You are on page 1of 16

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/231742158

Curium(III) citrate speciation in biological systems: A europium(III) assisted


spectroscopic and quantum chemical study

Article  in  Dalton Transactions · October 2012


DOI: 10.1039/c2dt31480k · Source: PubMed

CITATIONS READS

26 809

6 authors, including:

Anne Heller Astrid Barkleit


Technische Universität Dresden Helmholtz-Zentrum Dresden-Rossendorf
25 PUBLICATIONS   90 CITATIONS    64 PUBLICATIONS   613 CITATIONS   

SEE PROFILE SEE PROFILE

Harald Foerstendorf Satoru Tsushima


Helmholtz-Zentrum Dresden-Rossendorf Helmholtz-Zentrum Dresden-Rossendorf
77 PUBLICATIONS   1,172 CITATIONS    100 PUBLICATIONS   1,718 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Sorption of actinides and fission product in brine View project

Speciation-dependent cellular uptake and cytotoxicity of tri- and tetravalent lanthanides and actinides in/onto mammalian kidney cells (DFG) View project

All content following this page was uploaded by Satoru Tsushima on 17 May 2014.

The user has requested enhancement of the downloaded file.


Dalton Dynamic Article Links

Transactions
Cite this: Dalton Trans., 2012, 41, 13969
www.rsc.org/dalton PAPER
Curium(III) citrate speciation in biological systems: a europium(III) assisted
spectroscopic and quantum chemical study†
Anne Heller,* Astrid Barkleit, Harald Foerstendorf, Satoru Tsushima, Karsten Heim and Gert Bernhard
Received 6th July 2012, Accepted 3rd September 2012
DOI: 10.1039/c2dt31480k

Citrate complexes are the dominant binding form of trivalent actinides and lanthanides in human urine at
pH < 6. Hence, an accurate prediction of the speciation of these elements in the presence of citrate is
crucial for the understanding of their impact on the metabolism of the human organism and the
corresponding health risks. We studied the complexation of Cm(III) and Eu(III), as representatives of
trivalent actinides and lanthanides, respectively, in aqueous citrate solution over a wide pH range using
time-resolved laser-induced fluorescence spectroscopy. Four distinct citrate complexes were identified and
their stability constants were determined, which are MHCit0, M(HCitH)HCit2−, M(HCit)23−, and
M(Cit)25− (M = Cm, Eu). Additionally, there were also indications for the formation of MCit− complexes.
Structural details on the EuHCit0 and EuCit− complexes were obtained with FT-IR spectroscopy in
combination with density functional theory calculations. IR spectroscopic evidence for the deprotonation
of the hydroxyl group of the citrate ion in the EuCit− complex is presented, which also revealed that the
complexation of the Eu3+ ion takes place not only through the carboxylate groups, like in EuHCit0, but
additionally via the hydroxylate group. In both EuHCit0 and EuCit− the carboxylate binding mode is
mono-dentate. Under a very low metal : citrate ratio that is typical for human body fluids, the Cm(III) and
Eu(III) speciation was found to be strongly pH-dependent. The Cm(III) and Eu(III) citrate complexes
dominant in human urine at pH < 6 were identified to be Cm(HCitH)HCit2− and a mixture of
Eu(HCitH)HCit2− and EuHCit0. The results specify our previous in vitro study using natural human urine
samples (Heller et al., Chem. Res. Toxicol., 2011, 24, 193–203).

Introduction last few decades.2 Consequently, there are several possibilities


how humans potentially may get in contact with both An(III) and
Heavy metals, particularly radionuclides, represent a serious Ln(III).
health risk to humans in the case of incorporation. For the under- Actinide and lanthanide elements are hazardous to health due
standing of their (radio-) toxicity, distribution, deposition and to their radiological and/or chemical toxicity. Irrespective of the
elimination, it is crucial to investigate their aqueous speciation uptake pathway, i.e., inhalation, ingestion or cutaneous absorp-
and molecular transport mechanisms in biosystems. Unfortu- tion, the heavy metal ions are resorbed and transported by the
nately, only little is known about the behavior of trivalent acti- bloodstream prior to deposition in target organs or tissue. For all
nides (An(III)) and lanthanides (Ln(III)) in the human organism. An(III) and Ln(III) these are, in particular, the bones and liver,
An(III) are artificial, highly radioactive elements that are mainly respectively.3–5 Excretion of these elements is very low and
produced within the nuclear fuel cycle in nuclear power plants occurs mainly via the kidneys and, therefore, with urine.3–5 In
and, therefore, are contained in nuclear waste. Due to nuclear general, trivalent actinides and lanthanides exhibit analog chemi-
incidents, natural disasters, or non-professional storage of radio- cal properties due to their similar ionic radii and oxidation states.
active waste, An(III) can be released into the environment and the This, in turn, results also in a very similar behavior of An(III) and
biosphere. In contrast to this, Ln(III) are naturally occurring, non- Ln(III) in the human organism.
radioactive elements with a variety of applications in technology Apart from the well-studied effects of distribution, accumu-
and medicine.1 Especially the use of Ln(III) chelate complexes as lation and elimination, only a few investigations on the molecu-
contrast enhancing agents in magnetic resonance imaging and as lar chemical binding form of these elements in body fluids and
molecular sensors and probes has grown enormously during the tissues have been performed. The speciation of An(III) in human
blood was studied by some authors, who showed that these
metal ions mainly bind to the plasma proteins transferrin and
Helmholtz-Zentrum Dresden-Rossendorf, Institute of Resource Ecology, serum albumin.5,6 Additionally, a significant part of the heavy
P.O. Box 510119, 01314 Dresden, Germany. E-mail: a.heller@hzdr.de;
Fax: +49 351 260 3553; Tel: +49 351 260 2251
metal ions was found to bind to low molecular weight ligands,
† Electronic supplementary information (ESI) available. See DOI: such as phosphate and citrate.5,6 Preliminary investigations on
10.1039/c2dt31480k An(III) in urine suggested citrate complexation of the investigated

This journal is © The Royal Society of Chemistry 2012 Dalton Trans., 2012, 41, 13969–13983 | 13969
elements.7 Recently, we reported on the in vitro speciation of species are (i) citric acid HCitH3, (ii) dihydrogen citrate
curium(III) and europium(III), as representatives of An(III) and HCitH2−, (iii) hydrogen citrate HCitH2−, (iv) citrate HCit3− and
Ln(III), respectively, in human urine.8 We demonstrated that, in (v) citrate with the deprotonated hydroxyl group Cit4−
dependence on the pH, two different dominant species are (cf. Fig. 1a).
formed with each metal ion, out of which the one at slightly Literature data on the complexation of trivalent actinides and
acidic pH ≤ 5.7 is a curium(III) and europium(III) citrate lanthanides with citrate are quite numerous.9,13–27,28,29 While
complex, respectively.8 Hence, in all studies on the speciation of europium(III) is well studied,9,15,19–22,24,25 less work was con-
An(III) and Ln(III) in human body fluids published until now, ducted with curium(III).9,15,18,24,26 Luminescence spectroscopic
citrate is found to be at least partly involved in binding of the data of the complexes formed were reported once for europium(III)9
heavy metal ions. Furthermore, it is known that the presence of but to the best of our knowledge not for curium(III). Further-
citrate also has a large influence on the solubility and, therefore, more, the published literature data are inconsistent (see ESI,
on the mobility of trivalent actinides in aqueous solutions and Fig. S1 and Table S1†). For instance, the stability constant
liquid nuclear waste.9 reported for a certain complex species varies considerably
Citrate is a ubiquitous bioligand occurring in each body fluid, among the literature. In the case of 1 : 1 and 1 : 2 complexes of
not only in blood and urine, but also in sweat, liquor, and europium(III) with the citrate anion HCit3− at I = 0.1 M and
saliva.10 Each citrate molecule has three carboxylic functional- room temperature, the reported complex formation constants
ities as well as one hydroxyl group. Hence, several species exist vary in the range of log K11 = 6.725–8.020 and log K12 = 9.821–
in dependence on the pH. The deprotonation of citric acid is pre- 12.8,20 respectively. Moreover, there is also no agreement to
sented in Fig. 1a along with the respective pKa values.11,12 The what extent the hydroxyl group of the citrate ligand is involved
resulting speciation diagram is shown in Fig. 1b. In accordance in the heavy metal complexation and, thus, various structural
with Eberle and Moattar13 and Ohyoshi and Ohyoshi,14 citric models have been published for An(III) and Ln(III) citrate
acid is regarded as a polybasic acid of the general formula complexes.13,27–29
HxCitHy (x = 0–1, y = 0–3) within this paper. The first proton In the present study, we re-investigated the citrate complexes
reflects that of the hydroxyl group while the latter three represent of curium(III) and europium(III) as representatives of An(III) and
those of the carboxyl groups. Therefore, the five distinct ligand Ln(III), respectively, within a wide pH and ligand concentration
range. Due to the above-mentioned chemical similarities, it is
commonly assumed that Ln(III) mimic An(III) and can be used as
suitable nonradioactive chemical analogs.3,30 Because both
elements exhibit unique luminescence properties, their speciation
was studied with time-resolved laser-induced fluorescence spectro-
scopy (TRLFS). This technique is highly suitable to determine
the binding form of a luminescent metal ion at trace
concentrations.8,31–34 From both single and time-resolved experi-
ments, valuable parameters, such as emission wavelengths and
luminescence lifetimes, are obtained. These parameters are sensi-
tive to the chemical environment of the heavy metal ions and
vary in dependence on the ligand. Furthermore, the number of
water molecules in the first coordination shell of the heavy metal
ions can be estimated from the emission lifetime.35 Structural
details on the formed europium(III) citrate complexes were
obtained by Fourier-transform infrared (FT-IR) spectroscopy as
well as quantum mechanical calculations using the density func-
tional theory (DFT). This combined approach is highly suitable
to determine the functional group(s) of the ligand actually
binding the Eu3+ ion and the solution structure of the complex
species. Furthermore, the question whether the hydroxyl group
participates in the Eu3+ binding is addressed. The presented data
provide a comprehensive image of this complex ligand system
and the various species formed with curium(III) and europium(III)
in aqueous solutions with implications for their biological
relevance.

Experimental section
Solutions and reagents
Fig. 1 Stepwise deprotonation of citric acid (HCitH3) with the respective
pKa values for I = 0.1 M (a) and the resulting speciation diagram of 0.1 M Water free citric acid (≥99.5%, p.a., ACS) was purchased from
citric acid in aqueous solutions at I = 0.1 M and room temperature (b). Roth, EuCl3·6H2O (99.999) from Sigma, NaClO4 ( p.a.),

13970 | Dalton Trans., 2012, 41, 13969–13983 This journal is © The Royal Society of Chemistry 2012
NaCl ( p.a.), HCl ( p.a.) and NaOH ( p.a.) from Merck. Stock spectrometer from Bruker Optics, equipped with a mercury
solutions of citric acid were freshly prepared for each series of cadmium telluride detector. Details on the experimental setup are
measurement by dissolving in water and europium(III) stock solu- given elsewhere.37 The used ATR unit (DURA SamplIR II from
tions were prepared by dissolving EuCl3 in water. In the case of Smiths Inc.) is a horizontal diamond with nine internal reflec-
curium(III) a 2.36 × 10−4 M stock solution of the long-lived tions on the upper surface and an incidence angle of 45 °. An
248
Cm isotope (half-life = 3.4 × 105 years) in 1 M HClO4 ATR flow cell with a volume of 200 μL was used to ensure
was used. This radionuclide was supplied by the U.S. adequate background subtraction without external thermal inter-
Department of Energy, Office of Basic Energy Sciences, via the ference. Spectra were recorded in the range of 1800–800 cm−1 at
transplutonium element production facilities at Oak Ridge a spectral resolution of 4 cm−1. All experiments were performed
National Laboratory. A stock solution of 3 M NaClO4 (TRLFS under normal atmosphere at 24 ± 1 °C and a constant ionic
measurements) or 3 M NaCl (IR measurements) was used as the strength of 0.1 M (NaCl). Due to the high specific radioactivity
background electrolyte to maintain constant ionic strength in the of curium(III), IR measurements were only performed with
experiments. europium(III).
All stock solutions and samples were prepared in threefold As a first step, the spectral properties of the pure ligand were
distilled and carbonate free water. The pH of the samples was studied at pH 1 to 13.5. The complexation with europium(III)
measured with glass electrodes (InLab 427 combination pH was then investigated at a fixed metal to ligand ratio of 1 : 1
puncture electrode from Mettler-Toledo and BlueLine 16 pH (0.01 M) in the same pH range. In the case of the absorption
electrode from Schott, respectively) and corrected with spectra, a spectrum of pure water was used for background cor-
NaOH and HClO4 (TRLFS measurements) or HCl (IR rection. Difference spectra were calculated from single beam
measurements). spectra of the solutions containing either complexes or the pure
citric acid. Positive and negative bands in the difference spectra
represent vibrational modes of the complexes and the pure
TRLFS measurements ligand, respectively.
It should be noted that within IR measurements a colorless,
Time-resolved laser-induced fluorescence spectroscopy was per-
amorphous precipitation occurred in europium(III) citrate
formed using a pulsed flash lamp pumped Nd:YAG-OPO laser
solutions at pH 4 to 6. Using various techniques for elemental
system (Powerlite Precision II 9020 laser equipped with a Green
and thermal analysis, this precipitate was identified to be
PANTHER EX OPO) from Continuum. Details on the experi-
most probably of the empirical formulae EuHCit0·H2O.
mental setup are given elsewhere.36 The laser pulse energy was
Nevertheless, only IR spectra of the liquid phases are presented.
monitored using a photodiode and varied between 2 and 3 mJ.
The emission spectra were detected by an optical multichannel
analyzer consisting of a monochromator (Oriel MS 257), a spec-
trograph with different gratings (300 and 1200 lines per mm, Data analysis
respectively), and an ICCD camera (Andor iStar); all com-
ponents were purchased from the Lot-Oriel Group. A constant All absorption (IR) and emission (TRLFS) spectra were analyzed
time window of 1 ms and an excitation wavelength of 395 nm with the Origin 7.5G software38 to obtain peak positions.
were applied for all measurements. Curium(III) and europium(III) TRLFS spectra were baseline corrected and the steady-state
steady-state and time-dependent luminescence spectra were spectra were normalized. Since curium(III) exhibits a degenerated
recorded in the 570–640 nm (grating = 1200 lines per mm with ground state,39 this was applied to the intensity ( peak area) of
0.2 nm resolution) and 450–750 nm (grating = 300 lines per mm the whole spectrum. In the case of europium(III) normalization
with 0.7 nm resolution) ranges, respectively. For time-resolved was applied only to the peak area of the 5D0 → 7F1 band
measurements a total of 35 spectra with delay steps of 10–100 μs because the luminescence to this ground state is considered to be
were recorded per sample. independent from the chemical environment of the metal
In the case of curium(III) all experiments were run in a glove ion.32,40 Additionally, in the case of europium(III), the intensity
box under a nitrogen atmosphere, whereas europium(III) ratio (RE/M) between the electric dipole transition 5D0 → 7F2 and
measurements were carried out under nitrogen as well as normal the magnetic dipole transition 5D0 → 7F1 was determined as
atmosphere. All experiments were performed at 24 ± 1 °C and a follows:
constant ionic strength of 0.1 M (NaClO4). The complexation
reaction was studied at a fixed heavy metal concentration of RE=M ¼ Ið7 F2 Þ=Ið7 F1 Þ; ð1Þ
3 × 10−5 M europium(III) and 3 × 10−7 M curium(III), respecti- with I being the luminescence intensity ( peak area) of the
vely, by varying the citrate concentration from 10−5 to 0.1 M respective transition. Furthermore, Origin was also used to deter-
and the pH from 2 to 13. All samples were prepared directly in mine the luminescence lifetime of emitting species according to
the cuvette and the measurements were performed as spectro- the equation of exponential decay:
photometric titrations.
IðtÞ ¼ Σi I i expðt=τ i Þ: ð2Þ

ATR FT-IR measurements In eqn (2), I is the total luminescence intensity at the time t, Ii
the luminescence intensity of species i at t = 0, and τi the corre-
Attenuated total reflection Fourier-transform infrared spec- sponding lifetime. With this lifetime, the number of water mole-
troscopy was performed using a Vertex 80/v vacuum cules in the first coordination sphere of the heavy metal ions was

This journal is © The Royal Society of Chemistry 2012 Dalton Trans., 2012, 41, 13969–13983 | 13971
calculated using the following equations given by Kimura Additionally, the Gibbs energy of the reaction according to
et al.35 in which the lifetime is inserted in ms: eqn (5)–(8) was then determined as follows:

ΔR G 0 ¼ RT ln K 0 : ð12Þ
nH2 O + 0:5 ¼ 0:65=τ  0:88 for curiumðiiiÞ ð3Þ
Speciation calculations were carried out with Hyss200649
nH2 O + 0:5 ¼ 1:07=τ  0:62 for europiumðiiiÞ: ð4Þ using the same parameters as in Specfit and HypSpec as well as
the calculated stability constants (log β) of the heavy metal
Complex formation constants and emission spectra of the citrate complexes.
single species were determined independently using Specfit41
and HypSpec.42 These factor analysis programs decompose
spectra of mixtures into their components based on the spectro- DFT calculations
scopic properties of each species, which vary in dependence on
the pH or the ligand concentration. Furthermore, they calculate DFT calculations were performed using Gaussian 03.50
reasonable complex stability constants. The successful appli- Geometries were optimized in the aqueous phase at the B3LYP
cation of this software was demonstrated by earlier work on the level using the CPCM solvation model51 with UAHF radii.52
complexation of curium(III), europium(III), and americium(III) The large core effective core potential as well as the corres-
with various ligands.31,33,36,43 Input parameters for the data fitting ponding basis set suggested by Dolg et al.53 was used on
were the total concentrations of the heavy metals and citrate, the europium(III). For C, O, and H, the all-electron valence triple-ζ
stability constants of the heavy metal hydroxides,44 the protonation basis set plus polarization and diffuse functions have been
constants of the ligand species,11,12 the pH, and the measured used.54 More details of the DFT calculations have been described
sample spectra. For all constants, the values valid for I = 0.1 M in a previous publication.55
have been used and, where necessary, recalculated from I = 0 M
using the Davies equation. The measured luminescence spectra of
the Cm3+ and Eu3+ aqua ions were used as fixed normative spectra. Results and discussion
Based on the identified species, the formation constants for TRLFS spectra of europium(III) citrate complexes
the direct reactions, log Kpqr ( p = number of M3+ ions, q =
number of ligand molecules, r = number of protons), were calcu- The citrate ligand shows no luminescence in the investigated
lated according to the mass action law for the following wavelength range (570–640 nm), whereas the Eu3+ ion exhibits a
equations: characteristic emission spectrum. Hence, there is no interference
with the ligand and the detected luminescence results solely
M3þ þ HCit3 ! MHCit0 for log K 111 ð5Þ from the trivalent europium. To enable the complexation by
different citrate species (cf. Fig. 1a), europium(III) luminescence
spectra were measured at a constant metal concentration of
M3þ þ HCitH2 þ HCit3 ! MðHCitHÞHCit2 for log K 123 3 × 10−5 M and pH 2.0, 4.0, 5.5, 8.5, and 13.0 with varying
ð6Þ ligand concentrations of 10−5–10−1 M (Fig. 2).
The luminescence spectrum of the Eu3+ aqua ion is character-
M3þ þ 2HCit3 ! MðHCitÞ2 3 for log K 122 ð7Þ ized by emission bands at 585–600 nm and 610–625 nm repre-
senting the magnetic dipole transition 5D0 → 7F1 and the
hypersensitive electric dipole transition 5D0 → 7F2, respecti-
M3þ þ 2Cit4 ! MðCitÞ2 5 for log K 120 ¼ log β120 : ð8Þ
vely.8,32,56,57 The intensity ratio (RE/M) according to eqn (1) is
Extrapolation of the complex formation constants for zero 0.25–0.5 and the luminescence lifetime is 110 ± 10 μs corre-
ionic strength, log K0, was calculated with ionic strength sponding to 9 water molecules in the first hydration sphere of the
corrections using Specific Interaction Theory (SIT).45 The SIT Eu3+ ion.8,32,56,57 Spectroscopic changes indicating the com-
parameters of Eu3+ and citrate were taken from Vercouter et al.46 plexation of trivalent europium were observed at each pH. In
and De Stefano et al.,47 respectively. In the case of Cm3+, general, luminescence wavelengths remain nearly unaltered upon
no SIT parameter exists so far. Nevertheless, the SIT parameter complexation and a shift of the emission maxima was observed
of Nd3+, which is similar to almost all Ln(III), is recommended only for exceptional cases. The luminescence spectra of all
to be also used for Am3+.48 Therefore, this value was taken for europium(III) citrate complexes exhibit an additional emission
Cm3+. The log βpqr values for the overall reactions were calcu- band at around 578 nm corresponding to the Laporte forbidden
lated according to the mass action law for the following transition 5D0 → 7F0, which occurs only upon deformation of
equations: the spherical symmetry of the Eu3+ aqua ion. Furthermore, the
7
F1 and 7F2 bands are split and the RE/M is significantly enhanced.
M3þ þ Cit4 þ Hþ ! MHCit0 for log β111
Finally, also the emission lifetime prolongs upon complexation.
ð9Þ
In each sample, we observed mono-exponential luminescence
decay. All spectroscopic changes are summarized in Table 1.
M3þ þ 2Cit4 þ 3Hþ ! MðHCitHÞHCit2 for log β123 ð10Þ In the acidic region at pH 2.0 (cf. Fig. 2a), the luminescence
spectrum of europium(III) remains unaltered up to a ligand con-
centration of 2.5 × 10−4 M and all parameters correspond to
M3þ þ 2Cit4 þ 2Hþ ! MðHCitÞ2 3 for log β122 : ð11Þ those of the uncomplexed Eu3+ aqua ion. At higher citrate

13972 | Dalton Trans., 2012, 41, 13969–13983 This journal is © The Royal Society of Chemistry 2012
Fig. 2 Normalized steady-state luminescence spectra of 3 × 10−5 M europium(III) + citrate at pH 2.0 (a), 4.0 (b), 5.5 (c), 8.5 (d), and 13.0 (e), I = 0.1 M,
and room temperature in dependence on the ligand concentration (inset: enlargement of the 5D0 → 7F0 transition band) as well as the deconvoluted
emission spectra (f ) of the single europium(III) citrate complexes (Eu3+ without citrate = grey line).

Table 1 Luminescence spectroscopic parameters of the determined europium(III) species

λ/nma

pH [citrate]/M 7
F0 7
F1 7
F2 RE/M b τ/μs nH2Oc Assigned species Reference

1–6 0 — 591.7 616.4 0.5 110 ± 10 9.1 Eu3+ aqua ion Present work
2.0 >5 × 10−4 578.7 590.5/594.0 614.1/616.7 1.6 >145 <6.8 EuHCit0 Present work
4.0 ≤10−3 578.7 590.5/593.8 614.0/616.8 1.5 168 ± 4 5.8 EuHCit0 Present work
5.5 <4 × 10−4 578.7 590.5/593.9 614.0/616.8 1.5 162 ± 10 6.0 EuHCit0 Present work
4.0 >10−3 578.9 590.6/595.1 614.1/617.2 2.2 >245 <3.8 Eu(HCitH)HCit2− Present work
5.5 ≥4 × 10−4 578.9 590.5/595.0 614.0/617.3 2.0 250 ± 5 3.7 Eu(HCitH)HCit2− Present work
8.5 >5 × 10−4 579.3 589.5/595.3/589.9 614.8/619.7 2.4 490 ± 20d 1.6d Eu(HCit)23− Present work
13.0 ≥10−3 578.6 589.1/598.5 612.5/620.2 1.8 676 ± 5 1.0 EuCit25− Present work
≤5.7 — 578.7 590.2/594.5 613.9/616.7 2.2 153 ± 19 6.4 Eu(III) in human urine 8
a
Wavelengths of luminescence in the specified ground states with a standard deviation of ±0.1 nm. b Intensity ratio of the 7F2 (electric dipole
transition) over the 7F1 (magnetic dipole transition) luminescence band with a standard deviation of ±0.2 nm. c Number of water molecules calculated
according to Kimura et al.35 with a standard deviation of ±0.5. d Luminescence spectra show that Eu(HCit)23− and a hydrolyzed europium(III) species
co-exist at this pH. Therefore, both species contribute to the measured lifetime.

This journal is © The Royal Society of Chemistry 2012 Dalton Trans., 2012, 41, 13969–13983 | 13973
concentrations, the 7F0 peak evolves at 578.7 ± 0.1 nm. Further- 1.0 ± 0.5 water molecules. This indicates the formation of
more, at ≥5 × 10−3 M citrate, the RE/M is enhanced up to 1.6 ± another Eu(III) citrate species 4 at this pH.
0.1 and the luminescence lifetime is prolonged to >145 μs indi- In summary, four different europium(III) citrate species were
cating the formation of a Eu(III) citrate species 1. determined and characterized by varying the citrate concentration
At pH 4.0 and 5.5 (cf. Fig. 2b and 2c), the same spectral at pH 2.0–13.0. The estimation of the number of water molecules
changes were observed even with the smallest ligand addition. In left in the first coordination shell of the Eu3+ ion and literature
the range of 10−5–10−3 M citrate at pH 4.0 and 10−5–4 × 10−4 M data on several europium(III) citrate complexes9,15,19–22,24,25
at pH 5.5, all luminescence spectra are very similar with an RE/M allow a tentative assignment of the four europium(III) citrate
of 1.5 ± 0.1 and a luminescence lifetime of 165 ± 10 μs. Accord- species.
ing to eqn (4), this corresponds to a number of 5.9 ± 0.5 water The Eu(III) citrate species 1 is identified as the charge neutral
molecules in the first hydration shell of the Eu3+ ion. All spectro- 1 : 1 complex EuHCit0 with the citrate anion HCit3−. The lumi-
scopic parameters equal those of the Eu(III) citrate species 1 nescence lifetime of 165 ± 10 μs equals 3.2 ± 0.5 water mole-
determined at pH 2.0. Citrate concentrations of >10−3 M at pH cules that have been replaced by the ligand upon complexation.
4.0 and >4 × 10−4 M at pH 5.5 result in a further enhancement The Eu(III) citrate species 2 is found to be the 1 : 2 complex
of the RE/M to 2.1 ± 0.1 and a prolonged luminescence lifetime Eu(HCitH)HCit2−. In this species, one hydrogen citrate HCitH2−
of 250 ± 5 μs corresponding to 3.7 ± 0.5 water molecules in the and one citrate anion HCit3− bind to the Eu3+ ion. The lumines-
first coordination sphere of the trivalent europium. This indicates cence lifetime of 250 ± 5 μs indicates that 5.3 ± 0.5 water mole-
the formation of a new Eu(III) citrate species 2 at both pH values cules have been displaced by the ligands. The Eu(III) citrate
and a large excess of citrate. Furthermore, both complex species species 3 is assigned to the 1 : 2 complex Eu(HCit)23−. The
differ in the shape of the 7F0 peak. In the case of the Eu(III) citrate anion HCit3− is the exclusive ligand species at neutral to
citrate species 1, the sharp and quite intense peak occurs at alkaline pH and protonation upon complexation is very unlikely.
578.7 nm. In contrast, for the Eu(III) citrate species 2, it occurs at Furthermore, the luminescence lifetime of 490 ± 20 μs corre-
578.9 nm and is of much less intensity as well as broader sponds to a replacement of 7.5 ± 0.5 water molecules and, there-
(cf. insets of Fig. 2b and 2c). fore, also supports the binding of two ligand molecules, which
At pH 8.5 (cf. Fig. 2d), due to metal ion hydrolysis,56–58 spec- each displace at least three water molecules. Finally, Eu(III)
tral changes occur already in the europium(III) solution without citrate species 4 is found to be the 1 : 2 complex EuCit25−. This
citrate. Ligand addition induces further alterations, such as the results, first of all, from the fact that the complexation occurs
fine structures of the 7F1 and 7F2 bands being substantially only at high citrate excess indicating the binding of more than
different from those of the hydrolyzed Eu3+ ion or those of the one ligand molecule. Moreover, the luminescence lifetime of
two complex species determined at lower pH. This indicates 676 ± 5 μs is very long and equals the displacement of 8.1 ± 0.5
the formation of another Eu(III) citrate species 3. Strikingly, the water molecules by the ligand. However, also the hydrolysis of
splitting of the 7F1 band is enhanced to four single peaks upon water molecules from the first coordination sphere of the Eu3+
complexation. Since the 7F1 ground state of the Eu3+ ion can ion can not be excluded at such alkaline pH. Up to now, there
only be split into three Stark levels at the most,59 four single are no reference data on europium(III) citrate species at pH > 10.
peaks point to the simultaneous existence of two different euro- Finally, two series were measured at a constant metal to ligand
pium(III) species. Since one of the split peaks (592.5 nm) signifi- ratio (equimolar and with ligand excess) by varying the pH from
cantly decreases in intensity with increasing ligand concentration 2.0 to 12.0 (see ESI, Fig. S2†). In the series of ligand excess, all
and nearly disappears at the highest concentration of 10−3 M four complex species were identified and there were no indi-
citrate, it does not belong to the spectrum of the Eu(III) citrate cations for metal ion hydrolysis. At an equimolar metal to ligand
species 3 but to that of the hydrolyzed europium(III) species. ratio, only the EuHCit0 complex and an additional, yet uniden-
This demonstrates the competition of citrate complexation and tified, species were detected. In addition, metal ion hydrolysis
metal hydrolysis at this pH value. Furthermore, the measured occurred at very alkaline pH. Taking the equimolar metal to
emission lifetime is initially prolonged with increasing citrate ligand ratio into account, the unknown species is most probably
concentrations of ≤10−4 M (hydrolysis + citrate complexation) the 1 : 1 complex EuCit−, where the fully deprotonated citrate
but shortened at higher ligand concentrations ( predominantly anion Cit4− binds to the Eu3+ ion. Analog complexes are
citrate complexation). At the highest ligand concentrations, the reported for curium(III), americium(III), gadolinium(III), and neo-
lifetime is 490 ± 20 μs corresponding to 1.6 ± 0.5 water mole- dymium(III).13,17,23,26,29 All pH-dependent measurements
cules in the first hydration shell of the Eu3+ ion. demonstrate that the determined complexes subsequently equili-
At pH 13.0 (cf. Fig. 2e), no luminescence signal was detected brate with each other. In conclusion, the An(III) and Ln(III) citrate
in pure europium(III) solution and after the addition of ≤2.5 × species predominantly reported in the literature at the several pH
10−4 M citrate. This is due to the formation of the poorly water ranges and metal to ligand ratios were identified for europium(III)
soluble Eu(OH)3, which dominates the europium(III) speciation and the respective TRLFS spectra were assigned (see Table 1).
at high pH.57 At ligand concentrations exceeding 5 × 10−4 M The single component spectra are presented in Fig. 2f.
citrate, steady-state and time-resolved spectra are evaluable indi- To date, only one luminescence spectroscopic work on the
cating the partial re-solution of Eu3+ ions due to complexation complexation of europium(III) with citrate has been published.
with citrate. In contrast to the spectra at all other pH values, the Mathur et al. reported 7F0 → 5D0 excitation spectra for 10−4 M
luminescence intensity of the 7F2 band decreases with increasing europium(III) + 0.003 and 0.008 M citrate, respectively, at pH 3.6
ligand concentrations. At ≥10−2 M citrate, the RE/M is 1.8 ± 0.2 to 9.0 and a very high ionic strength of 6.6 M NaClO4.9 They
and the lifetime is prolonged to 676 ± 5 μs corresponding to detected a slight red shift of the excitation band from 579.1 nm

13974 | Dalton Trans., 2012, 41, 13969–13983 This journal is © The Royal Society of Chemistry 2012
at pH 3.6 to 579.9 nm at pH 9.09 comparable to that of the 7F0 different pH values ranging from 2 to 13.5 (Fig. 3). These
peak reported in this paper (cf. Table 1). Emission spectra poten- spectra serve as a reference for the interpretation of the spectra
tially serving as references for this work were not published. In of the europium(III) complexes. In general, the IR spectra of the
general, the lifetimes reported by Mathur et al. prolong with ligand species are in good agreement with published literature
increasing pH, which is in agreement with our findings (cf. data,60 and the tentative assignments of the vibrational bands are
Table 1). The lifetime of EuHCit0 determined in the present given in Table 3. All spectra are dominated by bands represent-
work is in very good agreement with the value reported by ing the carboxyl groups, observed at 1722 and 1226 cm−1 in the
Mathur et al.9 Regarding the identification of the other europium(III) spectra recorded at pH ≤ 4, and the carboxylate groups, observed
citrate species, discrepancies occur upon interpretation of the around 1568 and 1390 cm−1 at higher pH. At pH ≥ 6, no bands
data. Whereas in our work three complexes were determined in representing carboxyl groups are observed anymore. The spec-
the range of pH 3 to 9, Mathur et al. solely account for com- trum recorded at pH 13.5 shows an additional band at
plexation by the citrate anion HCit3–9 and, therefore, only for the 1016 cm−1, which is due to the deprotonation of the hydroxyl
formation of EuHCit0 and Eu(HCit)23− but not for that of the group ( pKa = 13.511). To the best of our knowledge, an IR
mixed Eu(HCitH)HCit2−. Moreover, the reported lifetimes for absorption spectrum, which evidences the deprotonation of the
Eu(HCit)23− vary significantly between 213 and 640 μs, which hydroxyl group in citric acid, has not been published before.
was explained by varying numbers of water molecules in the
first hydration shell of the Eu3+ ion.9 A comparison of the litera-
ture lifetimes with those determined in this study is presented in IR spectra and DFT calculations of europium(III) citrate
Table 2. On the basis of this comparison, it is reasonable to complexes
suppose that not only the EuHCit0 and Eu(HCit)23− species were
formed within the study of Mathur et al.9 but also the mixed The IR spectra of equimolar aqueous solutions of europium(III)
Eu(HCitH)HCit2− and possibly also the EuCit25− complex. and citrate recorded in the range of pH 2–12 are shown in

IR spectra of citrate

Prior to the measurement of the europium(III) citrate complexes,


the vibrational spectra of the pure ligand were recorded at

Table 2 Comparison of the luminescence lifetime of europium(III)


citrate species reported in the literature with those determined within this
study

Species determined in the


Species reported by Mathur et al.9 present work

Formulaea Lifetime/μs Lifetime/μs Formulae

EuHCit(H2O)60 180 ± 6 165 ± 10 EuHCit0


Eu(HCit)2(H2O)33− 213 ± 8 250 ± 5 Eu(HCitH)HCit2−
Eu(HCit)2(H2O)3− pH 7 470 ± 18 490 ± 20 Eu(HCit)23−
Eu(HCit)2(H2O)3− pH 9 636 ± 10 676 ± 5 EuCit25−
a
Renamed according to the ligand species used in the present work. Fig. 3 IR absorption spectrum of 10−2 M citrate at I = 0.1 M (NaCl)
and room temperature in dependence on the pH.

Table 3 Tentative assignment of the IR absorption bands of the ligand species

Wavenumber/cm−1 a

HCitH3 (pH 1–2) HCitH2−/HCitH2− (pH 4–6) HCit3− (pH 8–12) HCit3−/Cit4− (pH 13.5) Tentative assignmentb

1722 (s) 1722 (m) — — νs(HOCvO)


1633 (w) — — — δ(H–O–H)
— 1576 (s) 1568 (s) 1568 (m) νas(COO−)
1399 (w) 1396 (s) 1390 (s) 1390 (m) νs(COO−)
— — 1292 (w) — δ(OvC–O−) or δ(H–C–H) + δ(C–O–H)
— 1280 (w) 1280 (m) — ν(C–OH) + δ(C–O–H)
— — 1258 (w) — δ(OvC–O−) or δ(H–C–H) + δ(C–O–H)
1226 (s) 1226 (m) — — ν(OC–OH) + δ(C–O–H)
— 1106 (w) 1095 (w) — ν(C–OH)
— — — 1016 (s) ν(C–O−)
a
s = strong, m = medium, w = weak. b ν = stretching vibration, νs/νas = symmetric/antisymmetric stretching vibration, δ = deformation vibration.

This journal is © The Royal Society of Chemistry 2012 Dalton Trans., 2012, 41, 13969–13983 | 13975
Fig. 4a. In general, the absorption spectra potentially show con- the difference spectra changes up to pH 6, whereas at pH ≥ 8 no
tributions from uncomplexed citric acid. Therefore, difference significant spectral alterations are observed either in the absorp-
spectra were calculated from the respective single beam spectra tion or the difference spectra. Hence, from both series of spectra,
of the solutions containing europium(III) citrate complexes and two different europium(III) citrate complexes can be derived. The
the pure ligand (Fig. 4b). In these spectra, only modes of the first species dominates at low pH, while the second one is predo-
ligand undergoing alterations upon complexation are detected. minantly formed at higher pH. The spectra recorded at pH 6
Hence, negative bands in the difference spectra correspond to obviously show contributions from both complex types dominat-
vibrational modes of the pure ligand while positive ones re- ing the speciation at lower and higher pH values.
present the respective complex species. Table 4 summarizes the In the acidic region, at pH 2 and 4, the predominant species is
tentative assignments of the difference bands to the functional designated as the EuHCit0 species according to the TRLFS
groups. results. IR spectra of this complex are characterized by two
At pH ≤ 4, the presence of carboxyl groups is evidenced by strong bands representing vibrational modes of carboxylate
the absorption and difference spectra showing positive and nega- groups observed at 1600 and 1570 cm−1 for the ν3,as(COO)
tive bands respectively, at 1722 and 1226 cm−1, indicating con- mode and at 1417 and 1397 cm−1 for the ν3,s(COO) mode. The
tributions of uncomplexed citric acid molecules. The shape of splitting of these modes in the spectra of the europium(III)

Fig. 4 IR absorption (a) and difference (b) spectra of 10−2 M citrate + 10−2 M europium(III) at I = 0.1 M (NaCl) and room temperature in depen-
dence on the pH.

Table 4 Tentative assignment of the ATR-FT-IR absorption and difference bands of europium(III) citrate complex species (all data as wavenumber/
cm−1)

EuHCit0 (pH 2–4) Mixture (pH 6) EuCit− (pH 8–12)

Absa Diffb Absa Diffb Absa Diffb Tentative assignmentc Ligand/complex

1719 (m) 1720 (−) — 1711 (−) — — νs(HOCvO) Ligand


1600 (sh) 1600 (+) — 1604 (+) — 1579 (−) νas(COO−) Complex binding
1570 (s) 1568 (+) 1565 (s) 1550 (+) 1564 (s) 1550 (+) νas(COO−) Complex non-binding
1417 (s) 1419 (+) 1430 (sh) 1427 (+) 1425 (sh) 1427 (+) νs(COO−) Complex binding
1397 (s) 1397 (+) — — 1405 (s) 1406 (+) νs(COO−) Complex non-binding
— — 1400 (s) 1380 (−) 1385 (sh) 1385 (−) νs(COO−) Ligand
1305 (w) 1300 (+) 1297 (w) 1303 (+) 1293 (w) 1310 (+) δ(OvC–O−) or δ(H–C–H) + δ(C–O–H) Complex non-binding
1282 (m) 1276 (+) — 1279 (−) — 1278 (−) ν(C–OH) + δ(C–O–H) Ligand
1254 (w) 1254 (+) 1252 (w) 1252 (+) 1248 (w) 1246 (+) δ(OvC–O−) or δ(H–C–H) + δ(C–O–H) Complex non-binding
1237 (m) 1222 (−) — 1209 (−) — 1207 (−) ν(OC–OH) + δ(C–O–H) Ligand
1078 (m) 1078 (+) 1078 (w) 1074 (+) — — ν(C–OH) Complex
a
s = strong, m = medium, w = weak, sh = shoulder. b + = positive band/maximum, − = negative band/minimum. c ν = stretching vibration, νs/νas =
symmetric/antisymmetric stretching vibration, δ = deformation vibration.

13976 | Dalton Trans., 2012, 41, 13969–13983 This journal is © The Royal Society of Chemistry 2012
complex indicates the presence of coordinated and uncomplexed at 1278(−)/1246(+) cm−1. With respect to the spectra of the pure
carboxylate groups. ligand (Fig. 3), the band at 1278 cm−1 can be attributed to the
In contrast, at alkaline pH 8–12, predominantly the negatively hydroxylate group because it is clearly absent in the spectrum
charged EuCit− complex with the completely deprotonated recorded at pH 13.5, i.e., the Cit4− starts to predominate. There-
citrate anion Cit4− is formed. IR spectra of this complex show fore, from the spectral features observed in this frequency range,
strong absorption bands at 1564 and 1405 cm−1 representing the a contribution of the hydroxylate group to the binding of citrate
ν3,as(COO) and ν3,s(COO) mode, respectively. Because of the to the Eu3+ ion can be concluded.
broad shape of these bands, a more accurate interpretation can be In summary, the IR spectra suggest a predominant mono-
derived from the difference spectra (Fig. 4b). The ν3,as(COO) dentate binding of the Eu3+ ion via carboxylate groups in both
mode is represented by a difference band at 1578(−)/1550(+) cm−1, complex species EuHCit0 and EuCit−. Additionally, the hydro-
whereas the ν3,s(COO) mode is observed at 1427(+)/1385(−) cm−1 xylate group is contributing to the complexation of the Eu3+ ion
showing an interfering small band at 1406 cm−1. The difference in the EuCit− complex.
bands, again, clearly evidence the coordination of the carboxy- To check this assumption and to deduce the actual solution
late groups to the Eu3+ ion. structure of EuHCit0 and EuCit−, different complex structure
The type of coordination of carboxylate groups to heavy metal models were calculated with DFT and the resulting IR absorption
ions can be derived from vibrational spectra by the degree of the spectra were compared to the measured ones. For europium(III),
spectral splitting of the ν3(COO) modes (Δν3).61 The Δν3 value no solution structure for MHCit0 and MCit− has been reported
of the uncomplexed ligand serves as a reference and is until now, but for other metal ions, such as Gd(III),27 Dy(III),27
∼180 cm−1 throughout the pH range (cf. Fig. 4). In general, a bi- Pr(III),28 Am(III),13 U(VI),63 Cr(III),64 Fe(III),64,65 Al(III)66,67 and
dentate coordination of the carboxylate groups to the Eu3+ metal Ga(III),67 several structures have been published in the literature.
ion is expected to show a significantly lower spectral splitting Fig. 5 shows the structure models reported for such MHCit0
than a mono-dentate binding.61,62 For citrate, Δν3 values of (models 1–3) and MCit− (models 4–6) complexes, which vary
164–228 cm−1 were reported for mono-dentate complexes, with regard to the number and the nature of the functional
whereas Δν3 values ranging from 42 to 77 cm−1 were found for groups actually binding to the respective heavy metal ion. Con-
a bi-dentate binding.61,62 sequently, all these structure models were taken into account as
In the spectrum of the EuHCit0 complex, the bands of the car- possible EuHCit0 and EuCit− structures and carefully checked
boxylate groups coordinated to the heavy metal ion are observed with DFT calculations.
at 1600 and 1417 cm−1 resulting in a Δν3 value of ∼185 cm−1. In the case of EuHCit0, model 2 can be excluded, since the
This value is very close to that of the pure ligand and substanti- DFT calculated absorption spectrum of this complex species
ates a mono-dentate coordination of the carboxylate groups to exhibits a strong peak at 1502 cm−1 representing a vibrational
the Eu3+ ion. In the respective spectrum of the EuCit− complex, mode of the co-ordinated hydroxyl group that is not observed in
the maxima observed at 1550 and 1427 cm−1 reflect the differ- the measured IR spectra (see ESI, Fig. S4†). Likewise, model 3
ence bands of the ν3(COO) modes coordinated to the metal ion. can also be ruled out, because no bands were observed in the IR
The resulting Δν3 value of ∼125 cm−1 is smaller than the range absorption spectra that can be assigned to the presence of car-
reported for a mono-dentate complex. However, it is larger than boxyl groups. Hence, out of the published structure models for
those reported for a bi-dentate coordination.61,62 Hence, contri- MHCit0, only model 1 remains. The calculated IR spectrum for
butions of both binding modes have to be considered for the this complex structure is in very good agreement with the
EuCit− complex. measured spectrum at pH 2 and 4 (Fig. 6a). Both νas,s(COO)
In the spectral region between 1300 and 1200 cm−1, all modes and the characteristic bands in the 1300–1200 cm−1
spectra exhibit characteristic features. In the absorption spectra, region occur and are comparable to the measured IR spectrum.
three bands at 1305, 1282, and 1254 cm−1 are observed at pH ≤ This confirms that the Eu3+ ion is complexed according to model
4, which are partially superimposed by the strong band at 1 via one terminal and the central carboxylate group in the
1226 cm−1 in the spectrum recorded at pH 2 (cf. Fig. 4a). At EuHCit0 species. Nevertheless, TRLFS results show that up to
higher pH, only two bands at 1293 and 1248 cm−1 are observed three water molecules are replaced in the first coordination shell
and the corresponding difference spectra show a difference band of the Eu3+ ion by the citrate molecule HCit3−due to sterical

Fig. 5 Schematic structure models of reported MHCit0 (a: models 1–3) and MCit− (b: models 4–6) complexes of several heavy metal ions (M =
UO22+, Gd3+, Dy3+, Cr3+, Al3+, Fe3+, Ga3+, Pr3+ or Am3+).

This journal is © The Royal Society of Chemistry 2012 Dalton Trans., 2012, 41, 13969–13983 | 13977
Fig. 6 Comparison of the measured (dotted line) and the calculated (continuous line) IR spectra of EuHCit0 (a) and EuCit− (b).

Fig. 7 DFT calculated solution structures of the EuHCit0 (a) and EuCit− (b) complexes.

reasons. Another possibility is a change of the total coordination the central carboxylate group was found, whereas the hydroxyl
number from 9 to 8 upon complexation with citrate. group is not coordinated to the heavy metal ion in this species.
In the case of EuCit−, both models 5 and 6 can be excluded In contrast, an additional coordination of the hydroxylate group
since DFT calculations starting with these structures always con- to the Eu3+ ion was demonstrated for the EuCit− species. The
verge to other structure models (see ESI, Fig. S4†). Furthermore, calculated solution structures of both the EuCitH0 and EuCit−
with respect to model 6, complexation of the Eu3+ ion by all complexes are depicted in Fig. 7.
four functional groups is unlikely due to sterical reasons. There-
fore, out of the published structures for MCit−, only model 4
remains. The calculated absorption spectrum of this structure is TRLFS spectra of curium(III) citrate complexes
depicted in Fig. 6b and reveals very good agreement with the
measured spectrum at pH 8–12. Compared to the EuHCit0 In analogy to the europium(III) experiments, curium(III) spectra
species, the signal of the νs(COO) mode is less split in both the were recorded at a constant metal concentration of 3 × 10−7 M
calculated and the measured IR spectrum. Furthermore, in the and pH 2.4, 4.0, 5.5, 8.1, and 12.5 with varying ligand concen-
1300–1200 cm−1 region, the absorption band at 1280 cm−1 is trations of 10−5–10−3 M (Fig. 8).
present neither in the calculated nor in the measured IR The luminescence spectrum of the Cm3+ aqua ion is character-
spectrum. This verifies the postulated deprotonation of the ized by a single emission band at 593.5 ± 0.2 nm representing
hydroxyl group upon complexation and the direct binding of the 6D7/2 → 8S7/2 transition and a luminescence lifetime of 67 ±
this functionality to the Eu3+ ion. Hence, in the EuCit− 1 μs equalling 9 water molecules in the first hydration
species, the Eu3+ ion is complexed according to model 4 via sphere.8,33,39,69 Spectroscopic alterations that indicate the com-
one terminal and the central carboxylate as well as the hydroxy- plexation of the Cm3+ ion were observed at each pH. In general,
late group. the emission band is shifted to longer wavelengths. A splitting
Regarding the binding mode of the carboxylate groups, all of the luminescence band or the appearance of additional peaks
reported structure models have in common that the respective is not observed. Furthermore, the luminescence lifetime is pro-
heavy metal ion is bound in a mono-dentate fashion longed compared to that of the Cm3+ aqua ion. In each sample,
(cf. Fig. 5).13,27,28,63–68 To the best of the authors’ knowledge, a we observed mono-exponential luminescence decay. All spectro-
bi-dentate coordination of citrate has not yet been published. scopic changes are summarized in Table 5.
In conclusion, IR measurements and DFT calculations provide Curium(III) spectra develop in striking analogy to those of
new aspects of the solution structure of the EuHCit0 and EuCit− europium(III). Therefore, only a brief discussion is presented. At
complexes formed at equimolar metal to ligand ratio. For the pH 2.4 (cf. Fig. 8a), all spectra with ligand concentrations ≤7.5
EuHCit0 species, a mono-dentate binding of one terminal and × 10−5 M equal that of the uncomplexed Cm3+ aqua ion. At

13978 | Dalton Trans., 2012, 41, 13969–13983 This journal is © The Royal Society of Chemistry 2012
Fig. 8 Normalized steady-state luminescence spectra of 3 × 10−7 M curium(III) + citrate at pH 2.4 (a), 4.0 (b), 5.5 (c), 8.1 (d), and 12.5 (e), I = 0.1 M
(NaClO4) and room temperature in dependence on the ligand concentration as well as the deconvoluted emission spectra (f) of the single curium(III)
citrate complexes (Cm3+ without citrate = grey line).

Table 5 Luminescence spectroscopic parameters of the determined curium(III) species

pH [citrate]/M λ/nm τ/μs nH2Oa Assigned species Reference


3+
1–6 0 593.5 67 ± 1 8.9 Cm aqua ion Present work
2.4 ≥5 × 10−4 597.1 >80 <7.2 CmHCit0 Present work
4.0 <5 × 10−4 597.3 88 ± 1 6.5 CmHCit0 Present work
5.5 <2.5 × 10−4 597.2 90 ± 6 6.2 CmHCit0 Present work
4.0 >5 × 10−4 600.7 100 ± 5 5.6 Cm(HCitH)HCit2− Present work
5.5 ≥2.5 × 10−4 600.4 112 ± 5 4.9 Cm(HCitH)HCit2− Present work
8.1 >2.5 × 10−5 604.2 >205b <2.3b Cm(HCit)23− Present work
12.5 ≥5 × 10−4 607.5 155 ± 20c 3.3c CmCit25− Present work
≤5.8 — 600.1 122 ± 3 4.4 Cm(III) species in human urine 8
a
Number of water molecules calculated according to Kimura et al.35 with a standard deviation of ±0.5. b Luminescence spectra show that Cm(HCit)23−
and a hydrolyzed curium(III) species co-exist at this pH. Therefore, both species contribute to the measured lifetime. c Parameter reported with
reservations since only two analyzable spectra were measured.

higher citrate concentrations, the emission maximum is shifted to Measurements at pH 4.0 and 5.5 (cf. Fig. 8b and 8c) exhibit a
597.1 ± 0.1 nm and the lifetime is prolonged to >80 μs. This two-stage development with increasing ligand concentration
indicates the formation of a Cm(III) citrate species 1. indicating the formation of two different complex species.

This journal is © The Royal Society of Chemistry 2012 Dalton Trans., 2012, 41, 13969–13983 | 13979
Luminescence spectra at lower citrate concentrations equal those The Cm(III) citrate species 4 can not be clearly identified unequi-
of the Cm(III) citrate species 1 determined at pH 2.4 with an vocally, since only two analyzable spectra were available. Never-
emission maximum at 597.2 ± 0.1 nm and a lifetime of 89 ± theless, with regard to the europium(III) measurements, it is most
6 μs. According to eqn (3), this corresponds to a number of probable that this species is the 1 : 2 complex CmCit25−. As
6.4 ± 0.5 water molecules in the first hydration shell of the Cm3+ shown in the IR and DFT section, at very alkaline pH, the com-
ion. Higher ligand concentrations result in a further shift of the pletely deprotonated citrate anion Cit4− exists and binds to the
emission maximum to 600.5 ± 0.1 nm and a prolongation of the heavy metal ion. The 1 : 2 stoichiometry is supported by the
luminescence lifetime to 106 ± 10 μs corresponding to 5.0 ± 0.5 luminescence lifetime of about 155 ± 20 μs, which corresponds
water molecules. Hence, another Cm(III) citrate species 2 is to 5.6 ± 0.5 water molecules that have been displaced upon com-
formed at both pH values and a large excess of ligand. plexation. However, verification of this hypothesis is still due
At pH 8.1 (cf. Fig. 8d), spectral changes that result from metal and no reference data are reported for curium(III) citrate com-
ion hydrolysis70 occur already in the curium(III) solution without plexes at pH > 10, yet.
citrate. Ligand addition results in a red shift of the luminescence Finally, one pH-dependent series was measured with ligand
peak to 604.2 ± 0.1 nm and a prolonged emission lifetime of excess in the range of pH 2–13 (see ESI, Fig. S3†). All four
>205 μs, which equals <2.3 ± 0.5 water molecules left in the complex species were identified and no indications for the
first hydration shell of the Cm3+ ion. Since all spectroscopic hydrolysis of curium(III) were detected in the whole pH range
parameters do not correspond to one of the complex species under investigation. Strikingly, from pH 7.0 to 10.0, the time-
determined at lower pH, this points to the formation of a new resolved spectra revealed a bi-exponential luminescence decay.
Cm(III) citrate species 3 competing with the metal ion The shorter lifetime equals that of the Cm(HCitH)HCit2−
hydrolysis. species, while the longer lifetime corresponds to the Cm(HCit)23−
In the alkaline region, at pH 12.5 (cf. Fig. 8e), no lumines- complex. Hence, this is direct evidence that both species exist in
cence spectrum was obtained in pure curium(III) solution and a mixture. The pH-dependent measurements show that all
after the addition of <5 × 10−4 M citrate. This results from the species subsequently convert into each other. Moreover, they
formation of the dominant curium(III) species at high pH, which clearly demonstrate that at least two different curium(III) species
is the poorly water soluble Cm(OH)3.33 At ≥7.5 × 10−4 M simultaneously exist at each pH > 2.5.
citrate, luminescence is detectable again but steady-state and In conclusion, the An(III) and Ln(III) complexes predominantly
especially time-resolved spectra are very noisy. However, the reported in the literature at the respective pH values and metal to
emission maximum is even more red-shifted to 607.5 ± 0.1 nm ligand ratios were identified for curium(III) and their spectro-
than at pH 8.1 and the luminescence lifetime is about 155 ± scopic parameters determined (see Table 5). The single com-
20 μs corresponding to 3.3 ± 0.5 water molecules. This indicates ponent spectra of all curium(III) citrate species are depicted in
that Cm3+ ions are partially re-solved via complexation with Fig. 8f. To the best of our knowledge, no luminescence spectro-
citrate and a new Cm(III) citrate species 4 is formed. scopic reference data on the complexation of curium(III) with
In summary, like for europium(III), four different curium(III) citrate have been published until now.
citrate species were characterized by varying the citrate concen-
tration at pH 2–12.5. Identification of the complex species was
carried out in the same way as for europium(III) in consideration Complex stability constants and speciation of curium(III) and
of the FT-IR and DFT results and on the basis of the measured europium(III) citrate complexes
luminescence lifetimes. Literature data on curium(III) citrate
complexes9,15,19–22,24,25 were also taken into account and Based on the identified species, the formation constants for the
support our assignments. direct reactions, log Kpqr ( p = number of M3+ ions, q = number
All curium(III) complexes are in strikingly analogy to those of of ligand molecules, r = number of protons), and for the overall
europium(III). Thus, the Cm(III) citrate species 1 likely represents reactions, log βpqr, were calculated according to eqn (5)–(8) and
the charge neutral 1 : 1 complex CmHCit0 with the HCit3− (9)–(11), respectively. The resulting complex stability constants
anion. The emission lifetime of 89 ± 6 μs equals 2.5 ± 0.5 water log K, log K0 and log β are listed in Table 6. Calculation of the
molecules that have been displaced by the ligand upon com- values for CmCit25− was not successful, since only two analyz-
plexation. This fits to the IR spectroscopic results where the able spectra could be measured. The Gibbs energies at zero ionic
citrate anion HCit3− binds to the Cm3+ ion via two carboxylate strength are also summarized in Table 6.
groups in a mono-dentate way. The Cm(III) citrate species 2 turns A comparison of the log K values for curium(III), americium(III),
out to be the 1 : 2 complex Cm(HCitH)HCit2−, in which one and europium(III) citrate species determined in this study with
citrate HCit3− and one hydrogen citrate anion HCitH2− bind to those reported in the literature9,13–15,17–21 is given in Fig. 9. Our
the Cm3+ ion. The luminescence lifetime of 106 ± 10 μs indi- complex stability constants are in very good agreement
cates that 5.1 ± 0.5 water molecules in the first coordination shell especially with those reported by Stepanov.19 Discrepancies with
of the Cm3+ ion have been displaced by the ligands. Assuming a other literature data can be attributed to the different ionic
mono-dentate complexation as in the case of the CmHCit0 strengths (e.g. in the case of Mathur et al.9) and the complex
complex, this indicates the replacement of about 2.5 water mole- species that were considered to form. In particular, the existence
cules by each ligand molecule. The Cm(III) citrate species 3 is of the M(HCitH)HCit2− species is still controversial in the litera-
identified as the 1 : 2 complex Cm(HCit)23−. The luminescence ture. Therefore, some authors account for the formation of this
lifetime of >205 μs corresponds to a replacement of at least complex and some do not. This might contribute to the deviating
6.6 ± 0.5 water molecules and supports the 1 : 2 stoichiometry. stability constants reported in the literature (see ESI, Table S1†).

13980 | Dalton Trans., 2012, 41, 13969–13983 This journal is © The Royal Society of Chemistry 2012
Table 6 Thermodynamic parameters of the curium(III) and europium(III)
citrate complexes

Species log βa log Ka log K0 b ΔRG0/kJ mol−1 c

CmHCit0 21.0 ± 0.2 7.4 ± 0.2 9.3 ± 0.2 −23.1 ± 0.2


Cm(HCitH)HCit2− 43.8 ± 0.3 11.0 ± 0.3 12.9 ± 0.3 −32.0 ± 0.3
Cm(HCit)23− 38.4 ± 0.7 11.3 ± 0.7 13.2 ± 0.7 −32.7 ± 0.7
CmCit25− — — — —
EuHCit0 21.2 ± 0.2 7.5 ± 0.2 9.4 ± 0.2 −23.3 ± 0.2
Eu(HCitH)HCit2− 43.6 ± 0.5 10.8 ± 0.5 12.7 ± 0.5 −31.6 ± 0.5
Eu(HCit)23− 38.5 ± 0.4 11.4 ± 0.4 13.2 ± 0.4 −33.0 ± 0.4
EuCit25− 21.0 ± 0.2 21.0 ± 0.2 20.9 ± 0.2 −51.8 ± 0.2
a
Complex formation constants at I = 0.1 M (NaClO4). b Complex
formation constants at zero ionic strength. c Gibbs energy at zero ionic
strength.

Fig. 10 Speciation diagram of 3 × 10−5 M europium(III) + 10−3 M


citrate in aqueous solution at I = 0.1 M (NaClO4) and room temperature
in dependence on the pH.

EuCit25−. Therefore, also the speciation diagram of curium(III) in


aqueous citrate solutions should be very similar to the presented
one of europium(III). Fig. 10 shows that all complex species
determined within this study subsequently form and convert into
each other. Furthermore, it is obvious that, except for very acidic
and alkaline pH, always a mixture of three species exists. This,
in turn, reflects the difficulty in determining the spectroscopic
properties of a single complex species and accounts for the dis-
crepancies within the reported complex stability constants
(cf. Fig. 9).

Curium(III) and europium(III) citrate speciation in biological


systems

At physiological pH 7.4, the M(HCit)23− species of curium(III)


and europium(III) dominates the speciation to >80% (cf. Fig. 10).
Therefore, it is reasonable to assume that this complex might
also form to a certain extent in blood and other biofluids at near
neutral pH.
Fig. 9 Comparison of log K values for curium(III), americium(III), and In the case of human urine, a previous study on the in vitro
europium(III) citrate species (filled symbol = present work, open symbol = binding form of curium(III) and europium(III) in natural samples
literature data9,13–15,17–22,24). revealed that, at slightly acidic pH ≤ 5.7, both heavy metal ions
were, in fact, predominantly bound by citrate.8 Comparison of
the reported urinary citrate complexes with the single complex
Within our study, we were not able to give final evidence for the species determined within the present work is given in Tables 1
existence of the M(HCitH)HCit2− species, although this and 5.
complex was found to be the best species that fits to our experi- In the case of curium(III), it is obvious that the species formed
mental data by now. The question, therefore, remains still open in human urine is the Cm(HCitH)HCit2− complex. Both the
and needs further investigation. Furthermore, a conclusion luminescence wavelength and the emission lifetime are identical.
whether curium(III) or europium(III) form stronger complexes In contrast, for europium(III), the emission maxima and the fine
with citrate is not possible, since the stability constants of an structure of the urinary spectrum fit those of the Eu(HCitH)
analog species are quite identical within the range of error. HCit2− complex, while the luminescence lifetime equals that of
Fig. 10 depicts the representative speciation diagram of the EuHCit0 species. Therefore, a mixture of these two com-
europium(III) in aqueous citrate solutions with ligand excess in plexes exists in human urine in vitro at slightly acidic pH. Which
dependence on the pH, which has been calculated using the of the two species actually dominates most probably depends on
complex stability constants determined within the present work. the urinary citrate concentration. Thermodynamic modeling
Since curium(III) revealed striking analogy to europium(III) shows that at lower citrate concentrations, the 1 : 1 complex
throughout all experiments, it is reasonable to assume that the EuHCit0 is formed to a higher extent, whereas at higher citrate
log K120 value for CmCit25− should be comparable to those for concentrations the 1 : 2 complex Eu(HCitH)HCit2− dominates.

This journal is © The Royal Society of Chemistry 2012 Dalton Trans., 2012, 41, 13969–13983 | 13981
This is the first time that the molecular binding form of a triva- gastrointestinal tract) are in progress and shall emphasize the
lent actinide and/or lanthanide in a biological fluid is directly recent results and assumptions.
determined via experiment and demonstrates the great suitability
of TRLFS and IR spectroscopy for such in situ speciation inves-
tigations. Furthermore, knowledge of the actual An(III) and Acknowledgements
Ln(III) citrate complexes occurring in body fluids might be a pre-
This work was funded by the Deutsche Forschungsgemeinschaft
requisite to develop new and specific potential decontamination
under Contract No. BE 2234/10-1/2. We are indebted to the U.S.
agents based on ligands naturally produced in the human body.
Department of Energy, Office of Basic Energy Sciences, for the
use of 248Cm via the transplutonium element production facilities
at Oak Ridge National Laboratory, which was made available as
Conclusion part of a collaboration between the Helmholtz-Zentrum Dresden-
Rossendorf and the Lawrence Berkeley National Laboratory. We
For the first time, the complexation of curium(III) and europium(III)
acknowledge the Zentrum für Informationsdienste und Hochleis-
with citrate was investigated over a wide pH and ligand con-
tungsrechnen, Technische Universität Dresden, Germany, for use
centration range using the two complementary spectroscopic
of their supercomputers.
techniques TRLFS and IR as well as DFT calculations. In the
case of both elements, four complex species were identified with
TRLFS, out of which, for europium(III), two of them have been
Notes and references
characterized in more detail with ATR-FT-IR and DFT.
The spectral emission features of four distinct curium(III) and 1 J.-C. G. Bünzli, Nat. Chem., 2010, 2, 696; S. V. Eliseeva and
J.-C. G. Bünzli, Chem. Soc. Rev., 2010, 39, 189–227; H. Elsner, Com-
europium(III) complex species were assigned to MHCit0,
modity Top News, 2011, 36, 1–8; B. Karn, J. Environ. Monit., 2011, 13,
M(HCitH)HCit2−, M(HCit)25− and MCit25−. We demonstrated 1184–1189; M. Liedtke and H. Elsner, Commodity Top News, 2010, 31,
that, on the one hand, all species can be discriminated and ident- 1–6.
ified with regard to these features but, on the other hand, that an 2 M. F. Bellin and A. J. Van Der Molen, Eur. J. Radiol., 2008, 66,
160–167; A. Thibon and V. C. Pierre, Anal. Bioanal. Chem., 2009, 394,
exact characterization is hindered by the simultaneous formation 107–120; G. Muller, Dalton Trans., 2009, 9692–9707.
of several complexes in solution. Furthermore, the lifetimes 3 E. Ansoborlo, O. Prat, P. Moisy, C. Den Auwer, P. Guilbaud, M. Carriere,
measured with TRLFS give information about the number of B. Gouget, J. Duffield, D. Doizi, T. Vercouter, C. Moulin and V. Moulin,
water molecules in the first coordination shell of the Cm3+ and Biochimie, 2006, 88, 1605–1618.
4 P. W. Durbin, in Uranium, Plutonium, Transplutonic Elements, ed.
Eu3+ ion, respectively, which thereby give how many waters H. C. Hodge, J. N. Stannard and J. B. Hursh, Springer Verlag, Berlin,
have been displaced by ligand molecules upon complexation. 1973, pp. 739–896; D. M. Taylor and R. W. Leggett, Radiat. Prot.
This allows first estimations of the structure of the respective Dosim., 2003, 105, 193–198; F. Ménétrier, D. M. Taylor and A. Comte,
Appl. Radiat. Isot., 2008, 66, 632–647.
complex species but detailed structural information, such as the 5 J. R. Duffield, D. M. Taylor and D. R. Williams, in Handbook on the
functional groups of the ligand participating in the curium(III) Physics and Chemistry of Rare Earths, Vol. 18: Lanthanides/Actinides:
and europium(III) complexation, respectively, as well as the Chemistry, ed. K. A. Gschneidner Jr., L. Eyring, G. R. Choppin and
binding mode, cannot be obtained with TRLFS. G. H. Lander, Elsevier Science B. V., Amsterdam, 1994, pp. 591–621.
6 G. A. Turner and D. M. Taylor, Phys. Med. Biol., 1968, 13, 535–546;
IR spectroscopy combined with DFT calculation, in turn, J. R. Cooper and H. S. Gowing, Int. J. Radiat. Biol., 1981, 40, 569–572;
provide information on the mode of coordination of functional D. M. Taylor, J. Alloys Compd., 1998, 271, 6–10; D. M. Taylor,
groups to the metal ion. Thus, the two techniques are comp- J. R. Duffield, D. R. Williams, L. Yule, P. W. Gaskin and P. Unalkat,
Eur. J. Solid State Inorg. Chem., 1991, 28, 271–274.
lementary to TRLFS. We demonstrated that in the EuHCit0
7 G. N. Stradling, D. S. Popplewell and G. J. Ham, Health Phys., 1976, 31,
species, formed at acidic pH, only carboxylate groups bind the 517–519; D. S. Popplewell, G. N. Stradling and G. J. Ham, Radiat. Res.,
Eu3+ ion, whereas in the EuCit− complex, formed at alkaline pH, 1975, 62, 513–519.
additionally also the hydroxylate group is involved. Furthermore, 8 A. Heller, A. Barkleit and G. Bernhard, Chem. Res. Toxicol., 2011, 24,
193–203.
comparing IR with TRLFS spectra we showed that, due to steri- 9 J. N. Mathur, K. Cernochova and G. R. Choppin, Inorg. Chim. Acta,
cal reasons, each functional group binding to the Eu3+ ion 2007, 360, 1785–1791.
replaces more than one water molecule from its first hydration 10 W. M. Boothby and M. Adams, Am. J. Physiol., 1934, 107, 471–479;
shell. Moreover, DFT calculations aided us to interpret the A. C. Kuyper and H. A. Mattill, J. Biol. Chem., 1933, 103, 51–60;
C. C. Sherman, L. B. Mendel, A. H. Smith and M. C. Toothill, J. Biol.
measured IR spectra correctly and provided a basis to decide Chem., 1936, 113, 247–263.
about the actual complex solution structure. Therefore, the 11 A. M. N. Silva, X. L. Kong and R. C. Hider, Biometals, 2009, 22,
results obtained with each of the three techniques are in good 771–778.
12 W. Hummel, G. Anderegg, L. Rao, I. Puigdomènech and O. Tochiyama,
agreement with and verify each other. Chemical Thermodynamics, Vol. 9: Chemical Thermodynamics of Com-
In summary, the present work highlights the importance of pounds and Complexes of U, Np, Pu, Am, Tc, Se, Ni and Zr with Selected
citrate in the complexation of An(III) and Ln(III) in general and Organic Ligands, Elsevier B. V., Amsterdam, 2005.
curium(III) and europium(III) in particular. Furthermore it indi- 13 S. H. Eberle and F. Moattar, Inorg. Nucl. Chem. Lett., 1972, 8, 265–270.
14 E. Ohyoshi and A. Ohyoshi, J. Inorg. Nucl. Chem., 1971, 33,
cates the crucial role of citrate for their speciation in human 4265–4273.
body fluids, like urine and blood, since the dominant species at 15 R. Guillaumont and L. Bourderie, Bull. Soc. Chim. Fr., 1971, 8,
physiological pH is the strongly negatively charged M(HCit)23−. 2806–2809.
Further measurements of the curium(III) and europium(III) specia- 16 P. G. Daniele, A. Derobertis, C. Rigano and S. Sammartano, Ann. Chim.,
1985, 75, 115–120.
tion in natural biofluids (e.g. blood and saliva) as well as in 17 S. Bouhlassa and R. Guillaumont, J. Less Common Met., 1984, 99,
model solutions (e.g. for different fluids of the human 157–171.

13982 | Dalton Trans., 2012, 41, 13969–13983 This journal is © The Royal Society of Chemistry 2012
18 S. Hubert, M. Hussonno, L. Brillard, G. Goby and R. Guillaumont, 46 T. Vercouter, B. Amekraz, C. Moulin, E. Giffaut and P. Vitorge, Inorg.
J. Inorg. Nucl. Chem., 1974, 36, 2361–2366. Chem., 2005, 44, 7570–7581.
19 A. V. Stepanov, Zh. Neorg. Khim., 1971, 16, 2981–2985. 47 C. De Stefano, C. Foti, O. Giuffre and S. Sammartano, J. Chem. Eng.
20 H. Itoh, M. Fujisawa, Y. Ikegami and Y. Suzuki, Lanthanide Actinide Data, 2001, 46, 1417–1424.
Res., 1985, 1, 79–88. 48 R. J. Silva, G. Bidoglio, M. H. Rand, P. B. Robouch, H. Wanner and
21 A. V. Stepanov, Radiokhimiya, 1959, 1, 668–673. I. Puigdomènech, Chemical Thermodynamics, Vol. 2. Chemical Thermo-
22 S. Hubert, M. Hussonnois and R. Guillaumont, J. Inorg. Nucl. Chem., dynamics of Americium, Elsevier Science B. V., Amsterdam, 1995.
1973, 35. 49 P. Gans, A. Sabatini and A. Vacca, Protonic Software, Leeds, 2006.
23 S. Bouhlassa, M. Petitramel and R. Guillaumont, Bull. Soc. Chim. Fr., 50 M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A. Robb,
1984, 5–11. J. R. Cheeseman, J. A. Montgomery, Jr., T. Vreven, K. N. Kudin,
24 P. Thakur, Y. Xiong, M. Borkowski and G. R. Choppin, Radiochim. Acta, J. C. Burant, J. M. Millam, S. S. Iyengar, J. Tomasi, V. Barone,
2012, 100, 1–8. B. Mennucci, M. Cossi, G. Scalmani, N. Rega, G. A. Petersson,
25 B. S. Jensen and H. Jensen, Radiochim. Acta, 1988, 44/45, 45–49. H. Nakatsuji, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa,
26 A. Moutte and R. Guillaumont, Rev. Chim. Miner., 1969, 6, 603–610. M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai, M. Klene, X. Li,
27 A. P. G. Kieboom, J. M. Vandertoorn, J. A. Peters, W. Bovee, J. E. Knox, H. P. Hratchian, J. B. Cross, V. Bakken, C. Adamo,
A. Sinnema, C. A. M. Vijverberg and H. Vanbekkum, Recl. Trav. Chim. J. Jaramillo, R. Gomperts, R. E. Stratmann, O. Yazyev, A. J. Austin,
Pays-Bas, 1978, 97, 247–248; A. P. G. Kieboom, C. A. M. Vijverberg, R. Cammi, C. Pomelli, J. W. Ochterski, P. Y. Ayala, K. Morokuma,
J. A. Peters and H. Vanbekkum, Recl. Trav. Chim. Pays-Bas, 1977, 96, G. A. Voth, P. Salvador, J. J. Dannenberg, V. G. Zakrzewski, S. Dapprich,
315–316; C. A. M. Vijverberg, J. A. Peters, A. P. G. Kieboom and A. D. Daniels, M. C. Strain, O. Farkas, D. K. Malick, A. D. Rabuck,
H. Vanbekkum, Tetrahedron, 1986, 42, 167–174. K. Raghavachari, J. B. Foresman, J. V. Ortiz, Q. Cui, A. G. Baboul,
28 C. B. Yuan, J. Z. Liu, D. Q. Zhao, Y. J. Wu and J. Z. Ni, Polyhedron, S. Clifford, J. Cioslowski, B. B. Stefanov, G. Liu, A. Liashenko,
1995, 14, 3579–3583. P. Piskorz, I. Komaromi, R. L. Martin, D. J. Fox, T. Keith, M. A. Al-
29 G. E. Jackson and J. Dutoit, J. Chem. Soc., Dalton Trans., 1991, Laham, C. Y. Peng, A. Nanayakkara, M. Challacombe, P. M. W. Gill,
1463–1466. B. Johnson, W. Chen, M. W. Wong, C. Gonzalez and J. A. Pople,
30 E. Ansoborlo, L. Bion, D. Doizi, C. Moulin, V. Lourenco, C. Madic, Gaussian 03, Revision E.01, Gaussian, Inc., Wallingford CT, 2004..
G. Cote, J. Van der Lee and V. Moulin, Radiat. Prot. Dosim., 2007, 1–6. 51 V. Barone and M. Cossi, J. Phys. Chem. A, 1998, 102, 1995–2001.
31 A. Barkleit, G. Geipel, M. Acker, S. Taut and G. Bernhard, Spectrochim. 52 A. Bondi, J. Phys. Chem., 1964, 68, 441–451.
Acta A Mol. Biomol. Spectros., 2011, 78, 549–552. 53 M. Dolg, H. Stoll, A. Savin and H. Preuss, Theor. Chim. Acta, 1989, 75,
32 Lanthanide Probes in Life, Chemical and Earth Sciences: Theory and 173–194.
Practice, ed. J.-C. G. Bünzli and G. R. Choppin, Elsevier Science B. V., 54 R. Krishnan, J. S. Binkley, R. Seeger and J. A. Pople, J. Chem. Phys.,
Amsterdam, 1989. 1980, 72, 650–654.
33 H. Moll, V. Brendler and G. Bernhard, Radiochim. Acta, 2011, 99, 55 A. Barkleit, S. Tsushima, O. Savchuk, J. Philipp, K. Heim, M. Acker,
775–782. S. Taut and K. Fahmy, Inorg. Chem., 2011, 50, 5451–5459.
34 R. N. Collins, T. Saito, N. Aoyagi, T. E. Payne, T. Kimura and 56 C. Moulin, J. Wei, P. Van Iseghem, I. Laszak, G. Plancque and
T. D. Waite, J. Environ. Qual., 2011, 40, 731–741; T. E. Keyes, V. Moulin, Anal. Chim. Acta, 1999, 396, 253–261.
R. J. Forster and C. Blackledge, Spectrosc. Prop. Inorg. Organomet. 57 G. Plancque, V. Moulin, P. Toulhoat and C. Moulin, Anal. Chim. Acta,
Compd., 2010, 41, 211–261. 2003, 478, 11–22.
35 T. Kimura and G. R. Choppin, J. Alloys Compd., 1994, 213, 313–317; 58 C. M. Andolina, R. A. Mathews and J. R. Morrow, Helv. Chim. Acta,
T. Kimura, G. R. Choppin, Y. Kato and Z. Yoshida, Radiochim. Acta, 2009, 92, 2330–2348.
1996, 72, 61–64; T. Kimura and Y. Kato, J. Alloys Compd., 1998, 277, 59 J.-C. G. Bünzli, in Metal Ions in Biological Systems, Vol. 42: Metal Com-
806–810. plexes in Tumor Diagnosis and as Anticancer Agents, ed. A. Sigel and
36 H. Moll, A. Johnsson, M. Schäfer, K. Pedersen, H. Budzikiewicz and H. Sigel, Marcel Dekker Inc., New York, 2004, pp. 39–75.
G. Bernhard, Biometals, 2008, 21, 219–228. 60 M. Lindegren, J. S. Loring and P. Persson, Langmuir, 2009, 25,
37 K. Müller, H. Foerstendorf, S. Tsushima, V. Brendler and G. Bernhard, 10639–10647; S. P. Pasilis and J. E. Pemberton, Geochim. Cosmochim.
J. Phys. Chem. A, 2009, 113, 6626–6632; B. Li, J. Raff, A. Barkleit, Acta, 2008, 72, 277–287; K. Lackovic, B. B. Johnson, M. J. Angove and
G. Bernhard and H. Foerstendorf, J. Inorg. Biochem., 2008, 104, 718–725. J. D. Wells, J. Colloid Interface Sci., 2003, 267, 49–59; S. P. Pasilis and
38 Origin Lab Corporation, Northhampton, 2006. J. E. Pemberton, Inorg. Chem., 2003, 42, 6793–6800.
39 J. I. Kim, R. Klenze and H. Wimmer, Eur. J. Solid State Inorg. Chem., 61 G. B. Deacon and R. J. Phillips, Coord. Chem. Rev., 1980, 33, 227–250;
1991, 28, 347–356. M. Kakihana, T. Nagumo, M. Okamoto and H. Kakihana, J. Phys.
40 J. V. Beitz, in Handbook on the Physics and Chemistry of Rare Earths, Chem., 1987, 91, 6128–6136.
Vol. 18: Lanthanides/Actinides: Chemistry, ed. K. A. Gschneidner Jr, 62 G. Montavon, C. Hennig, P. Janvier and B. Grambow, J. Colloid Interface
L. Eyring, G. R. Choppin and G. H. Lander, Elsevier Science B. V., Sci., 2006, 300, 482–490.
Amsterdam, 1994, pp. 159–196; W. D. Horrocks Jr. and M. Albin, in 63 K. S. Rajan and A. E. Martell, Inorg. Chem., 1965, 4, 462–469.
Progress in Inorganic Chemistry, ed. S. J. Lippard, John Wiley Hoboken, 64 Y. Z. Hamada, B. L. Carlson and J. T. Shank, Synth. React. Inorg.
1984, vol. 31, pp. 1–104. Met.-Org. Chem., 2003, 33, 1425–1440.
41 R. A. Binstead, A. D. Zuberbühler and B. H. Jung, Spectrum Software 65 A. M. N. Silva, X. Kong, M. C. Parkin, R. Cammack and R. C. Hider,
Associates, 2004. Dalton Trans., 2009, 8616–8625.
42 P. Gans, A. Sabatini and A. Vacca, Protonic Software, Leeds, 2008. 66 W. H. Kuan, M. K. Wang, P. M. Huang, C. W. Wu, C. M. Chang and
43 A. Heller, A. Barkleit, G. Bernhard and J.-U. Ackermann, Inorg. Chim. S. L. Wang, Water Res., 2005, 39, 3457–3466; R. J. Motekaitis and
Acta, 2009, 362, 1215–1222; A. Heller, O. Rönitz, A. Barkleit, A. E. Martell, Inorg. Chem., 1984, 23, 18–23.
G. Bernhard and J.-U. Ackermann, Appl. Spectrosc., 2010, 64, 930–935; 67 M. Clausen, L. O. Ohman and P. Persson, J. Inorg. Biochem., 2005, 99,
M. Glorius, H. Moll and G. Bernhard, Polyhedron, 2008, 27, 2113–2118. 716–726.
44 W. Hummel, U. Berner, E. Curti, F. J. Pearson and T. Thoenen, Nagra/PSI 68 T. Toraishi, S. Nagasaki and S. Tanaka, J. Mol. Struct. (THEOCHEM),
Chemical Thermodynamic Data Base 01/01, Nagra, Wettingen, 2002; 2005, 757, 87–97.
R. Guillaumont, T. Fanghänel, J. Fuger, I. Grenthe, V. Neck, D. A. Palmer 69 N. M. Edelstein, R. Klenze, T. Fanghänel and S. Hubert, Coord. Chem.
and M. H. Rand, Chemical Thermodynamics, Vol. 5: Update on the Rev., 2006, 250, 948–973.
Chemical Thermodynamics of Uranium, Neptunium, Plutonium, 70 H. Wimmer, R. Klenze and J. I. Kim, Radiochim. Acta, 1992, 56, 79–83;
Americium and Technetium, Elsevier Science B. V., Amsterdam, 2003. T. Fanghänel, J. I. Kim, P. Paviet, R. Klenze and W. Hauser, Radiochim.
45 L. D. Pettit, I. Puigdomenech, H. Wanner, I. Sukhno and V. Buzko, Acta, 1994, 66–7, 81–87; T. Rabung, M. Altmaier, V. Neck and
Academic Software http://www.acadsoft.co.uk, 2004. T. Fanghänel, Radiochim. Acta, 2008, 96, 551–559.

This journal is © The Royal Society of Chemistry 2012 Dalton Trans., 2012, 41, 13969–13983 | 13983
View publication stats

You might also like