You are on page 1of 8

www.elsevier.

nl/locate/ica

Inorganica Chimica Acta 293 (1999) 53 – 60

Thermodynamic and structural study of inclusion complexes


between trivalent lanthanide ions and native cyclodextrins
Nicolas Fatin-Rouge, Jean-Claude G. Bünzli *
Uni6ersité de Lausanne, Institut de Chimie Minérale et Analytique, Unité de Recherche sur les Eléments f, BCH 1402,
CH-1015 Lausanne, Switzerland

Received 12 February 1999; accepted 4 May 1999

Abstract

The interaction of trivalent lanthanide ions (Ln3 + ) with native cyclodextrins (CDs) is investigated in acidic and basic aqueous
media. At low pH, the association constants for 1:1 complexes are in the range log K =2–4 (m= 0.1 M, NaCl or TMACl), as
determined by pH-potentiometric study of the hydrolysis of the Ln ions in absence and in presence of CDs. The thermodynamic
parameters of the inclusion reaction show that the complexation of Ln3 + ions inside the CD cavity is entropically driven and does
not depend upon the host. 1H and 13C NMR spectra point to inclusion of La3 + cations in the cavity of a-CD occurring in the
narrowest part of the host molecule, close to the C5 carbon atoms. a-CD acts like a crown ether, its anomeric oxygen atoms being
the donor atoms. Complexation of Tb3 + by partially deprotonated a-CD was investigated at pH 12.33: 13C NMR measurements
show that complexation occurs at the OH-3 secondary hydroxyl groups of a-CD while UV – Vis spectrophotometric titration leads
to an apparent constant log Kapp :4.2 (1:1 complex). Lifetime measurements of the Eu(5D0) and Tb(5D4) levels confirm these
findings. © 1999 Elsevier Science S.A. All rights reserved.

Keywords: Cyclodextrin complexes; Thermodynamics; Lanthanide complexes

1. Introduction complexation remains a challenge for synthetic


chemists.
The search for efficient luminescent sensors and CDs have moderate solubility in water and are
switches for the analysis of biomaterials is presently known to include organic molecules or fragments in
stimulating a wealth of interest [1,2]. In this context, their cavity. This property can be used to generate close
both trivalent lanthanide ions (Ln3 + ) and cyclodextrins contacts between these fragments and complexed
(CDs) hold a distinctive place, because the first have cations, as demonstrated with modified CDs [7,8]. In
luminescent properties that are environment sensitive this respect, the recent elaboration of rules for selective
[3] and the second provide specific and protective mi- substitutions at the hydroxyl positions of CDs [9] led to
cro-environments for small molecules included in their the design of efficient supramolecular devices and archi-
cavity [4,5]. Lanthanide-centered luminescence is easy
tectures [5,10,11]. CDs can also include coordination
to sensitize by the antenna effect [6], but avoiding
complexes [12] with sizable binding constants as large
de-activation through high energy vibrations of bound
as 2000 M − 1 [13]. They are expected to reduce deacti-
ligands or of solvent molecules is more difficult [3].
vation of luminescent excited states of LnIII ions and to
Indeed, small water molecules interact strongly with
Ln3 + ions and the design of a coordination environ- enhance the luminescence intensity of chromophores
ment bearing chromophores suitable for an efficient [14,15] and/or the yield of the energy transfer onto the
antenna effect and protecting the LnIII ions from water Ln3 + ion [8].
For these reasons, we have launched a research pro-
gram aiming at investigating second sphere Ln3 + com-
* Corresponding author. Tel.: + 41-21-692 3821/3850; fax: +41-
21-692 3825. plexation by CDs and its effect on the photophysical
E-mail address: jean-claude.bunzli@icma.unil.ch (J.-C.G. Bünzli) properties of the lanthanide ions. In absence of funda-

0020-1693/99/$ - see front matter © 1999 Elsevier Science S.A. All rights reserved.
PII: S 0 0 2 0 - 1 6 9 3 ( 9 9 ) 0 0 2 2 7 - 3
54 N. Fatin-Rouge, J.-C.G. Bünzli / Inorganica Chimica Acta 293 (1999) 53–60

mental data on Ln3 + – CDs interaction, but for one 6.3013.210, 10 ml, accuracy 0.03 ml), was used along
partial study on the dissolution of praseodymium hy- with a Metrohm 6.0238.000 glass electrode. The latter
droxide by a- and g-CDs in basic medium [16], we have was prepared with the same electrolyte and at the same
started our study by looking at inclusion complexes of ionic strength (0.1 M) than the solutions to study in
Ln3 + aqua-ions with native CDs. In this paper, we order to minimize junction potentials; traces of silver
report thermodynamic and structural data on the inser- nitrate were added and after measurement the electrode
tion of La3 + (aq) in a-, b-, and g-CD and of Ln3 + (aq) was left in an acidic solution (HCl, pH 2.5); the elec-
in a-CD in acidic solutions, as well as of the association trode was used 48 h after conditioning. The following
of Tb3 + (aq) with a-CD at basic pHs. Metrohm buffer solutions were used to calibrate the
electrode: 62305.010 (pH 4.00), 62305.020 (pH 7.00),
62305.020 (pH 9.00), and 0.025 M Na2CO3-NaHCO3
2. Experimental (m=0.1 M, NaCl or TMACl; pH 10.01 [19]). The data
(60 points per curve, drift 1 mV mn) were treated
2.1. Sol6ents and stating materials mathematically by the program SUPERQUAD [20]
using a Marquardt algorithm. The following values of
All the chemicals used in this study were of analytical the ionic product of water were taken into account
grade and were used without further purification: a-, b- (m= 0.1 M): pKw = 13.72 (NaCl [21]) and pKw =13.93
and g-CDs, tetramethylammonium chloride (TMACl), (TMACl), as determined from HCl–TMAOH titration.
tetramethylammonium hydroxide (TMAOH) (Fluka),
NaCl, Na2B4O7, Na2CO3, NaHCO3 (Merck). Stock so- 2.3. Spectrophotometric titrations
lutions of Ln(ClO4)3 ( 8 mmol l − 1) were prepared
from lanthanides oxides (99.99%, Rhône Poulenc). Electronic spectra were recorded with a Perkin–
Ln3 + solutions were prepared just before use from Elmer l-900 spectrophotometer at 25.0°C (205–500
freshly boiled doubly distilled water saturated with N2. nm, 100 nm min − 1 scan-speed) using 0.1 cm quartz
They were acidified systematically with HCl to pH 4 cells. Titration of a 0.009 M a-CD solution (20 ml) by
before complexometric titration, to avoid hydroxide Tb(ClO4)3·nH2O (0.467 M, pH 4) was performed at pH
precipitation; standardized Na2H2EDTA solution in 12.3090.05 in a thermostated (25.09 0.1°C) glass
urotropine buffered medium was used along with jacketed vessel and under Ar atmosphere. The pH was
xylene orange as indicator. Stock solutions of CDs maintained constant during the titration by addition of
( 8 mmol l − 1) were prepared from freshly boiled KOH 0.5 M. The Ln3 + solution was added with fast
doubly distilled water saturated with N2. The concen- mixing to avoid local excess and precipitation. Under
trations of CDs were calculated from the optical rota- these conditions, equilibrium was attained within 10
tion of sodium light at 25.0°C (a-CD =150.5 9 0.5°, min. The [Tb3 + ]/[a-CD] ratio ranged from 0 to 1.3. A
b-CD =162.590.5° and g-CD = 177.5 9 0.5° [17]). 15 min time delay was imposed after each 20.0 ml
Substances used for 1H NMR were dried under vacuum addition, before 0.3 ml of solution was transferred into
after dissolution in D2O. a quartz cell with a Teflon® syringe. The apparent
complexation constant was determined using the SPEC-
2.2. pH Potentiometry FIT program [22].

The hydrolysis constants of Ln3 + and the association 2.4. NMR measurements
constants with non deprotonated CDs were determined
1
by pH potentiometric titrations of 0.002 M metal solu- H and 13C spectra were collected on DPX-400 or
tions in absence and in presence of 1 equiv. of CD [18]. AM-360 Bruker spectrometers. 1H NMR spectra were
Care was exercised to analyze the pH domain in which recorded in D2O and 13C spectra were recorded in
CDs are non deprotonated. The titrations were carried H2O/D2O 50/50 v/v using methanol as external refer-
out using 20 ml sample solutions in a thermostated ence. Assignment of the 13C spectra was done using 1H
(25.0 9 0.1°C) glass-jacketed vessel under Ar atmo- decoupling, DEPT (distortionless enhancement by po-
sphere. The ionic strength of the solutions was held larization transfer) 135 [23] and previously published
constant (m =0.1 M) by adding an aliquot of NaCl or results [24]. 1H NMR spectrum of a-CD was assigned
TMACl. Reactants were mixed in the thermostated following Woods et al. results [25]. Chemical shifts of
vessel and acidified to pH 4.0090.02 (HCl) at least 30 a-CDs H2-6 protons were obtained using selective irra-
min before titration. Titrations were carried out with an diation. The a-CD titration with La3 + (aq) in acidic
automatic Metrohm Titrino 736 GP potentiometer (res- solution was monitored by 13C NMR as follows: five
olution 0.1 mV, accuracy 0.2 mV) using a constant samples of a-CD  0.05 M (0.107 g, 2 ml) were pre-
volume addition (0.1 ml) program and linked to an pared and La(ClO4)3·nH2O was added to yield [La3 + ]t /
IBM PS/2 computer. An automatic burette (Metrohm [a-CD]t ratios ranging between 0 and 2. Both the LaIII
N. Fatin-Rouge, J.-C.G. Bünzli / Inorganica Chimica Acta 293 (1999) 53–60 55

salt and the a-CD were titrated just before measure- means that CDs do not exchange protons in the pH
ment to determine their exact hydration number. The range analyzed. Therefore, a special model was adopted
pH was adjusted to 4.5 using 1 to 7 ml of H2SO4(aq) for La3 + , taking into account partially deprotonated
and a glass electrode. a-CDs. Below pH 10, speciation diagrams show that in
For titration in basic medium, Tb(ClO4)3·nH2O 0.467 the worst case (i.e. g-CD) monodeprotonated CD ac-
M in H2O was added to a solution of a-CD (7.2 ml, counts for less than 5% and no complex between this
0.15 M in H2O/D2O 50/50 v/v, pH 12.33 90.03) ther- species and La3 + ions was detected. Examination of
mostated at 25.090.1°C in a glass jacketed vessel Fig. 1 shows that titration curves for La3 + display
under Ar atmosphere. The pH and the solvent compo- some difference between a-CD on one hand and b- and
sition were adjusted and 0.5 ml were syringed from the g-CDs on the other hand. The association constants
titration vessel for analysis. pH values were measured with the three CDs are in the range log K=3.8–4
using a calibrated glass electrode in H2O/D2O 50/50 v/v (Table 1) and appear to be almost independent of the
with KOH solution. A slight turbidity appeared for nature of the receptor despite variations in diameter
[Tb3 + ]t /[a-CD]t \ 1.09, which increased with further [29] and polarity [30] of the cavity. Enthalpy (DH) and
metal addition and these were filtered before analysis. entropy (DS) changes were estimated from the tempera-
ture dependence of K in a limited temperature range
2.5. Lifetime determination (298.2–310.5 K). For a-CD, DH =28.591.8 kJ mol − 1
and DS =1699 6 J K − 1 mol − 1 while these values
Eu(5D0) and Tb(5D4) lifetimes were measured on amount to 299 4 kJ mol − 1 and 169 916 J K − 1
solutions 0.025–0.1 M in both bidistilled and degassed mol − 1, respectively, for both b- and g-CDs. These
water and deuterated water. Selective pulsed excitations
were achieved with a tunable dye laser (Lambda Physik
FL-3002) pumped with a XeCl excimer laser (Lambda
Physik EMG 101 MSC, pulse width: 18 ns). Rho-
damine 6G was used for Eu which was excited though
the 5D0 ’ 7F0 (578.5 nm) or 5D0 ’ 7F1 transition (588
nm) while coumarine 102 was used to excite Tb through
its 5D4 ’ 7F6 transition at 488 nm. The signal was
averaged with a SR-430 multichannel scaler coupled to
a SR-400 gated photon counter from Stanford Re-
search Systems. Decay curves were treated mathemati-
cally on an IBM PS/2-80 computer using a home made
algorithm. Data are averages of 3 – 5 determinations.

3. Results and discussion

3.1. Insertion of Ln 3 + ions in CDs at acidic pHs

The constants for the association of Ln3 + ions and


CDs were determined from potentiometric titrations of
metal ion solutions in absence and in presence of CDs.
A chloride salt was chosen as electrolyte because this
anion has the smallest association constant with CDs
(for a-CD, 1.5 M − 1 versus 30 – 50 M − 1 for ClO4− [26]).
The pH-potentiometric titration of Ln3 + ions alone
gave accurate values of the acidic hydrolysis constants,
very close to the literature values [27]. Analysis of the
pH titration curves obtained in presence of one equiva-
lent of CD was made introducing an association con-
stant [28] between the Ln3 + cation and CDs:
K = [Ln3 + ·CD]/[Ln3 + ]·[CD]. This model fitted conve-
Fig. 1. Potentiometric titration curves of La3 + (top) and Lu3 +
niently the titration curves and except for La3 + (aq), the
(bottom) ions in absence (1) and in presence of a-CD (2), b-CD (3),
pKa value of which is especially large, addition of CD g-CD (4). [Ln3 + ]0 =[CD]0 =0.002 M, V0 =20 ml, [TMAOH]tit =
to the solutions did not influence the number of base 0.034 M, m= 0.1 M (TMACl), T=25.0 90.1°C (the starting pH
equivalents needed to titrate the Ln3 + ions, which value was reached by adding acid before titration).
56 N. Fatin-Rouge, J.-C.G. Bünzli / Inorganica Chimica Acta 293 (1999) 53–60

Table 1
Association constants (Log K 92s) and thermodynamic parameters
( 9 2s) for the interaction of La3+ with a-, b-, and g-CDs (H2O,
m= 0.1 M TMACl)

T (K) a-CD b-CD g-CD

298.0 3.859 0.12 3.779 0.14 3.77 9 0.14


302.0 3.919 0.10 3.82 90.12 3.82 9 0.12
310.5 4.059 0.10 3.97 90.12 3.97 9 0.12
DH (kJ mol−1) 28.5 91.8 2994 29 9 4
DS (J K−1 mol−1) 1699 6 1699 16 169 9 16

Fig. 2. Association constants of Ln3 + cations with a-CD as deter-


thermodynamic parameters are indicative of an entropi- mined by competitive potentiometric titrations at 25.0 90.1°C, m =
cally driven association. Large and positive values for 0.1 M (NaCl), vs. the reciprocal of the ionic radii for a coordination
DH point to a weak metal – ligand interaction energy as number of 9 [3].
expected for complexation by neutral donor atoms. For
Lu3 + , titration curves with the three CDs are superim- 5, the following chemical shift differences were ob-
posable, leading to identical values of log K. In view of served: − 0.016, − 0.020, + 0.040, + 0.014, +0.105,
the low dependence of K(Ln3 + ) upon the host, associa- − 0.011, and + 0.009 for H1, H2, H3, H4, H5, H6a and
H6b, respectively (see Scheme 1). The largest variations
tion constants for the entire Ln3 + series were deter-
occur for the internal protons H3 and H5 and similarly
mined with a-CD only (Table 2). A plot of these
to what was observed for aromatic guest inclusion [32],
constants versus the inverse ionic radii for coordination
these protons are de-shielded upon addition of
number 9 (Fig. 2) displays variations along the lan- La3 + (aq). The external protons H2 and H4 sustain
thanide series which are reminiscent of a tetrade effect small shifts only and we can reasonably infer that the
[31] and which point to the influence of solvation and metal ion is located near the H5 protons, which are
coordination number change on the association pro- localized in the narrow part of the conically shaped
cess. Except for La3 + , for which K is the largest, the CD. To confirm this assignment, we have examined the
association constants are slightly larger when TMACl is 13
C NMR spectra and indeed, they show that carbon
used as inert electrolyte instead of NaCl, which reflects atoms C5 have the largest chemical shift variation and
some competition between Ln3 + and Na + . become shielded (Fig. 4). These data are consistent with
To determine whether the Ln3 + – CD association oc- steric polarization [33], the H5 and C5 chemical shifts
curs on the outside or inside the host cavity, we have varying in opposite direction upon metal addition.
recorded 1H and 13C NMR spectra at pH 4.5 in D2O as Moreover, the resulting shortening of the C5 –H5 bond
a function of R =[La3 + ]tot/[a-CD]tot (Fig. 3). For R= is confirmed by the large variation of the coupling
constant 1JCH with added metal ion (Table 3). To
Table 2 explain these data more quantitatively, we apply the
Hydrolysis constants (log b*1–3) a of Ln3+ ions and association con- theory of steric polarization of C–H bonds by a proton
stants (log Kass) with a-CD (H2O, T= 25.09 0.1°C); 92s indicated H% to the system C5 –H5 –La3 + (Scheme 1). The 13C
within parentheses
chemical shift (dC13) is a function of the angle C–H–La
Ln m = 0.1 M (NaCl) m =0.1 M (TMACl)
(u) and of the H–La distance (r) [33]:

log b*
1–3 log K log b*1–3 log K

La −26.12 (5) 3.98 (4) −26.41 (5) 3.85 (6)


Pr −24.00 (4) 2.91 (8) −24.28 (4) 3.25 (8)
Nd −23.54 (4) 2.8 (1) −23.88 (5) 3.40 (7)
Sm −22.34 (4) 2.7 (1) −22.78 (5) 3.18 (7)
Eu −22.21 (5) 2.6 (1) −22.74 (5) 2.79 (9)
Gd −22.16 (5) 2.5 (1) −22.76 (5) 2.78 (9)
Tb −21.61 (4) 2.78 (9) −22.32 (5) 3.20 (9)
Dy −21.45 (4) 3.23 (9) −22.14 (5) 3.77 (8)
Ho −21.46 (6) 3.66 (8) −22.08 (4) 3.74 (9)
Er −21.33 (5) 3.46 (9) −22.11 (5) 3.6 (1)
Tm −20.86 (8) 2.65 (9) −21.97 (6) 3.0 (1)
Yb −20.68 (6) 2.44 (9) −21.72 (7) 3.0 (1)
Lu −20.71 (6) 2.3 (1) −21.04 (8) 3.0 (2)
Fig. 3. Part of 1H NMR spectra of a-CD alone in D2O (a) and in
a
b* + 3
1–3 =[Ln(OH)3]·[H3O ] /[Ln
3+
]. presence of 5 equiv. of La3 + (b) at pH 4.5.
N. Fatin-Rouge, J.-C.G. Bünzli / Inorganica Chimica Acta 293 (1999) 53–60 57

dC13(ppm)= k× cos u× exp(−c ×r) (1)


3+
in which k and c are positive constants. If La sits in
the vicinity of C5, the angle u is such that cos u is
negative, henceforth the shielding of C5. Chemically
speaking, this can be explained both with an internu-
clear repulsion between La3 + (or its hydration sphere)
and the H5 protons. The large steric polarization effect
on the C5 –H5 bond indicates that the angle (C5 –H5 –
La) is close to 180° and, on the other hand, the less
intense de-shielding of the H3 protons and a weak steric
polarization effect on the C3 –H3 bond are consistent
with a longer H3 –La distance and a C3 –H3 –La angle
close to 90°. To substantiate this interpretation we have
performed molecular mechanic calculations of the Y3 + /
Scheme 1. Top: numbering of carbon atoms in the a-CD glucopyra- a-CD interaction with the AM1 package [34]. They
nose residue. Bottom: schematic representation of the host guest show the cation positioned at the level of the C5 carbon
position in the inclusion complex between a-CD and La3 + cation.
and of the anomeric oxygen atoms inside the cavity.
The reason for this positioning could be due to direct
(or to hydrogen bonded) interactions with the anomeric
oxygen atoms of the host. Analyzing the positions of
the C1 – 6 carbon atoms in the conically shaped a-CD we
get, as expected, a shielding for C5 and C6, and a
de-shielding for C2 and C3. The insertion of the Ln3 +
ions into the cavity induces some conformational
changes which are further reflected in variation of the
chemical shifts of C1 and C4, as reported for aromatic
host inclusion [32]. The lifetimes of the Eu(5D0) and
Tb(5D4) levels remain unchanged upon addition of one
equivalent CD to acidic solutions of the aqua ions
(0.1139 0.002 and 0.449 0.02 ms, respectively), which
means that the included cations are hydrated com-
pletely and that the Ln3 + –CD interaction is mainly an
outer-sphere process.

3.2. Association between tri6alent Tb and a-CD in


basic medium

Above pH 10, deprotonated secondary hydroxyl


groups of CDs were demonstrated to be able to bind
metal cations [35–38]. In our case, complexation of
terbium by partially deprotonated a-CD was detected
in a 0.002 M equimolar solution of the metal and the
Fig. 4. Top: Part of the 13C NMR spectra of a-CD in D2O/H2O host at pH\ 10.5. We have investigated the nature of
50/50 v/v at pH 4.5 in absence and in presence of La3 + ; R =[La3 + ]t / this interaction by spectrophotometry and NMR spec-
[a-CD]t : 0 (a), 0.25 (b), 0.5 (c), 1 (d), 1.25 (e), 2 (f); [a-CD]tot =0.02 troscopy. A pH of 12.30–12.33 was chosen because it is
M; the chemical shifts are referenced with respect to external
close to the average pKa of the host (12.33) and it
methanol. Bottom: Variation of 13C chemical shifts with R.

Table 3
Variation of the 1JCH coupling constants of a-CD upon La3+ addition (D2O/H2O 50/50 v/v, pH 4.5, 20°C)

Eq. of La3+ 1
JCH (Hz)

C1H1 C2H2 C3H3 C4H4 C5H5 C6H6

0 168.00 147.46 142.32 145.98 140.85 143.4


2 168.73 145.60 145.25 143.80 147.46 144.52
D(1JCH) (Hz) 0.73 −1.86 2.93 −2.18 6.61 1.1
58 N. Fatin-Rouge, J.-C.G. Bünzli / Inorganica Chimica Acta 293 (1999) 53–60

Fig. 5. Absorbance variation at 230 nm during the spectrophotomet-


ric titration of partially deprotonated a-CD with Tb3 + in D2O/H2O
50/50 v/v, pH 12.309 0.05, T= 25.090.1°C, l= 0.1 cm, [a-CD]0 =
9.2× 10 − 3 M.

allows us to use a simpler model for the interpretation


of the experimental data. The absorption spectra dis-
play an isosbestic point at 211 nm up to [Tb]tot/[a-
CD]tot =0.92. Beyond this ratio, the absorbance
increases linearly with the addition of the metal salt and
some turbidity appears for ratios R larger than 1. The
Fig. 6. 13C NMR titration curves for the addition of Tb3 + to
metal ion addition induces an absorption band at 230 partially de-protonated a-CD in D2O/H2O 50/50 v/v at pH 12.33 9
nm, corresponding to the 4f 8 “4f 75d transition of 0.03.
Tb3 + . Absorbance variations recorded at this wave-
length versus the number of metal equivalents clearly 13
C NMR titration of a-CD by Tb3 + at pH 12.33 in
points to a 1:1 stoichiometry of the complex (Fig. 5), in
D2O/H2O 50/50 v/v confirms the stoichiometry of the
agreement with the findings of Yashiro et al. for Pr3 +
complex (Fig. 6), but contrary to what was observed
[16]. Factor analysis of the spectrophotometric data
with transition metal ions, chemical shift variations
concluded to the presence of three main absorbing
point to a Tb3 + ion associated to the deprotonated
species, that we label Tb(aq), a-CD(aq) and [Tb(a-
secondary OH-3 hydroxyl function of the host
CD)](aq), since the exact composition and charges of
molecule: the C1 and C3 carbon atoms are de-shielded
these species cannot be determined. The following ap-
and their chemical shifts increase linearly up to R:1.
parent complexation constant (see Eq. (2)) was ob-
We interpret this as follows. At pH 13, the carbon
tained from the least-squares fit: log Kapp =4.24 90.06.
atoms of a-CD are de-shielded (0.42 ppm for C1), an
For each wavelength, the sum of residual absorbances
effect resulting from a conformational change [40].
was less than 0.025 and residues were dispersed ran-
Indeed, the first de-protonation occurs at the OH-2
domly. The fit yielded a molar absorbance coefficient
positions [41], so that the H–O(2)···H–O(3) hydrogen
o 230 nm = 430920 M − 1 cm − 1 typical of TbIII com-
bonds between adjacent glucosic units [42] are short-
plexes in water [39].
ened, resulting in a reduction of the larger ring section.
Therefore, a-CD is undergoing a conformational
Tb(aq)+ a-CD(aq) = [Tb(a-CD)]aq change, from its conical conformation in the fully pro-
tonated form to a cyclindrical conformation in the
[Tb(a-CD)] de-protonated form (Fig. 7). X-ray structures of metal
Kapp = (2) complexes with de-protonated a-CD evidence this con-
[Tb]·[a-CD]

Fig. 7. Schematic representation of the conformational changes occurring for a-CD upon de-protonation and subsequent complexation.
N. Fatin-Rouge, J.-C.G. Bünzli / Inorganica Chimica Acta 293 (1999) 53–60 59

formation [36]. At the pH at which the Tb3 + titration [2] A.P. de Silva, H.Q.N. Gunaratne, T. Gunnlaugsson, A.J.M.
has been performed, all the OH-2 groups are de-proto- Huxley, C.P. McCoy, J.T. Rademacher, T.E. Rice, Chem. Rev.
97 (1997) 1515.
nated so that the additional de-shielding of C1 (0.3 [3] J.-C.G. Bünzli, G.R. Choppin (Eds.), Lanthanide Probes in Life,
ppm) point to a further conformational change. Model Chemical and Earth Sciences, Elsevier Science, Amsterdam, 1989.
considerations show that an inverted cone structure in [4] V.T. D’souza, K.B. Lipkowitz, Chem. Rev. 98 (1998) 1743.
which the OH-3 groups are orientated conveniently to [5] C.M. Rudzinski, W.K. Hartmann, D.G. Nocera, Coord. Chem.
bind Tb3 + is the most probable. Lifetime measure- Rev. 171 (1998) 115.
[6] N. Sabbatini, M. Guardigli, J.M. Lehn, Coord. Chem. Rev. 123
ments of the Eu(5D0) and Tb(5D4) excited levels upon
(1993) 201.
addition of one equivalent CD (tEu =0.230 ms and [7] I. Tabushi, N. Shimizu, T. Sugimoto, M. Shiozuka, K. Yamamura,
tTb =1.08 ms, as compared to 0.113 and 0.442 ms, J. Am. Chem. Soc. 99 (1977) 7100.
respectively, for the aqua ions [43]) reflect the loss of [8] C.M. Rudzinski, D.S. Engebreton, W.K. Hatmann, D.G. Nocera,
3 – 4 inner-sphere water molecules upon complexation. J. Phys. Chem. A 102 (1998) 7442.
This result is consistent with the 13C NMR data and [9] A.R. Khan, P. Forgo, K.J. Stine, V.T. D’souza, Chem. Rev. 98
(1998) 1977.
model considerations which suggest that several coordi- [10] R. Corridini, A. Dossena, G. Galaverna, R. Marchelli, A. Panagia,
nating groups located at the O-3 position of the par- G. Sartor, J. Org. Chem. 62 (1997) 6283.
tially deprotonated CD enter the first coordination [11] F. Charbonnier, T. Humbert, A. Marsura, Tetrahedron Lett. 39
sphere of the cation. (1998) 3481.
[12] J.F. Stoddart, R. Zarzycki, Recl. Trav. Chim. Pays-Bas 107 (1988)
515.
[13] Y. Wang, S. Mendoza, A.E. Kaifer, Inorg. Chem. 37 (1998) 317.
4. Conclusion [14] S. Li, W.C. Purdy, Chem. Rev. 92 (1992) 1457.
[15] G.H. Zhao, S.G. Zhao, J.Z. Gao, J.W. Kang, Bull. Soc. Chim.
Several factors are invoked usually to account for the Belg. 106 (1997) 197.
association of CDs with various molecules [44]. In this [16] M. Yashiro, S. Miyama, T. Takarada, M. Komiyama, J. Inclus.
Phenom. Mol. Recogn. Chem. 17 (1994) 393.
work, we have demonstrated that at acidic pH, an
[17] D. French, M.L. Levine, J.H. Pazur, E. Norberg, J. Am. Chem.
entropically driven inclusion of trivalent Ln cations Soc. 71 (1949) 353.
occurs with the three native a-, b-, and g-CDs, leading [18] L Szente, in: J.L. Atwood, J.E.D. Davies, D.D. Mac Nicol, F.
to 1:1 complexes with log K in the range 2 – 4. NMR Vötgle, J.M. Lehn (Eds.), Comprehensive Supramolecular Chem-
analysis shows the metal ion coordinated to anomeric istry, Pergamon, Oxford, vol. 3, 1996, p. 274.
O-atoms of a-CD in the vicinity of the C5 carbon atom. [19] D.A. Skoog, F.J. Holler, T.A. Nieman, Principles of Instrumental
Analysis, 5th ed., Harcourt Brace College Publishers, Philadelphia,
On the other hand, in basic medium, partially deproto- 1998, p. 617.
nated a-CD molecules seem to act as first sphere coor- [20] P. Gans, A. Sabatini, A. Vacca, J. Chem. Soc., Dalton Trans. (1985)
dination ligand and the formation of a 1:1 complex 1195.
with trivalent terbium is accompanied by a large reor- [21] L.G. Sillèn, A.E. Martell, in: E. Högfeldt (Ed.), Stability Constants
ganization of the host molecule, so that the bonding of Metal – Ion Complexes, Pergamon, Oxford, 1982, Special Pub-
lication no. 21, Part A, 33.
can be considered as being on the outside of the
[22] H. Gampp, M. Maeder, C.J. Meyer, A.D. Zuberbühler, Talanta
cyclodextrin. In both kinds of complexes hydrogen 33 (1986) 943.
bonds probably play a major role in stabilizing the [23] A. Popov, K. Hallenga, Modern NMR Techniques and Their
supramolecular structures. This opens interesting per- Applications in Chemistry, Practical Spectroscopy Series, vol. 11,
spectives for the protection of Ln containing species Marcel Dekker, New York, 1991, p. 391.
against solvent interaction by native or derivatized [24] P. Colson, H.J. Jennings, I.C.P. Smith, J. Am. Chem. Soc. 96 (1974)
8081.
CDs. [25] D.J. Woods, F.E. Hruska, W. Saenger, J. Am. Chem. Soc. 99 (1977)
1735.
[26] Y. Matsui, M. Ono, S. Tokunaga, Bull. Chem. Soc. Jpn. 70 (1997)
Acknowledgements 535.
[27] E.N. Rizkalla, G.R. Choppin, K.A. Gschneidner Jr., L. Eyring,
(Eds.), Handbook on the Physics and Chemistry of Rare Earth,
This work is supported by grants from the Swiss
vol. 5, Elsevier Science, Amsterdam, 1991 (Chapter 103), pp. 393ff.
National Science Foundation. We thank the Fondation [28] K.A. Connors, in: J.L. Atwood, J.E.D. Davies, D.D. Mac Nicol,
Herbette (Lausanne) for the gift of spectroscopic equip- F. Vötgle, J.M. Lehn (Eds.), Comprehensive Supramolecular
ment and V. Foiret for measuring the luminescence Chemistry, vol. 3, Pergamon, Oxford, 1996, p. 205.
lifetimes. [29] F.W. Lichtenthaler, S. Immel, J. Inclus. Phenom. Mol. Recogn.
Chem. 25 (1996) 3.
[30] K.W. Street, W.E. Acree, Appl. Spectrosc. 42 (1988) 1315.
[31] M. Majdan, Monat. Chem. 119 (1988) 1079.
References [32] Y. Inoue, Ann. Rep. NMR Spectrosc. 27 (1993) 59.
[33] D.M. Grant, B. Vernon Cheney, J. Am. Chem. Soc. 89 (1967) 5315.
[1] R. Bergonzi, L. Fabrizzi, M. Licchelli, C. Mangano, Coord. Chem. [34] M.J.S. Dewar, E.V. Zoebisch, E.F. Healy, J.J.P. Stewart, J. Am.
Rev. 170 (1998) 31. Chem. Soc. 107 (1985) 3902.
60 N. Fatin-Rouge, J.-C.G. Bünzli / Inorganica Chimica Acta 293 (1999) 53–60

[35] Y. Matsui, T. Kurita, M. Yagi, T. Okayama, K. Mochida, Y. Handbook on the Physics and Chemistry of Rare Earth, vol. 24,
Date, Bull. Chem. Soc. Jpn. 48 (1975) 2187. Elsevier Science, Amsterdam, 1979, p. 171.
[36] P. Klüfers, H. Piotrowski, J. Uhlendorf, Chem. Eur. J. 3 (1997) [40] R.I. Gelb, L.M. Schwartz, J.J. Bradshaw, D.A. Laufer, Bioorg.
601. Chem. 9 (1980) 299.
[37] B.U. Nair, G.C. Dismukes, J. Am. Chem. Soc. 105 (1983) 124. [41] G. Wenz, Angew. Chem., Int. Ed. Engl. 33 (1994) 803.
[38] Y. Matsui, T. Kurita, Y. Date, Bull. Chem. Soc. Jpn. 42 (1972) [42] M. Vincendon, Bull. Chem. Soc. Fr. 3 (1981) 129.
3229. [43] J.-C.G. Bünzli, V. Foiret, unpublished results.
[39] W.T. Carnall Jr., in: K.A. Gschneidner Jr., L. Eyring (Eds.), [44] K.A. Connors, Chem. Rev. 97 (1997) 1325.

. .

You might also like