm3+, Yb3+ and Lu3+—A Raman Study. Molecules 2019, 24, 1953. https://doi.org/10.3390/molecules24101953

AMA Style

Rudolph W, Irmer G. On the Hydration of Heavy Rare Earth Ions: Ho3+, Er3+, Tm3+, Yb3+ and Lu3+—A Raman Study. Molecules. 2019; 24(10):1953. https://doi.org/10.3390/molecules24101953

Chicago/Turabian Style

Rudolph, Wolfram, and Gert Irmer. 2019. "On the Hydration of Heavy Rare Earth Ions: Ho3+, Er3+, Tm3+, Yb3+ and Lu3+—A Raman Study" Molecules 24, no. 10: 1953. https://doi.org/10.3390/molecules24101953

Article Metrics

Back to TopTop