Polyhedron: Xiao-Yan Chen, George S. Goff, Brian L. Scott, Wolfgang Runde

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 7

Polyhedron 68 (2014) 80–86

Contents lists available at ScienceDirect

Polyhedron
journal homepage: www.elsevier.com/locate/poly

Comparison of structural variations of Ln(III) compounds


with (pyrazol-1-yl)acetic acid
Xiao-Yan Chen a, George S. Goff a, Brian L. Scott b, Wolfgang Runde c,⇑
a
Chemistry Division, Los Alamos National Laboratory, Los Alamos, NM 87545, United States
b
Material Physics and Applications Division, Los Alamos National Laboratory, Los Alamos, NM 87545, United States
c
Science Program Office, Los Alamos National Laboratory, Los Alamos, NM 87545, United States

a r t i c l e i n f o a b s t r a c t

Article history: As a continuation of our previous studies of light lanthanides (La, Ce, Pr, and Nd) with (Pyrazol-1-yl)acetic
Received 1 August 2013 acid (L), we reacted L with the heavier lanthanides in aqueous solution at pH 5 and synthesized lantha-
Accepted 14 October 2013 nide complexes of the general formula [Ln(L)3(H2O)2]H2O (Ln = Eu, 1; Gd, 2; Dy, 3; Ho, 4; Er, 5; Yb, 6; Lu,
Available online 19 October 2013
7) and [HoL(NO3)2(H2O)3] (8). All complexes were characterized by single crystal X-ray diffraction anal-
ysis revealing one-dimensional chain formations. Three distinct crystallographic structures are governed
Keywords: by the different coordination modes of the carboxylate groups in L: terminal bidentate, bridging biden-
Rare earth compounds
tate and tridentate coordination in 1–3, terminal and bridging bidentate coordination in 4–7, and bridg-
(Pyrazol-1-yl)acetic acid
Crystal structures
ing bidentate coordination in 8. The solid state UV–Vis–NIR diffuse reflectance spectra and the solution
Optical spectroscopy UV–Vis–NIR absorption spectra differ, suggesting different coordination environments in solution and
solid state. The coordination-induced shifts of the 13C NMR signals for [Lu(L)3(H2O)2]H2O (7) in D2O show
that the carboxylate groups of the ligand are coordinated with the Lu(III) ion in solution.
Ó 2013 Elsevier Ltd. All rights reserved.

1. Introduction hydrothermal conditions. Among the known Ln complexes, twelve


coordination modes were observed including terminal and bridg-
The pyrazole heterocycle possesses various biological proper- ing carboxylate groups or bridging and chelating pyrazole N and
ties such as being anti-inflammatory [1], anti-tubercular [2], anti- carboxylate groups.
microbial [2,3], and anti-cancer [4] etc., which have increased the Our research group has been developing task-specific ionic
interest of pyrazole-metal complexes for pharmaceutical applica- liquids (TSILs) with covalently tethered functional groups to dis-
tions [5]. Since the first report of cis-dichlorobis(pyrazole)plati- solve f-elements for the processing of used nuclear fuel [12]. Tak-
num(II) which possesses cancer-targeting abilities in 2000 [6], ing the advantage of wider electrochemical window of
pyrazole-based compounds with palladium, platinum, copper, pyrazolium ILs than some other common ILs such as imidazolium,
gold, zinc, cobalt, nickel, iron, silver and gallium have been inves- sulfonium or guanidinium ILs [13], we incorporated (pyrazol-1-
tigated for therapeutic drug development [5]. Among these transi- yl)acetic acid (L) into ILs, which can be deprotonated as a
tion metal complexes, over 20 different coordination modes of functionalized pyrazolate anion. Prior to our work, the only re-
pyrazole have been revealed [7]. ported crystal structure for L with a transition metal was a Cu(II)
With the increasing applications of lanthanides (Ln) in industry, complex [14]. We have previously reported four light lanthanide
technology, and medicine [8], studies of lanthanide compounds complexes with L of the general formula [Ln(L)3(H2O)2]nH2O
with pyrazole [9] have drawn more and more attention. However, (Ln = La, Ce, Pr, n = 2; Ln = Nd, n = 1) [15]. All of the complexes re-
pyrazoles can only be deprotonated by treatment with lanthanides veal a one-dimensional chain structure with tridentate and biden-
and diarylmercurials or iodine activated lanthanides [7,9]. In order tate bridging ligands connecting adjacent lanthanide centers along
to react with trivalent lanthanides at mild conditions, the pyrazole the a-axis; however, two different structure types are found with
ring has been modified by incorporating functional groups such as the structural break occurring between Pr3+ and Nd3+. In addition
carboxylate groups. To date, crystal structures have been reported to terminal bidentate and bridging tridentate ligands found in La,
for lanthanides with pyrazole-3,5-dicarboxylic acid [10] and 4-iod- Ce and Pr complexes, a third coordination mode (bidentate bridg-
opyrazole-3,5-dicarboxylic acid [11], both synthesized under ing) is observed in the Nd complex caused by the lanthanide con-
traction. As a continuation of this study, we report herein eight
complexes with intermediate and heavy lanthanides of the general
⇑ Corresponding author. Tel.: +1 505 667 3350. formula [Ln(L)3(H2O)2]H2O (Ln = Eu, 1; Gd, 2; Dy, 3; Ho, 4; Er, 5;
E-mail address: runde@lanl.gov (W. Runde).

0277-5387/$ - see front matter Ó 2013 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.poly.2013.10.015
X.-Y. Chen et al. / Polyhedron 68 (2014) 80–86 81

Yb, 6; Lu, 7) and [HoL(NO3)2(H2O)3] (8) which have been synthe- 5; Yb, 6; Lu, 7) and [HoL(NO3)2(H2O)3] (8). Photographs of the crys-
sized and structurally characterized by single crystal X-ray diffrac- tals are shown in Fig. 1.
tion. Spectroscopic methods (Raman, UV–Vis–NIR absorbance,
diffuse reflectance, and 13C NMR spectroscopies) were used to ex- 3.1. Single crystal X-ray diffraction studies
plore the coordination chemistry in solution and solid state.
Crystal data collection and refinement details for complexes
1–8 are listed in Tables 1 and 2, while selected bond lengths and
2. Experimental methods angles are provided in the Supporting information (Table S1). Com-
pared to the previously reported structures of lanthanides with
Pyrazole (98%) was used as received from Acros. All other chem- pyrazole-3,5-dicarboxylic acid [10], L only has one carboxylate
icals were purchased from Fisher Scientific, Inc., and solutions were group and the acetic acid is substituted directly to one of the
prepared with distilled deionized water (specific resis- N atom of the pyrazole ring. This substitution excludes the other
tance P 18.0 MX cm). (Pyrazol-1-yl)acetic acid (L) was prepared N atom from chelating lanthanides, leading to the formation of
as described in the literature [14]. Stock solutions of L (1 M) were one-dimensional chains for complexes 1–8. The different coordina-
prepared gravimetrically by dissolving L in water with the pH ad- tion modes of L observed in 1–3 (terminal bidentate, bridging
justed to 5 by adding NaOH. bidentate and tridentate), 4–7 (terminal and bridging bidentate),
Lanthanide stock solutions were prepared by dissolving and 8 (bridging bidentate) result in two distinct crystallographic
Ln(NO3)36H2O (99.99% purity) in water. Seven lanthanide(III) structures.
complexes (Eu (1), Gd (2), Dy (3), Ho (4), Er (5), Yb (6), and Lu
(7)) with L were prepared under mild conditions by adding 1 mL 3.1.1. Structural characterization of heavy lanthanide complexes of the
of a 1 M Ln(III) nitrate stock to 3 mL of the L stock solution (1 M) type [Ln(L)3(H2O)2]H2O
in a 20 mL borosilicate scintillation vial. The vial was capped and Complexes 4–7 are isostructural, crystallizing in the monoclinic
crystallization of the lanthanide compounds occurred at room tem- space group P21/n. Two coordination modes of L are evident in
perature within a couple of days. During the crystallization, the these complexes: terminal bidentate (Scheme 1a) and bridging
solution pH remained constant at 5. [HoL(NO3)2(H2O)3] (8) was bidentate (Scheme 1b). Fig. 2 shows the coordination environment
prepared in the same method using a 1:1 metal to ligand ratio. of Ho3+ in complex 4 as a representative example. The unit cell of 4
Crystals of 1–8 were isolated and their structures were determined contains the one lanthanide, three L and three water molecules.
by single crystal X-ray diffraction. The Ho3+ atom is eight-coordinate with two oxygen atoms from
Crystals were removed from the solutions, mounted using a ny- terminal bidentate L, four oxygen atoms from bridging bidentate
lon cryoloop and Paratone-N oil, and cooled with a liquid nitrogen L, and two water molecules completing the square antiprismatic
vapor stream. The data were collected on a Bruker D8 diffractom- coordination geometry. Bidentate bridging of L in 4 connects Ho1
eter with an APEX II charge-coupled-device (CCD) detector and a and Ho1B with a distance of 4.525(1) Å. The least squares planes
Bruker Kryoflex low temperature device. The instrument was of the pyrazole and carboxylate group L show a dihedral angle of
equipped with a graphite monochromatized Mo Ka X-ray source 74.55(2)°. The O3–C6–O4 angle is 122.4(2)°. In the second biden-
(k = 0.71073 Å). A hemisphere of data was collected using x scans, tate bridging L, the carboxylate group bridges Ho1 and Ho1A with
with 10-s frame exposures and 0.5° frame widths. Data collection the Ho–O bond lengths of 2.336(2) and 2.264(2) Å. The resulting
and initial indexing and cell refinement were handled using APEX Ho1  Ho1A distance is 4.895(1) Å. The difference between the
II software [16]. Frame integration, including Lorentz-polarization two C–O bond lengths in a single carboxylate group of L follows
corrections, and final cell parameter calculations were carried out a different order in 4: terminal bidentate (0.01 Å) > bridging biden-
using SAINT+ software [17]. The data were corrected for absorption tate (0.001 Å). Such a trend indicates the weakest electron delocal-
using redundant reflections and the SADABS program [18]. Decay of ization occurs in the carboxylate group of the terminal bidentate L,
reflection intensity was not observed as monitored via analysis of totally opposite from the previous observation. The bidentate
redundant frames. The structure was solved using direct methods bridging L links neighboring Ho3+ ions to form a one-dimensional
and difference Fourier techniques. Hydrogen atom positions were chain structure with alternating Ho  Ho distances of 4.525(1)
idealized on methyl and pyrazole rings, but were not refined on and 4.895(1) Å (Fig. 3).
water molecules. The final refinement included anisotropic tem-
perature factors on all non-hydrogen atoms. Structure solution, 3.1.2. Structural characterization of [HoL(NO3)2(H2O)3]
refinement, and creation of publication materials were performed Complex 8 was obtained by varying the metal to ligand ratio to
using SHELXTL [19]. 1:1 and crystallizes in the monoclinic space group P21/c. In con-
Raman spectra of solid samples were collected for all com- trast to the eight-coordinate Ho in (4), the Ho in 8 is nine-coordi-
pounds using a Thermo Scientific DXR SmartRaman spectrometer nate with two oxygen atoms from two bridging bidentate L
with a 780 nm excitation laser. Diffuse reflectance spectra were molecules (Scheme 1b), four oxygen atoms from two nitrate an-
measured for ground single crystals using a Cary 5 UV–Vis–NIR ions, and three coordinated water molecules (Fig. 4). The coordina-
spectrophotometer with a diffuse reflectance attachment, while tion geometry of Ho3+ can be described as a tricapped trigonal
solution absorbance spectra were obtained by using a Cary 5 prism. Bidentate L bridges the Ho3+ ions to form an extended chain
UV–Vis–NIR spectrophotometer. Complex 7 was characterized by structure along the c axis with the Ho1–O1 and Ho1A–O2 bond
13
C NMR spectroscopy in D2O at room temperature on a Bruker lengths of 2.293(2) and 2.256(1) Å, respectively. The difference be-
ARX-300 spectrometer equipped with 5- and 10-mm multinuclear tween the bond lengths of C1–O1 (1.250(2) Å) and C1–O2
probes. The frequency was 75 MHz for the 13C nuclei. (1.264(2) Å) is larger than that of the bidentate bridging ligands
in complex 4, indicating a weaker electron delocalization. The
Ho  Ho distance is 5.704(1) Å, which is significantly larger than
3. Results and discussion that in complex 4 (4.895(1) Å). The average Ho–O bond length
for coordinated water molecules of 2.391(2) Å is comparable to
The reaction of seven lanthanides with (pyrazol-1-yl)acetic acid that in 4 (2.378(2) Å). The pyrazole rings in the one-dimensional
(L) resulted in the formation of crystalline complexes with the gen- chain are almost parallel to each other with the dihedral angle of
eral formula [Ln(L)3(H2O)2]H2O (Ln = Eu, 1; Gd, 2; Dy, 3; Ho, 4; Er, 1.86(9)°.
82 X.-Y. Chen et al. / Polyhedron 68 (2014) 80–86

Fig. 1. Photographs of single crystals [Ln(L)3(H2O)2]H2O, Ln = Eu (A), Gd (B), Dy (C), Ho (D), Er (E), Yb (F), and Lu (G) and [HoL(NO3)2]3H2O (H).

Table 1
Selected crystal data and structure refinements for [Ln(L)3(H2O)2]H2O (Ln = Eu, 1; Gd, 2; Dy, 3; Ho, 4).

1 2 3 4
T (K) 140(1) 140(1) 140(1) 140(1)
Formula C15H21EuN6O9 C15H21GdN6O9 C15H21DyN6O9 C15H21HoN6O9
Formula weight 581.34 586.63 591.88 594.31
Crystal system monoclinic monoclinic monoclinic monoclinic
Space group P21/n P21/n P21/n P21/n
a (Å) 9.2515(5) 9.2490(7) 9.2976(8) 9.348(1)
b (Å) 15.3810(8) 15.3574(11) 15.2803(13) 15.2182(16)
c (Å) 14.4142(7) 14.4077(11) 14.3756(13) 14.3622(15)
b (°) 100.2910(10) 100.360(1) 100.969(1) 101.452(1)
V (Å3) 2018.11(18) 2013.1(3) 2005.0(3) 2002.5(4)
Z 4 4 4 4
q (g/cm3) 1.913 1.936 1.961 1.971
l (mm 1) 3.170 3.357 3.790 4.014
F(0 0 0) 1152 1156 1164 1168
Crystal size (mm) 0.24  0.06  0.02 0.22  0.04  0.02 0.24  0.06  0.04 0.20  0.06  0.04
h (°) 1.95–28.45 1.96–28.43 1.96–28.35 1.97–28.52
Index ranges 12 6 h 6 12, 12 6 h 6 12, 12 6 h 6 12, 12 6 h 6 12,
20 6 k 6 20, 20 6 k 6 20, 20 6 k 6 20, 20 6 k 6 20,
19 6 l 6 19 18 6 l 6 19 19 6 l 6 18 19 6 l 6 19
Minimum and maximum transmission 0.5166 and 0.9393 0.5254 and 0.9359 0.4632 and 0.8632 0.5007 and 0.8559
Goodness-of-fit (GOF) on F2 1.493 0.773 1.569 1.128
R1, R2 [I > 2r(I)] 0.0218, 0.0619 0.0243, 0.0844 0.0220, 0.0634 0.0232. 0.0474
R1, R2 (all data) 0.0263, 0.0637 0.0315, 0.0948 0.0259, 0.0650 0.0302, 0.0501
3
Largest difference in peak and hole (e Å ) 1.032, 0.765 0.948, 1.530 0.889, 0.842 0.666, 0.715

3.1.3. Structural characterization of intermediate lanthanide [Nd(L)3(H2O)2]H2O [15]. The Eu1–O3 and Eu1–O4 bond lengths
complexes of the type [Ln(L)3(H2O)2]H2O are 2.451(2) and 2.716(2) Å, respectively, which are more diverged
Complexes 1–3 are isostructural with the previously published than the corresponding Nd–O distances (2.501(2) and 2.705(2) Å)
structure of [Nd(L)3(H2O)2]H2O [15], crystallizing in the mono- in [Nd(L)3(H2O)2]H2O [15]. The Eu1B–O4 bond length
clinic space group P21/n, with bidentate and tridentate bridging (2.406(2) Å) with the bridging l3–oxygen atom is shorter than
(Scheme 1c) ligands forming an extended one-dimensional chain. those of Eu1–O3 and Eu1–O4. The Eu1–O4–Eu1B bond angle is
In the following, complex 1 is chosen to discuss the crystallo- 116.18(7)°, which is similar to that of Nd–O–Nd (115.93(8)°) in
graphic features found in complexes 1–3. The unit cell contains [Nd(L)3(H2O)2]H2O [15]. For the tridentate bridging L, the C-O
one unique Eu3+ center, three deprotonated ligands as monoanions bond lengths of the carboxylate group are 1.247(3) and
and two coordinated water molecules with one lattice water mol- 1.268(3) Å, which suggest that there is some electron delocaliza-
ecules (Fig. 5). As observed previously for [Nd(L)3(H2O)2]H2O [15], tion in the carboxylate groups compared with the C–O bond
three coordination modes of L are present in complex 1: terminal lengths of the carboxylate group in the free ligand (1.207(2) and
bidentate, bridging bidentate and tridentate modes. The tridentate 1.309(2) Å) [14]. Such electron delocalization is even larger than
bridging L links Eu1 and Eu1B with a distance of 4.351(1) Å, which that in [Nd(L)3(H2O)2]H2O (C–O, 1.243(3) and 1.274(3) Å) [15].
is slightly shorter than the Nd  Nd distance (4.379(1) Å) in Within the tridentate ligand, the dihedral angle of the least squares
X.-Y. Chen et al. / Polyhedron 68 (2014) 80–86 83

Table 2
Selected crystal data and structure refinements for [Ln(L)3(H2O)2]H2O (Er, 5; Yb, 6; Lu, 7) and [HoL(NO3)2(H2O)3] (8).

5 6 7 8
T (K) 293(2) 140(1) 140(1) 140(1)
Formula C15H21ErN6O9 C15H21N6O9Yb C15H21LuN6O9 C5H11HoN4O11
Formula weight 596.64 602.42 604.35 468.11
Crystal system monoclinic monoclinic monoclinic monoclinic
Space group P21/n P21/n P21/n P21/c
a (Å) 9.4051(7) 9.4451(7) 9.4740(8) 11.0734(8)
b (Å) 15.1458(11) 15.0869(12) 15.0368(13) 13.8862(11)
c (Å) 14.3526(10) 14.3477(11) 14.3392(12) 8.7857(7)
b (°) 101.954(1) 102.482(1) 102.857(1) 112.267(1)
V (Å3) 2000.2(3) 1996.2(3) 1991.5(3) 1250.21(17)
Z 4 4 4 4
q (g/cm3) 1.981 2.005 2.016 2.487
l (mm 1) 4.259 4.748 5.020 6.399
F(0 0 0) 1172 1180 1184 896
Crystal size (mm) 0.20  0.14  0.06 0.20  0.04  0.02 0.12  0.08  0.06 0.20  0.06  0.02
h (°) 1.98–28.43 1.98–28.35 1.99–28.37 1.99–28.47
Index ranges 12 6 h 6 12, 12 6 h 6 12, 12 6 h 6 12, 14 6 h 6 14,
19 6 k 6 20, 19 6 k 6 19, 19 6 k 6 19, 17 6 k 6 18,
19 6 l 6 19 18 6 l 6 18 18 6 l 6 19 11 6 l 6 11
Minimum and maximum transmission 0.4830 and 0.7842 0.4502 and 0.9110 0.5841 and 0.7527 0.3610 and 0.8827
Goodness-of-fit (GOF) on F2 1.153 1.206 1.482 1.295
R1, R2 [I > 2r(I)] 0.0204, 0.0429 0.0244, 0.0608 0.0182, 0.0537 0.0140, 0.0341
R1, R2 (all data) 0.0265, 0.0450 0.0331, 0.0642 0.0213, 0.0551 0.0155, 0.0350
3
Largest difference in peak and hole (e Å ) 0.592, 0.608 1.173, 0.854 0.855, 0.725 0.635, 0.648

Scheme 1. Coordination modes of (Pyrazol-1-yl)acetic acid in complexes 1–8.

Fig. 3. One-dimensional chain of [Ho(L)3(H2O)2]H2O (4). Ho, green; C, grey; O, red;


N, blue. The hydrogen atoms and the lattice water molecules have been removed for
clarity. (Color online.)

The carboxylate group in the bidentate bridging L bridges Eu1


and Eu1A with the Eu–O bond lengths of 2.321(2) and 2.403(2) Å
and the Eu1  Eu1A distance is 4.991(1) Å. The least squares planes
of the pyrazole and carboxylate group of the bidentate bridging L
show a larger dihedral angle of 80.97(20)° than that in the triden-
Fig. 2. The coordination environment of Ho(III) in [Ho(L)3(H2O)2]H2O (4) showing tate ligand. In the bidentate terminal ligand, the dihedral angle of
the relevant atom labeling scheme. The hydrogen atoms and the lattice water the least squares planes of the pyrazole and carboxylate group in
molecules have been removed for clarity.
the bidentate L is 88.07(30)°, which is much larger than found in
the tridentate L. The difference between the two C–O bond lengths
planes of the pyrazole ring (N3, N4, C8, C9 and C10) and carboxyl- in a single carboxylate group of L decreases in the following order:
ate group (C6, O3 and O4) is 75.25(20)°, which is smaller than tridentate bridging (0.021 Å) > bidentate bridging (0.01 Å) > termi-
those in the free L (80.76(6)°) and in [Nd(L)3(H2O)2]H2O nal bidentate (0.001 Å). Such a trend indicates that the strongest
(79.68(12)°) [15]. electron delocalization occurs in the carboxylate group of the
84 X.-Y. Chen et al. / Polyhedron 68 (2014) 80–86

lanthanide contraction [20]. The lanthanide series can be divided


into three groups according to ionic radii: the lighter La–Pm
(Group 1), the intermediate Sm–Dy (Group 2), and the heavier
Ho–Lu (Group 3). As shown in Fig. 7, the average Ln–O bond
lengths with the terminal bidentate L, bridging bidentate L, and
coordinated water molecules shorten with decreasing ionic radii
[21] in accordance with the well-known lanthanide contraction
[20]. For the carboxylate group in the tridentate bridging L, the
Ln–O3 bond lengths decrease while the Ln–O4 (l3–bridging oxy-
gen) bond lengths increase. With the continuous elongation of
the Ln–O4 bond, the breakage of the Ln–O4 bond was observed
at Ho (4) resulting in the coordination mode change of the triden-
tate bridging to bidentate bridging. The Ln  O4 distances for Ho,
Er, Yb and Lu are shown as unconnected red squares in Fig. 7.

3.2. Spectroscopic characterization of solid-state and solution species

3.2.1. Solid-state Raman spectroscopy


The solid-state FT-Raman spectra for all the lanthanide com-
pounds are very similar (1 is shown in Fig. 8 as an example). We
have previously reported the Raman spectroscopic characteriza-
tion of (pyrazol-1-yl)acetic acid [15]. Consistent with a longer
C@O bond length (1.247(3)–1.261(3) Å), the carboxylate symmet-
Fig. 4. The coordination environment of Eu(III) in [Eu(L)3(H2O)2]H2O (1) showing ric stretching (m(C@O)) mode in the Eu3+ complex (1) at
the relevant atom labeling scheme. The hydrogen atoms and the lattice water 1587 cm 1 is shifted to lower frequency compared to the free li-
molecules have been removed for clarity.
gand (1707 cm 1, 1.207(2) Å) [14]. The Raman shift at 1463 cm 1
is assigned to the pyrazole ring stretching mode [22]. Upon coordi-
nation with Eu3+, the C–O symmetric and asymmetric stretching
modes are shifted slightly to 1287 and 1167 cm 1 compared to
the free ligand (1300 and 1174 cm 1) [15]. The peak at
1327 cm 1 can be assigned as either the pyrazole ring stretching
mode or the d(C–OH) bending mode. Similar peak shifts and inten-
sity changes have been reported for the light lanthanide complexes
with L [15]. The medium–strong to strong peaks at 1394 and
1097 cm 1 in 1 are due to the pyrazole ring stretching modes
[22]. These ring stretching peaks shift to lower frequencies with
additional shoulder peaks caused by the different coordination
modes of L in the lanthanide complexes compared with those in
the free ligand (1406 and 1098 cm 1) [15]. The strong peak at
1066 cm 1 is assigned as C–H in-plane bending vibration mode
[23]. Similar peak shifts and splitting can be seen for the pyrazole
ring bending mode with multiple peaks around 946 and 930 cm 1
in 1. The weak Raman band at 797 cm 1 is attributed to the d(CH)
bending mode. The multiple peaks around 722 and 709 cm 1 can
be assigned as the bending mode of d(O@CO), which shift to higher
frequency compared to the free ligand (664 cm 1) [15]. In the low
wavenumber region of the complex, we observe a weak band at
431 cm 1, which might be a combination of ring bending modes
and a new Eu–O vibration mode [23,24].

3.2.2. UV–Vis–NIR spectroscopy


Fig. 5. One-dimensional chain of [Eu(L)3(H2O)2]H2O (1). Eu, green; C, grey; O, red; In order to compare coordination modes in solid-state and solu-
N, blue. The hydrogen atoms and the lattice water molecules have been removed for tion, we compared the solid-state diffuse reflectance spectra of
clarity. (Color online.) [Ln(L)3(H2O)2]H2O (Ln = Eu (1), Dy (3), Ho (4), and Er (5)) with
the UV–Vis–NIR absorption spectra of the corresponding solid dis-
terminal bidentate L, as observed previously for the light lantha- solved in water (Figs. S1–S3). The electronic absorbance spectra for
nide complexes with L [15]. The two coordinated water molecules trivalent lanthanide complexes are characterized by a number of
are cis to each other with an O7–Eu1–O8 angle of 67.93(7)°. Each Laporte-forbidden f–f transitions [25]. In the solution UV–Vis–
Eu3+ ion is connected to the neighboring Eu3+ ions by alternate tri- NIR spectra, the molar absorptivities of all complexes are similar
dentate and bidentate bridging L forming one-dimensional chain with those of the Ln3+(aq) ion in dilute perchloric acid. The
along the a-axis (Fig. 6), with the Eu  Eu distances alternating be- UV–Vis–NIR spectrum of compound 4 dissolved in water (Fig. 9)
tween 4.991(1) and 4.351(1) Å. displays the characteristic transitions for the Ho3+ aquo ion at
361 (5I8 to 3H6), 416 (5I8 to (5G, 3G)5), 451 (5I8 to 5F1), 468 (5I8 to
5
3.1.4. Structural comparison across the lanthanide series G4), 473 (5I8 to 5F2), 485 (5I8 to 5F3), 537 (5I8 to 5F4 (5S2)), 641
The structural break between Dy (3) and Ho (4) revealed by sin- (5I8 to 5F5), and 656 (5I8 to 5F5) nm. The molar absorptivity of 2.5
gle crystal X-ray diffraction analysis can be explained by the L mol 1 cm 1 at 451 nm is comparable to that reported for the
X.-Y. Chen et al. / Polyhedron 68 (2014) 80–86 85

Fig. 6. Thermal ellipsoids (50% probability) of [HoL(NO3)2(H2O)3] (8). The hydrogen atoms and the lattice water molecules have been removed for clarity.

Fig. 9. Solid state UV–Vis–NIR diffuse reflectance spectrum (red) of [Ho(L)3(H2-


Fig. 7. Comparison of the average Ln–O bond lengths with decreasing ionic radii.
O)2]H2O (4) and UV–Vis–NIR absorbance spectrum of [Ho(L)3(H2O)2]H2O (4)
Ln–Oterminal bidentate, blue; Ln–O3, purple; Ln–O4, red; Ln–O5, green; Ln–Owater,
dissolved in water (black). The Ho3+(aq) absorbance peaks in dilute perchloric acid
black. The Ho–Obridging bidentate bond length in 8 is shown as green triangle. Ionic
are shown as dotted lines [25a]. (Color online.)
radii for eight- and nine-coordinate lanthanide cations were taken from Shannon
[21]. (Color online.)

3.2.3. 13C NMR Spectroscopy of Solution Species


13
C NMR spectroscopy was used to further explore the coordi-
nation of L with the Ln(III) ions in solution. The 13C NMR spectrum
of the diamagnetic Lu(III) complex 7 dissolved in D2O (pH 5) dis-
plays 13C NMR features at 54.38, 106.02, 132.09, 139.86, and
176.22 ppm, which shifted from those of the free ligand
(52.24 ppm (CH2), 106.61 (CH), 132.90 (CH), 139.81 (CH),
172.23 ppm (COOH)) [15]. The coordination-induced shift (Dd) of
3.99 ppm for the COOH group is caused by the coordination of
the Ln(III) ion with L via the carboxylate groups and the associated
deshielding of the carboxylate carbon atom. A similar Dd value of
4.08 ppm (COOH) was observed for the La3+ complex reported pre-
viously [15]. The methylene carbon (CH2COOH) exhibits a larger
Dd value of 2.14 ppm than the pyrazole ring carbon (maximum
Dd of 0.81 ppm) due to its closer proximity to the Lu3+ ion. Only
one set of signals was present in the 13C NMR spectrum of 7, indi-
Fig. 8. Solid-state Raman spectrum of [Eu(L)3(H2O)2]H2O (1). cating that all coordinated ligands in the Lu3+ solution complex are
equivalent. The absence of NMR shifts characteristic of the uncom-
plexed ligand indicates that all ligand is bound to Lu3+ in solution.
Ho(aq)3+ ion in dilute perchloric acid (2.9 L mol 1 cm 1) [25a].
However, the solid state diffuse reflectance spectra of 4 does not
match the absorbance of the corresponding solution species at 4. Conclusion
pH 5, suggesting different coordination environments around the
lanthanide ion in solution and solid state. Both relative intensity We have synthesized and characterized eight novel lanthanide
ratios and peak shapes differ in the solution and solid state spectra complexes with (pyrazol-1-yl)acetic acid (L) of the general formula
with peaks shifted up to 4 nm. [Ln(L)3(H2O)2]H2O (Ln = Eu, 1; Gd, 2; Dy, 3; Ho, 4; Er, 5; Yb, 6; Lu,
86 X.-Y. Chen et al. / Polyhedron 68 (2014) 80–86

7) and [HoL(NO3)2(H2O)3] (8). Complexes 1–7 were prepared using (b) H.V.R. Dias, K.H. Batdorf, M. Fianchini, H.V.K. Diyabalanage, S. Carnahan, R.
Mulcahy, A. Rabiee, K. Nelson, W.L.G. van, J. Inorg. Biochem. 100 (2006) 158.
a 1:3 metal to ligand ratio. Although complexes 1–7 possess a one-
[4] (a) B. Insuasty, A. Tigreros, F. Orozco, J. Quiroga, R. Abonia, M. Nogueras, A.
dimensional chain structure along the a-axis, two different struc- Sanchez, J. Cobo, Bioorg. Med. Chem. 18 (2010) 4965;
ture types are found with the structural break occurring between (b) D. Pal, S. Saha, S. Singh, Int. J. Pharm. Pharm. Sci. 4 (2012) 98.
Dy3+ (3) and Ho3+(4). Three coordination modes of L were found [5] F.K. Keter, J. Darkwa, Biometals 25 (2012) 9.
[6] K. Sakai, Y. Tomita, T. Ue, K. Goshima, M. Ohminato, T. Tsubomura, K.
in 1–3: terminal bidentate, bridging bidentate, and bridging triden- Matsumoto, K. Ohmura, K. Kawakami, Inorg. Chim. Acta 297 (2000) 64.
tate. The tridentate bridging L changes to bidentate bridging in 4–7 [7] M.A. Halcrow, Dalton Trans. (2009) 2059.
caused by the smaller ionic radii of the Ln(III) ions heavier than [8] (a) W. Xu, K. Kattel, J.Y. Park, Y. Chang, T.J. Kim, G.H. Lee, Phys. Chem. Chem.
Phys. 14 (2012) 12687;
Dy3+. The change from a tridentate bridging to bidentate bridging (b) W. Deng, E.M. Goldys, Langmuir 28 (2012) 10152;
mode also decreases the coordination number of the Ln3+ ions from (c) C. Liu, L. Zhang, Q. Zheng, F. Luo, Y. Xu, W. Weng, Sci. Adv. Mater. 4 (2012)
nine (1–3) to eight (4–7). Complex 8 was prepared using a 1:1 me- 1;
(d) C. Vidaud, D. Bourgeois, D. Meyer, Chem. Res. Toxicol. 25 (2012) 1161;
tal to ligand ratio, and contains only bidentate bridging ligands in (e) P. Dissanayake, D.J. Averill, M.J. Allen, Sci. Synth. Knowl. Updates (2011) 1;
the crystal structure. (f) L.J. Charbonniere, Curr. Inorg. Chem. 1 (2011) 2.
The shifts and intensity changes between the solid state UV– [9] (a) M. Viciano-Chumillas, S. Tanase, L.J. de Jongh, J. Reedijk, Eur. J. Inorg. Chem.
(2010) 3403;
Vis–NIR diffuse reflectance spectra and the solution UV–Vis–NIR (b) R. Kawahata, T. Tsukuda, T. Yagi, A. Fuyuhiro, S. Kaizaki, J. Alloys Compd.
spectra suggest a different coordination environment of Ln(III) 408–412 (2006) 976;
with L in solution and solid state. A single set of ligand NMR shifts, (c) G.B. Deacon, C.M. Forsyth, A. Gitlits, B.W. Skelton, A.H. White, Dalton Trans.
(2004) 1239;
different from those of the free ligand in solution, indicates one
(d) G.B. Deacon, A. Gitlits, B.W. Skelton, A.H. White, Chem. Commun. (1999)
type of Ln3+–L solution species. Both NMR and vibrational spectro- 1213;
scopic data clearly identify that the carboxylate group in L partic- (e) C.C. Quitmann, V. Bezugly, F.R. Wagner, K. Müller-Buschbaum, Z. Anorg.
ipates in the Ln(III) bonding. This information on the structure and Allg. Chem. 632 (2006) 1173;
(f) G.B. Deacon, P.C. Junk, A. Urbatsch, Aust. J. Chem. 65 (2012) 802.
bonding of lanthanide(III) with (pyrazol-1-yl)acetic acid in solution [10] (a) J. Zhao, L.-S. Long, R.-B. Huang, L.-S. Zheng, Dalton Trans. (2008) 4714;
and solid state aids our fundamental understanding of the lantha- (b) J. Xia, B. Zhao, H.-S. Wang, W. Shi, Y. Ma, H.-B. Song, P. Cheng, D.-Z. Liao, S.-
nide coordination with new TSILs composed of deprotonated (pyr- P. Yan, Inorg. Chem. 46 (2007) 3450;
(c) X.-H. Zhou, Y.-H. Peng, X.-D. Du, C.-F. Wang, J.-L. Zuo, X.-Z. You, Cryst.
azol-1-yl)acetic acid as anion for recycling of trivalent f-elements Growth Des. 9 (2009) 1028.
in used nuclear fuel. [11] X.-H. Zhou, Y.-H. Peng, Z.-G. Gu, J.-L. Zuo, X.-Z. You, Inorg. Chim. Acta 362
(2009) 3447.
[12] (a) D.P. Fagnant, G.S. Goff, B.L. Scott, W. Runde, J.F. Brennecke, Inorg. Chem. 52
Acknowledgements (2013) 549;
(b) X.-Y. Chen, G.S. Goff, M. Quiroz-Guzman, D.P. Fagnant, J.F. Brennecke, B.L.
The authors gratefully acknowledge the Laboratory-Directed Scott, W. Runde, Chem. Commun. 49 (2013) 1903.
[13] (a) J. Caja, T.D.J. Dunstan, V. Katovic, Proc. Electrochem. Soc. 19 (2002) 1014;
Research and Development Program and the G.T. Seaborg Institute (b) Y. Abu-Lebdeh, A. Abouimrane, P.-J. Alarco, M. Armand, J. Power Sources
for Transactinium Science at Los Alamos National Laboratory for 154 (2006) 255;
financial support during this project. M. Chai, Y. Jin, S. Fang, L. Yang, S.-i. Hirano, K. Tachibana, Electrochim. Acta 66
(2012) 67.
[14] A.N. Boa, J.D. Crane, R.M. Kowalczyk, N.H. Sultana, Eur. J. Inorg. Chem. (2005)
Appendix A. Supplementary data 872.
[15] X.-Y. Chen, G.S. Goff, B.L. Scott, M.T. Janicke, W. Runde, Inorg. Chem. 52 (2013)
3217.
CCDC 943459–943466 contain the supplementary crystallo- [16] APEX II, Bruker AXS. Inc., Madison, 2004.
graphic data for complex 1–8. These data can be obtained free of [17] SAINT+, Bruker AXS. Inc., Madison, 2003.

charge via http://www.ccdc.cam.ac.uk/conts/retrieving.html, or [18] G. Sheldrick, SADABS, University of Gottingen, Gottingen, Germany, 2001.
[19] SHELXTL; Bruker AXS, Inc., Madison, WI, 1997.
from the Cambridge Crystallographic Data Centre, 12 Union Road, [20] T. Moeller, The Chemistry of the Lanthanides, Pergamon Press, Oxford, 1973.
Cambridge CB2 1EZ, UK; fax: +44 1223-336-033; or e-mail: depos- [21] R.D. Shannon, Acta Crystallogr., Sect. A 32 (1976) 751.
it@ccdc.cam.ac.uk. Supplementary data associated with this article [22] S.N. Udaya, K. Chaitanya, M.V.S. Prasad, V. Veeraiah, A. Veeraiah, J. Mol. Struct.
1019 (2012) 68.
can be found, in the online version, at http://dx.doi.org/10.1016/
[23] I. Kostova, V.K. Rastogi, W. Kiefer, J. Optoelectron. Adv. Mater. 10 (2008) 1345.
j.poly.2013.10.015. [24] N. Sundaraganesan, E. Kavitha, S. Sebastian, J.P. Cornard, M. Martel,
Spectrochim. Acta, Part A 74A (2009) 788.
[25] (a) W.T. Carnall, The absorption and fluoresence spectra of rare earth ions in
References
solution, in: K.A. Gschneider Jr., L. Eyring (Eds.), Handbook on the Chemistry
and Physics of the Rare Earths, North-Holland Publishing Company, 1979, p.
[1] M.J. Alam, M.J. Ahsan, O. Alam, S.A. Khan, Lett. Drug. Des. Discov. 10 (2013) 171. Chapter 24;
776. (b) W.T. Carnall, P.R. Fields, J. Rajnak, J. Chem. Phys. 48 (1968) 4424.
[2] S. Mor, R. Mohil, D. Kumar, M. Ahuja, Med. Chem. Res. 21 (2012) 3541.
[3] (a) A.G. Yadav, V.N. Patil, A.L. Asrondkar, A.A. Naik, P.V. Ansulkar, A.S. Bobade,
A.S. Chowdhary, Rasayan J. Chem. 5 (2012) 117;

You might also like