Abstract
Recent advances in the coordination chemistry of Eu2+ are reviewed. Common synthetic routes for generating discrete Eu2+-containing complexes reported since 2000 are summarized, followed by a description of the reactivity of these complexes and their applications in reduction chemistry, polymerization, luminescence, and as contrast agents for magnetic resonance imaging. Rapid development of the coordination chemistry of Eu2+ has led to an upsurge in the utilization of Eu2+-containing complexes in synthetic chemistry, materials science, and medicine.
Keywords: Lanthanides, Coordination chemistry, Imaging agents, Luminescence, Polymerization
Introduction
Properties of Eu2+
Among the divalent lanthanides, Eu2+ has the most accessible divalent oxidation state because of its half-filled 4f7 electronic configuration and, consequently, a high stabilization from exchange energy (Figure 1).[1] While Eu2+ is the most accessible of the divalent lanthanides, outside of the solid state it has a propensity to oxidize [Eu0(Eu3+/Eu2+) = −0.35 V vs. the normal hydrogen electrode (NHE)],[2] which necessitates handling under an inert atmosphere. Four decades ago, there were only a few discrete Eu2+-containing complexes reported. Most of these complexes were halides, chalcogenides, and organometallic compounds that are generally insoluble in organic solvents such as tetrahydrofuran (THF) because of the formation of extended structures.[3]
Figure 1.
Relative reduction potentials for Ln3+→Ln2+ (V vs. NHE); values from ref.[2]
Despite oxidation and oligomerization, which present obstacles to the preparation and characterization of Eu2+-containing complexes, the interesting catalytic, photophysical, and magnetic properties of this ion have spurred a great deal of research. The unique properties of Eu2+ are influenced largely by the spacing of the energy levels of the 4f and 5d orbitals and the reducing properties of the ion.
The f orbitals make the electronic properties of Eu2+ unique in comparison to those of elements in the d block. For example, using spectroscopic techniques, Adin and Sykes demonstrated that electron transfer from f orbitals is more difficult than that from d orbitals.[4] They observed a rate constant for the reaction of Eu2+ with V3+ of 0.013 M−1s−1 (1.0 3 HClO4, 25 °C, ionic strength 2.0 M), while the rate constant for the reaction of Cr2+ with V3+ under the same conditions is 0.85 M−1s−1.[4] The smaller rate constant observed for Eu2+ relative to that for Cr2+ can be attributed to the shielding of the valence 4f orbitals from the environment by the electrons in the fully occupied 5s and 5p orbitals. An additional influence of the f orbitals is exemplified by the photophysical properties of Eu2+ that stem from the lowest-energy and first-excited-state configurations of 4f7 and 4f65d1, respectively.[5] While the 4f orbitals of Eu2+ remain largely unperturbed by the presence of ligands, the energy of the 5d orbitals is influenced readily by ligands. Consequently, the luminescence properties of Eu2+ can be tuned by using coordination chemistry, and the characteristic emission properties of this divalent ion include a broad emission band (390–580 nm)[6] and a short radiative lifetime (ca. 1 μs) that are attributed to the Laporte-allowed 4f65d1→4f7 transitions.[5] In addition to these allowed transitions, sharp emission bands, which appear between 354 and 376 nm[7,8] and have longer radiative emission lifetimes (ca. 1 ms) that correspond to the Laporte-forbidden 4f→4f transitions are observed, similar to those observed in the Eu3+ ion.[5] Beyond the unique luminescence properties of Eu2+, the f orbitals are responsible for the interesting magnetic characteristics of this divalent ion that include a high magnetic moment (7.63–8.43 μB) associated with seven unpaired electrons in an 8S7/2 ground-state configuration.[3,9]
In addition to the desirable optical and magnetic properties of Eu2+, this ion displays interesting redox chemistry. Divalent lanthanides including Eu2+ act as one-electron reductants, and detailed discussions of the reductive chemistry of divalent lanthanides were published by Evans in 2000 and 2002.[2,10] Since then, research efforts have been directed toward developing Eu2+-containing species that act as multielectron reductants in synthetic chemistry.[11] Additionally, there is interest in studying the influence of ligands on the redox properties of Eu2+.[12]
Early Eu2+ Complexes
Discrete Eu2+-containing complexes were reported as early as 1964.[13] However, the number of these complexes is small relative to their Eu3+ analogues, which form air-stable complexes with most electronegative atoms such as oxygen and nitrogen, partially because Eu3+ is a hard Lewis acid. Eu2+ is a softer Lewis acid than Eu3+ because of its lower charge density; consequently, Eu2+-containing complexes often include ligands with relatively soft atoms such as carbon and phosphorus.
Among the first discrete Eu2+-containing complexes prepared was the metallocene-like complex Cp2Eu, where Cp is cyclopentadiene. This complex did not garner much interest because of its insolubility in polar organic solvents such as THF.[14] This insolubility was ascribed to a polymerization that can be prevented with the use of bulkier ligands.[9,14,15] Templeton and co-workers were able to isolate crystals from THF or toluene of Cp*2Eu (1) in which Cp* is methyl-substituted Cp (Figure 2).[9] Evans and co-workers investigated the bond characteristics and the electronic spin–spin paramagnetic relaxation rate of Eu in Cp*2Eu(THF) (2) and Cp*2Eu(THF)(Et2O) (3). The Eu–ligand bonds of 2 and 3 are ionic in character as indicated by the isomer shifts of these complexes: 151Eu Mössbauer spectroscopy, a technique used to study the oxidation state and the local environment of Eu in the solid state, provides the information that the shifts of 2 and 3 are not different from those of ionic Eu2+ halide complexes.[16] Moreover, the data obtained from the spherical relaxation model fits of the 151Eu Mössbauer spectra of 2, 3, and [Cp*Eu(THF)2(μ-I)]2 (4) in the solid state indicate that the Eu–Eu distances affect the electronic spin–spin paramagnetic relaxation of these complexes: longer Eu–Eu distances lead to slower spin–spin relaxation rates.[16] In addition to Cp*, the sterically demanding ligand bis(trimethylsilyl)amide, [{(CH3)3Si}2N]−, was used as a precursor in preparing Eu2+ complexes. The six-coordinate complex [{(CH3)3Si}2N]2Eu(CH3OCH2CH2-OCH3)2 (5) and the four-coordinate complex [{(CH3)3Si}2-N]2Eu(bpy) (6), where bpy is 2,2′-bipyridine, are both soluble in THF, pentane, toluene, and 1,2-dimethoxyethane (DME).[3] In general, the problem of insolubility in hydrocarbons is addressed by using sterically demanding ligands to increase hydrophobicity and to prevent the formation of coordination polymers.[9,14,15]
Figure 2.
Structures of Eu2+-containing complexes 1–15.
Phosphane complexes of Eu2+ have been investigated since the early 1980s and display unique structural and electronic properties.[17–19] In some cases, phosphorus-containing ligands form dative bonds with Eu2+ in the presence of oxygen-containing bases.[20] Templeton and co-workers reported the synthesis of phosphane complexes of Eu2+: in their procedure, NaEu[N{Si(CH3)3}2]3 was prepared from EuI2 and NaN[Si(CH3)3]2 in diethyl ether, and the relatively weak donor ligand [N{Si(CH3)3}2]− was displaced by a phosphorus-containing ligand such as 1,2-bis(dimethyl-phosphanyl)ethane (DMPE) to give a complex with an empirical formula of Eu[N{Si(CH3)3}2]2(DMPE)1.5.[19] Another Eu2+-containing phosphane complex was reported by the same group with Cp* in place of [N{Si(CH3)3}2]−. The resulting complex, EuCp*2(DMPE) (7), was not soluble in noncoordinating solvents, potentially because of the role of DMPE in bridging complexes to form coordination polymers. To reduce the potential for aggregation, ethylene-bridged DMPE was replaced with methylene-bridged 1,2-bis(dimethylphosphanyl)methane (DMPM).[21] As a result, complex EuCp*2(DMPM) (8) was isolated, but analogous complexes with monodentate phosphanes such as P(CH3)3 or P(nBu)3 were not reported, likely because the steric bulk of these phosphanes prevented the formation of a coordinatively saturated metal ion.[21] Another Eu2+-containing phosphane complex was reported by Rabe and coworkers.[18] Their four-coordinate, homoleptic Eu2+-containing phosphane complex Eu[{μ-P(tBu)2}2–Li(THF)]2 (9) assumes a heavily distorted tetrahedral geometry. The oxidation state of Eu in the complex was confirmed by 151Eu Mössbauer spectroscopy with an isomer shift of −11.7 mms−1 as opposed to isomer shifts of 0–1 mms−1 for compounds that contain Eu3+ (Figure 3).[18] One interesting feature of this complex is that an alkali ion, Li+, interacts with the phosphorus in P(tBu)2 and the oxygen in THF in a similar way to the structure of complex NaEu[N{Si-(CH3)3}2]3, where Na+ interacts with the nitrogen atom of [N{Si(CH3)3}2]−.[15,18] Also, an agostic interaction between Eu2+ and the tert-butyl groups in P(tBu)2 was observed in 9.[18] The authors postulated that this interaction is driven by the tendency of the metal to form a coordinatively saturated environment and also by crystal packing forces because this interaction is only observed in the solid state.[15,18]
Figure 3.
151Eu Mössbauer spectrum of 9 at 4.2 K. Reprinted with permission from ref.[18] Copyright (1997) American Chemical Society.
In addition to the softer C- and P-containing ligands, polyoxadiazamacrobicyclic ligands (cryptands) form stable complexes (cryptates) of Eu2+ (complexes 10–13), which were first reported in the late 1970s. Cryptands are used to encapsulate Eu2+ to oxidatively stabilize this ion and enable studies of the luminescence properties of Eu2+. Gansow and co-workers demonstrated the oxidative stabilization of Eu2+, by using cyclic voltammetry, when the ion is bound by cryptands.[22] This stabilization was attributed to a good fit of the Eu2+ ion into the cryptand cavity,[22] and the thermodynamic stability constants (log K) for some Eu2+-containing cryptates were reported by Burns and Baes to be 10.2–13.0.[23] Besides the thermodynamic properties of these compounds, these Eu2+-containing cryptates are substitutionally inert towards the Na+, Ba2+, Ca2+, Mg2+, Zn2+, and tetraethylammonium cations.[24,25] Beyond stabilizing the Eu2+ ion thermodynamically and kinetically, cryptands also exclude a large number of solvent molecules from the inner sphere of the ion. This solvent exclusion is important because coordinated solvent molecules quench the luminescence of Eu2+, thus this ion does not exhibit strong luminescence in aqueous solution at room temperature when not chelated.[26]
In the 1990s, Eu2+ complexes with a wide variety of ligands were synthesized. Shore and co-workers reported Eu2+-containing borohydrides (NH3)xEu(B10H14), (CH3CN)2Eu(BH4)2, and (C5H5N)1.8Eu(BH4)2 as precursors for preparing metal borides.[20,27] However, all complexes that contain borohydrides are unstable as solids and tend to decompose to produce closo-[B10H10]2−.[20] The conversion from Eu2+ borohydrides to Eu2+ borides was performed successfully at high temperatures (> 200 °C) under vacuum to obtain the stable boride phase EuB6.[27] Although Eu2+-containing complexes with borohydrides as ligands were successfully synthesized, the coordination modes of Eu2+ in these complexes were not fully explored because of the lack of X-ray crystallographic data. In addition to Eu2+ borohydrides, Eu2+ selenolates and tellurolates were also prepared. For example, Eu[ESi{Si(CH3)3}3]2-(TMEDA)2 (14) and [Eu{ESi{Si(CH3)3}3}2(DMPE)2]2(μ-DMPE) (15) were reported by Arnold and co-workers; in these compounds TMEDA is (CH3)2NCH2CH2N(CH3)2 and E is Se or Te.[28] The crystal structure of 15 reveals a seven-coordinate Eu2+ ion that is bridged by DMPE to form a dimeric complex that was not observed for the TMEDA-containing complex. The difference in coordination chemistry of the two complexes was attributed to the larger covalent radius of P relative to N and, consequently, a less crowded metal center in complex 15.[28]
The seminal studies of the coordination chemistry of Eu2+ from the 1960s through the 1990s built a foundation for the design of new ligands that will be driven by the functions of the resulting Eu2+-containing complexes. While Eu2+-containing clusters and polymers are reported in the literature,[29–33] these compounds are not described in this review, because the focus of this review is recent advances in the coordination chemistry of discrete Eu2+ complexes. Synthetic strategies reported in the 21st century are presented, followed by a description of the reactivity and applications of the resulting Eu2+-containing complexes.
Synthesis of Recent Eu2+-Containing Complexes
The rapid increase in the reports on Eu2+-containing complexes over the last two decades has resulted in the publication of several synthetic pathways to prepare these complexes (Scheme 1). These methods fall into three categories regarding the europium-based starting materials used: (1) metallic europium; (2) europium(III) salts including halides, triflates, oxides, or nitrates; (3) europium(II) halides. This section is organized on the basis of these three types of starting materials: oxidation of Eu metal, reduction of Eu3+, and metathesis of Eu2+.
Scheme 1.
Synthetic methods for obtaining Eu2+-containing complexes arranged by source of europium where L is a ligand, N is either Hg or Sn, and M is an alkali metal ion.
Oxidation of Eu0
Oxidation of Eu metal can be accomplished by activation of Eu0 with Hg, iodine, or liquid ammonia, by redox trans-metallation, or by pseudo-Grignard-compound formation. It is crucial to perform this type of synthesis under an inert atmosphere, because Eu metal will oxidize uncontrollably in air to produce mixtures of Eu2+- and Eu3+-containing complexes.
Syntheses of Eu2+ complexes from Eu0 often require activation. Mercury, which activates Eu0 through amalgamation, was used to prepare the homoleptic and tetranuclear Eu2+ complex Eu4[(tBu)2pz]8 (16) in which (tBu)2pz is 3,5-di-tert-butylpyrazolate (Figure 4).[34] Structural elucidation of linear complex 16 revealed different coordination modes (η2, μ-η5:η2, and μ-η2:η2) of the pyrazolate groups to Eu2+. The outer Eu2+ ions are bonded in a η2 fashion by the two terminal pyrazolate groups; a μ-η5:η2 coordination mode is observed for the four pyrazolate groups that bridge the inner and outer Eu2+ ions; and a μ-η2:η2 binding mode is exhibited by the two pyrazolate groups that bridge the two inner Eu2+ centers.[34]
Figure 4.
Structures of Eu2+-containing complexes 16–19, 21–28, 30–35, 38, and 40, which were prepared by the oxidation of Eu0.
An alternative to Hg as an activating agent is iodine, which was reported to be crucial in the preparation of several lanthanide-based complexes including the dimeric [Eu(Odip)(μ-Odip)(THF)2]2 (17) in which Odip is 2,6-diisopropylphenolate.[35] In addition to mercury and iodine, dissolution in liquid ammonia can be used in the activation of Eu0. Such is the case in the preparation of Eu2+-containing alkoxide Eu[OC6H3-2,6-(tBu)2]2(THF)3·0.75C7H8 (18) and aryloxides [Eu4(μ-OC10H7)6(OC10H7)2(THF)10] ·2THF (19) and Eu(OC6H4OCH3)2 (20).[36–38] Activation of Eu0 by using liquid ammonia was also employed for the synthesis of [Eu2{(tBu)2pz}4(THF)2], which is characterized by centrosymmetric Eu2+ centers bonded in μ-η2:η5 and η2 modes by the bridging and terminal pyrazolate ligands, respectively.[39] These coordination modes can also be found in pyrazolate-based Eu2+ complex 16 and demonstrate the ability of pyrazolate ligands to coordinatively saturate Eu2+. Liquid ammonia was also used in the one-pot synthesis of an organohydroborate complex of Eu2+, [(THF)4Eu-{(μ-H)2BC8H14}2] (21), which features an agostic interaction between Eu2+ and a hydrogen atom in one of the hydroborate ligands. Structural and IR spectroscopic data suggest the presence of agostic interactions in solution and in the solid state, and these interactions may have influenced the cis arrangement of the hydroborate ligands in the octahedral complex.[40]
In the preparation of Eu2+ complexes through redox transmetallation, mercury- and tin-based complexes are sometimes used as oxidants, and activating agents such as Hg are used to improve the yields of this type of synthesis. The preparation of the (perfluoroaryl)europium(II) complex Eu(C6F5)2(OC4H8)5 (22) entails the oxidation of Eu0 by using bis(pentafluorophenyl)mercury, Hg(C6F5)2, in THF. Complex 22, characterized by a pentagonal bipyramidal geometry with axial C6F5 groups and five THF molecules in the equatorial position, has relatively long Eu–C bonds (2.82 Å); this is likely due to the inductive electron-withdrawing effect of the fluorine substituents of the two phenyl groups.[41] The same oxidant, Hg(C6F5)2, was used to prepare the homoleptic and dinuclear {Eu(iPr2pz)2-(iPr2pzH)2}2 (23) in which iPr2pz is 3,5-diisopropylpyrazolate. An X-ray crystal structure of pyrazolato Eu2+ complex 23 shows a different coordination mode (μ-η5:η1) to Eu2+ by the bridging 3,5-diisopropylpyrazolate centers. This coordination mode of the bridging pyrazolate groups in addition to the η1 binding mode of the terminal pyrazolate groups was ascribed to intramolecular hydrogen binding.[42] It is worth noting that redox transmetallation reactions carried out in nonpolar solvents including toluene (for example, in the preparation of complex 23) can produce homoleptic Eu2+ complexes. In addition to complexes 22 and 23, Hg(C6F5)2 was used as oxidant in the preparation of Eu2+-containing complexes with tri-azanides as ligands [Dmp(Tph)N3EuC6F5] (24) and [Eu{N3Dmp(Tph)}2] (25) in which Dmp is 2,6-[2,4,6-(CH3)3C6H2]2C6H3 and Tph is 2-(2,4,6-iPr3C6H2)C6H4. Of particular note are the coordination modes of the ter-phenyl and biphenyl groups in complexes 24 and 25 that exhibit π-arene–Eu interactions with Eu–C distances of 3.088–3.233 and 3.011–3.311 Å in complex 24 and 25, respectively. In complex 24, both the mesityl ring of the terphenyl group and the triisopropylphenyl ring of the biphenyl moiety show an η5 interaction with the metal, whereas in complex 25, the mesityl rings of terphenyl groups exhibit η5 and η3 binding modes. The assignment of hapticity was based on the shortest metal–centroid separation or the smallest angle between the M–centroid vector and the normal of the arene plane.[43,44]
When Hg(SCN)2 was used as oxidant, [Eu(NCS)2-(DME)3] (26) was obtained from redox transmetallation with Eu0 in DME, and bimetallic [Eu(NCS)2(THF)4]2 (27) was obtained when THF was used as solvent. The crystal structure of the eight-coordinate 26 shows configurational isomers, and the relative amount of each isomer is dependent on temperature. Racemic mixtures in a single crystal can be obtained at lower temperature (5 °C), while a mixture of enantiomerically pure Λ (left-hand) or Δ (right-hand) products was observed in a single crystal at room temperature.[45] The coordination environment in 27 features two bridging thiocyanate ligands and terminal THF molecules, adapting a distorted pentagonal bipyramidal geometry at each Eu2+ center.
Besides mercury-based oxidants, trialkyltin(IV) compounds were explored for redox transmetallation with Eu0. Either the trivalent [Eu3+(Ph2pz)3(DME)2]·2DME (28) or the divalent cis-[Eu2+(Ph2pz)2(DME)2] (29) can be made from the reaction of Sn(CH3)3(Ph2pz), where Ph2pz is 3,5-diphenylpyrazolate, with Eu0 in DME by changing the amount of excess Eu0.[46,47]
In contrast to redox transmetallation, the route toward the formation of a pseudo-Grignard compound has been successful in preparing Eu2+ complexes even without the use of Hg, iodine, or ammonia. Thiolate-containing complex Eu(SAr*)2(THF)0.5 (30), in which Ar* is 2,6-[2,4,6-(iPr)3C6H2]2C6H3, was prepared from Eu0, Ar*SH, and 2-iodobenzotrifluoride.[48] This thiolate complex is characterized by η6-π interactions of Eu2+ with two o-2,4,6-triisopropylphenyl rings of the terphenyl groups, as evidenced by IR spectroscopy and X-ray crystallography.[48] Another example of a Eu2+-containing complex that shows η1-π interactions with Eu2+ and was prepared from europium metal is Eu(Dpp)(THF)2 (31), where Dpp is 2,6-Ph2C6H3. Complex 31 was synthesized from Eu0 and DppI in THF and exhibits a distorted tetrahedral geometry with two aryl ligands and two THF molecules.[49]
While the oxidation of Eu0 through activation of the metal surface, redox transmetallation, and formation of pseudo-Grignard compounds is used to prepare homo-metallic Eu2+ complexes, charge-separated and neutral heterobimetallic Eu2+-containing complexes can also be synthesized from Eu0. Treatment of Eu0 and another rare earth metal (Y0, Nd0, or Ho0) with 2,6-diphenylphenol (OHdpp) at elevated temperature (190 °C) in the presence of Hg yielded charge-separated complexes [Eu2(Odpp)3]-[Y(Oddp)4] (32), [Eu2(Odpp)3][Nd(Oddp)4] (33), and [Eu2-(Odpp)3][Ho(Oddp)4] (34). The Eu2+ centers in [Eu2-(Odpp)3]+ are bridged by aryloxide ligands and have coordination modes of η4:η2:η1 and η1:η6:η3 with π-arene–Eu interactions.[50] In contrast to the charge-separated complexes, heterobimetallic complexes [CaEu(Oddp4)] (35), [SrEu(Oddp)4] (36), and [BaEu(Oddp)4] (37) were produced upon reaction of Eu0 and an alkaline-earth metal (Ca0, Sr0, or Ba0) with OHdpp at high temperatures (210–235 °C). However, when Eu0 was treated with HOdpp in the presence of MOdpp (M = Na0, K0, and Li0) at elevated temperature (210 °C), heterobimetallic Eu2+-containing complexes [NaEu(Oddp)3]·PhCH3 (38) and [KEu(Oddp)3]·2.5PhCH3 (39) and the mixed-valent complex [Eu2(Oddp)3]-[Eu(Oddp)4] (40) were obtained. These heterobimetallic complexes also show π-arene–Eu interactions that coordinatively saturate Eu2+. The coordination mode of the Eu2+ center is similar in 35 and 36 (η2:η3), while a different binding mode was observed in 37 (η3:η5). The coordination modes in the Eu2+-containing complexes with alkaline metals are η4:η2:η1 (complex 38) and η3:η1:η1 (complex 39) and are similar to that of [Eu2(Odpp)3]+ in charge-separated complexes 32–34.[51] Structural data for the [Eu2(Oddp)3]+ unit in 40 depict aryloxides binding in η1:η4:η2 and η3:η2:η3 fashions.[51] Purification of charge-separated or heterobimetallic complexes requires solvent extraction to remove the unreacted metal and homometallic coproducts.[50,51]
Despite the propensity of Eu metal to undergo uncontrolled oxidation to both Eu2+ and Eu3+, the use of Eu metal can produce modest to excellent yields (23–96%) in the preparation of Eu2+-containing complexes. Also, activating Eu0 with Hg has been shown to increase the yield of the desired Eu2+ complexes.
Reduction of Eu3+
The reduction of Eu3+ by either a ligand that is also a reductant or by chemical or electrochemical methods is more common than the oxidation of Eu metal. Eu3+ halides, triflates, oxides, or nitrates are typically used as the precursors for the preparation of Eu2+-containing complexes with a variety of ligands including aryloxides, amino-diboranates, crown ethers, and cryptands.[25,52–58] An example of a Eu2+-containing complex that was prepared from Eu3+ chlorides with a reducing species is the dinuclear, distorted pentagonal bipyramidal complex [Eu{H3BN-(CH3)2BH3}2(THF)2]2 (41), which was obtained by using N,N-dimethylaminodiboranate as the reductant (Figure 5).[52] However, the preparation of 41 produced mixtures of divalent and trivalent species.[52] In contrast, only divalent species were obtained when EuI2 was used as a precursor instead of EuCl3.[52] Qi and co-workers used Eu3+ chlorides as starting materials and a Na–K alloy as reductant to prepare [LEu(HMPA)2]2(THF)4 (42), in which L is 2,2′-methylene bis(6-tert-butyl-4-methylphenoxo) and HMPA is hexamethylphosphoric triamide. Complex 42 is characterized by two phenolate moieties bridging two Eu2+ atoms.[53] Each Eu2+ in 42 adopts a distorted trigonal bipyramidal geometry and a slightly asymmetric bridging mode of the phenolate ligand. This asymmetry is also observed in the bond length of Eu–O: the axial Eu–O bond (2.44 Å) is shorter than the equatorial Eu–O bond (2.52 Å).[53] Also, the variation of the Eu–O–C angles in 42 have been associated with the existence of agostic interactions in this coordinatively unsaturated complex.[53] In addition to the use of Na–K alloy as a noncoordinating reductant in the preparation of Eu2+-containing complexes, Zn was used as reductant in the synthesis of complexes 10 and Eu2+–18-crown-6 (43) from Eu3+ triflate.[54] The reduction of Eu3+ by Zn in the presence of 18-crown-6 or [2.2.2]cryptand was monitored by using the absorption bands at 250 and 330 nm, which were assigned to the Eu2+ species.[54] Additionally, Eu3+ oxides serve as starting materials for preparing Eu2+-containing complexes by electrochemical reduction. For example, Eu2O3 was reduced electrochemically in the syntheses of the Eu2+ complex of benzo-18-crown-6 and of triethanolamine.[55,56] With the strategy of reducing Eu3+, our laboratory prepared several Eu2+-containing cryptates: 10,11, and 44–47.[57] In general, the advantage of using the reduction approach is that the metal source is air-stable, so the problem of surface oxidation is avoided. Nevertheless, the possibility of incomplete reduction of Eu3+ can limit the yield and purity of the desired Eu2+-containing complexes.
Figure 5.
Structures of Eu2+-containing complexes 41–47, which were synthesized by the reduction of Eu3+.
Metathesis of Eu2+ Complexes
While the other synthetic methods involve redox reactions to obtain Eu2+, the metathesis approach makes use of Eu2+ halides and alkali-metal-containing ligands as precursors to generate the desired Eu2+-containing complex. The metathesis reaction has been successful in preparing several Eu2+-containing complexes including the bimetallic [Eu2- (Ap*py)3I(THF)] (48), where Ap*py is deprotonated 6-methylpyridin-2-yl-[6-(2,4,6-triisopropylphenyl)pyridin-2-yl]amine,[59] a monometallic Eu2+-containing trans-N,N′-di-methyl-meso-octaethylporphyrinogen complex (49),[60] a Eu2+-containing benzamidinate complex [Eu{PhC-{NSi(CH3)3}{2,6-(iPr)2NC6H3}}2(THF)2] (50),[61] a Eu2+ complex with olefin-substituted cyclopentadienyl ligands [{C5(CH3)4Si(CH3)2CH2CH=CH2}2Eu] (51),[62] bis(diphosphanylamido) complex [{(Ph2P)2N}2Eu(THF)3] (52),[63] a Eu2+ complex containing bis(phosphinimino)methanide ligands [{{(CH3)3SiNPPh2}2CH}EuI(THF)]2 (53),[64] an aminotroponiminate complex of composition [{(iPr)2ATI}-Eu{N{Si(CH3)2}}(THF)2] (54), where (iPr)2ATI is N-isopropyl-2-(isopropylamino)troponiminate,[65] and a monometallic, heteroleptic complex [(DIP2pyr)Eu(THF)3] (55), where DIP2pyr is 2,5-bis[(N-2,6-diisopropylphenyl)iminomethyl]pyrrolyl (Figure 6).[66] Complexes 48–55 were prepared by the metathesis reaction of [EuI2(THF)x], where labile THF, iodine, or both are displaced by the desired ligand. Complex 50 exhibits cisoid and transoid isomers with respect to the position of coordinated THF.[61] The bidentate diphosphanylamides, (Ph2P)2N−, in complex 52 are η2 coordinated to the Eu2+ center through the phosphorus and nitrogen atoms. The structure of the bimetallic Eu2+-containing complex 53 displays the tridentate bis(phosphinimino)methanide ligands bound to Eu2+ through two nitrogen atoms and a methine carbon atom. Two of these ligands are coordinated to a Eu2+ center to form a six-membered metallacycle with a twist boat conformation.[64] When complex 53 was treated with K{CH(PPh2NSi(CH3)2}, the six-coordinate amido Eu2+-containing complex [{{(CH3)3-SiNPPh2}2CH}Eu{(Ph2P)2N}(THF)] (56) was obtained. As observed in 53, the methine carbon atom was coordinated to Eu2+ in 56, and the reported Eu–C (methine carbon) bond lengths were 2.945 Å and 2.878 Å for complexes 53 and 56, respectively.[64] In contrast to 53, in which iodine bridged two Eu2+ centers, complex 56 is monometallic; a nonbridging iodine atom and THF occupy the axial positions and the tridentate (DIP2pyr)− ligand is bound to Eu2+ through its three nitrogen atoms to afford a distorted pentagonal pyramidal complex. As an alternative to the metathesis reaction of [EuI2(THF)x] in preparing Eu2+-containing complexes, we used the commercially available EuI2 to prepare cryptates 10, 11, and 57–59.[25,58] A similar europium halide salt was used to generate four-coordinate [Eu{(ArN)2CN(iPr)2}2] (60), where Ar is 2,6-diisopropylphenyl.[67] Roesky and co-workers used a bimetallic Eu2+-containing complex [{CH(PPh2NSiMe3)2}Eu(THF)(μ-I)]2 and NaBPh4 as precursors to prepare [{CH{PPh2NSi-(CH3)3}2}Eu(THF)3]BPh4 (61).[68] The resulting complex was used to prepare other Eu2+-containing complexes by metathesis of the labile borate ligand.[68] Another example of a borate-containing Eu2+ complex, [Cp*(μ-η6:η1-Ph)2-EuBPh2] (62), was prepared from {Eu(Cp*)2} in the presence of [Et3NH][BPh4]. The structural motif of complex 62 features one coordinated Cp* and two of the phenyl rings of BPh4− ligand in a μ-η6:η1 binding mode, with the ligands in a pyramidal geometry around the coordination sphere of Eu2+. While density functional theory (DFT) calculations predict a trigonal planar shape for complex 62, the large size of Eu2+ prefers intermolecular packing over intramolecular packing in the solid state, and agostic interactions may also have contributed to this deviation of geometry.[69] In general, with the metathesis approach of preparing Eu2+-containing complexes, the presence of other oxidation states of Eu is not a problem as long as synthesis is performed under an inert atmosphere. The success of this third approach of metathesis relies on the completeness of the displacement of the alkali metal or a practical method of purification to remove byproduct salts. Steric limitations introduced by the incoming ligand may pose a problem in preparing the target Eu2+ complex.
Figure 6.
Structures of Eu2+-containing complexes 48–62, which were obtained from the metathesis of Eu2+-containing complexes.
Each of the three synthetic methods described here have distinct advantages and limitations, and each route must be carefully considered when choosing a method to synthesize a new complex. In addition, structural studies of the recently synthesized Eu2+-containing complexes presented in this section demonstrate the rich coordination chemistry of Eu2+. Differences in the coordination environment in Eu2+-containing complexes can be attributed to steric and electronic effects of the coordinated ligands as well as the size and electronic properties of Eu2+, which are responsible, in part, for the agostic interactions observed in some Eu2+ complexes.
The development of other Eu2+-containing complexes with use of other divalent lanthanides in generating complexes with Ln2+–Ga+ or Ln2+–Al+ (Ln = Sm, Eu, Yb, or Tm) bonds is an active area of research.[70–72] The applications of these compounds as reductants are currently being explored.[70]
Applications of Eu2+-Containing Complexes
The growing interest in Eu2+-containing complexes stems from their unique catalytic, photophysical, and magnetic properties. The remainder of this review describes the use of Eu2+-containing complexes in four different applications.
Reductants
Eu2+ is a one-electron reductant for many systems. However, it was reported that a bimetallic Sm2+-containing complex has the potential to be a two-electron reductant per molecule.[73] This discovery led to the pursuit of divalent lanthanide species, including Eu2+-containing complexes, that could act as multielectron reductants. This active area of research is important in establishing a more complete understanding in the reduction chemistry of Eu2+. Lee and co-workers reported Eu2+-containing complex 50 that serves as a one-electron reductant to diphenyldichalcogenide PhSeSePh to produce [Eu({PhC{NSi(CH3)3}{2,6-(iPr)2-NC6H3}})2(μ-SePh2)]2 (63), in which the two Eu3+ centers are bridged by diphenyl diselenides (Figure 7).[61] Each Eu3+ ion in 63 assumes a distorted octahedral geometry and is coordinated by two η2-bound benzamidinate ligands. However, when the same Eu2+-containing bisamidinate complex was allowed to react with diphenyl ditellurides and iodine, both reactions were unsuccessful, indicating that the reducing strength of Eu2+ was not sufficient to reduce diphenylditellurides and iodine.[61] Also, in the case of the macrocyclic ligand system trans-N,N′-dimethyl-meso-octa-ethylporphyrinogen, the Eu2+-containing complex cannot reduce tert-butyl-1,4-diazabuta-1,3-diene (tBu-DAB).[60] However, the analogous Sm2+ complex forms a bimetallic complex when reacted with tBu-DAB.[60] The formation of the bimetallic Sm3+ complex bridged by the reduced [tBu-DAB]2− is driven by the reducing power of Sm2+ towards tBu-DAB, but the steric constraints due to the bulky tert-butyl moieties makes this reaction reversible.[60] Because Eu2+ has a more positive Ln2+/Ln3+ oxidation potential (−0.35 V vs. NHE) than Sm2+ (−1.55 V vs. NHE), the driving force to form an analogous bimetallic Eu3+ complex is hindered, presumably because the reducing strength of Eu2+ is not sufficient to reduce tBu-DAB. These examples of controlled redox potential could be useful in selective reductions during multistep syntheses and for developing the first Eu2+-containing complex that could act as multielectron reductant.
Figure 7.
Structure of Eu2+-containing reductant 63.
An important factor that can control the reducing properties of Eu2+ is the nature of the ligands. As an example, the reaction with Cp*2Eu(OEt2) with C6F5-substituted diazabutadiene (DADC6F5)− led to the oxidation of Eu2+ and formation of the radical anion (DADC6F5)·−.[74] The presence of Eu3+ in the resulting complex was demonstrated by crystallographic data, magnetic susceptibility measurements, and IR spectroscopy. The electron-withdrawing characteristic of the C6F5 moiety and its proximity to the metal center is likely the cause of the facile oxidation of Eu2+. On the other hand, neither oxidation of Eu2+ nor formation of radical anion was observed when tert-butyl-substituted diazabutadiene was used.[74] In general, the diazabutadienyl ligand can have rich redox chemistry when coupled with Eu2+ if this ligand is modified by using substituents of different electronic properties.
Polymerization Initiators
As discussed in the previous section, ligands can influence the rich reduction chemistry of Eu2+, and the resulting complex can play an important role in several applications including polymer chemistry. In the preparation of polymers, Eu2+-containing complexes can act as electron-transfer agents to initiate several polymerization processes. An example of this application is the activity of several Eu2+-containing complexes that contain indenyl ligands in the initiation of the polymerization of methyl methacrylate (MMA) and ε-caprolactone.[11,12]
Methyl Methacrylate Polymerization
Eu2+-containing complexes 64–73, which contain functionalized indenyl ligands are used in the polymerization of MMA (Figure 8).[11,75–77] The activity of these complexes in polymerizations and the stereochemistries of the resulting polymers are dependent on temperature and solvent (Figure 7). High activity is observed at low temperatures (−30 to −60 °C for complexes 64 and 65; 0 to −60 °C for complexes 66–69, 72, and 73; −30 to −45 °C for complexes 70 and 71), and these complexes have solvent-dependent activities. Complexes 64, 65, and 72 have high activities when THF or dimethoxyethane is used as solvent instead of toluene. Complex 66 has high activity in dimethoxyethane, THF, and toluene; 67 shows good activity in THF; and complexes 68, 69, and 73 have high activity in dimethoxyethane. The molecular weights of the resulting MMA polymers are high (up to 443 kDa) at low temperatures, which indicates that chain propagation is favored at low temperatures.[11,75–77] The mechanism of MMA polymerization was postulated to be initiated through reductive dimerization of MMA by electron transfer from Eu2+, which results in the formation of Eu3+ enolates.[11,75–77] Propagation did not go through an insertion mechanism, because the indenyl ligand was not detected on the polymer chains by 1H NMR spectroscopy.[11,76]
Figure 8.
Structures of Eu2+-containing complexes 64–82, which were used as polymerization initiators.
The stereochemistry of the MMA polymers was determined by 1H NMR spectroscopy. The majority of the MMA polymers produced by using Eu2+-containing functionalized indenyl complexes 64, 65, 72, and 73 as initiators is syndiotactic.[11,77] However, when complexes 66–68 were used in dimethoxyethane or THF, a 1:1:1 ratio of isotactic, syndiotactic, and atactic polymers was obtained. However, predominantly syndiotactic polymers were obtained when using 66 in toluene, and a decrease in the ratio of syndiotactic to isotactic polymers was observed with 70 when the solvent was changed from THF or dimethoxyethane to toluene, which suggests that a solvent effect influences the stereochemistry of these polymers. The presence of several tacticities in the synthesized polymers was attributed to the rac/meso interconversion, which is favored by solvents including THF and dimethoxyethane.[77] While the stereochemistry of the polymers showed dependency on the solvent used, it is still unclear whether this stereoregularity is influenced by the nature of the initiator, specifically by the metal or ligands. A similar study that uses different initiators (for example, different divalent lanthanide metal ions or a ligand system other than indene compounds) could answer this question.
Ring-Opening Polymerization of ε-Caprolactone
Eu2+-containing complexes 74–79, which contain indenyl, fluorenyl, and iminopyrrolyl ligands, showed catalytic activity in the ring-opening polymerization of ε-caprolactone.[12,78–80] With these catalysts, high activity is observed at high temperatures (0 to 30 °C for complexes 75 and 76; −30 to 60 °C for complex 74; 0 to 60 °C for complex 77; and 30 to 60 °C for complexes 78 and 79), and the activities of these complexes depend on the solvent. Complexes 77–79 have high activities when THF or toluene is the solvent; complex 74 exhibits high activity in dimethoxyethane, THF, and toluene, and 75 and 76 show good activity in toluene. Furthermore, the steric bulk of the substituents in the ligands may have an effect on the activity of the complexes and the molecular weights of the resulting polymers. When the silyl substituent in the furfuryl- or tetrahydrofurfuryl-functionalized indenyl ligands of complexes 75 and 76 was absent, the two complexes were not active. However, it is not clear what the role of the silyl substituent is in the polymerization process.[78] The use of more sterically bulky ligands in 79 relative to the ligands in 78 has been implicated to affect the chain propagation process, thus affecting the molecular weights of the polymers.[80] Wang and co-workers proposed that the polymerization of ε-caprolactone is initiated by the formation of Eu3+-containing alkoxyl radical species formed from the oxidation of the Eu2+ metal center (Scheme 2).[12] The alkoxyl radical species is believed to take a second Eu2+-containing complex and produce a bimetallic Eu3+-containing complex that is bridged by the opened caprolactone.[12] Subsequently, chain propagation occurs when subsequent molecules of ε-caprolactone are added to the Eu3+ complex by a coordination–insertion process.[12]
Scheme 2.
Proposed mechanism of polymerization of ε-caprolactone.[12] Ind is the indenyl ligand. Adapted with permission from ref.[12] Copyright 2007 Wiley-VCH Verlag GmbH & Co. KGaA.
Because the role of the ligands on the polymerization of MMA and ε-caprolactone has yet to be elucidated, several Eu2+ complexes, [{η5:η1-{1-(CH3)3Si-3-(C5H9OCH2)-C9H5}}2Eu] (80),[81] [(η5:η1 -C5H9OCH2C13H8)2Eu] (81),[82] and [(η5:η1-C5H10NCH2CH2C13H8)2Eu] (82),[82] containing functionalized indenyl and fluorenyl ligands were investigated in terms of their activity towards MMA and ε-caprolactone polymerization. The silyl substituent in the indenyl ligands in complex 80 is likely a key factor in the two polymerization processes.[81] Without the silyl substituent in 80, no activity was observed under any reaction conditions.[81] However, high activity was observed with 80 for both MMA and ε-caprolactone polymerizations. The effect of the ligand is also evident in the case of complexes 81 and 82. Complex 81 with N-piperidineethyl-functionalized fluorenyl ligands has higher activity in ε-caprolactone polymerization than complex 81 with tetrahydro-2H-pyranyl ligands under the same reaction conditions. However, this ligand effect was not observed on the polymerization of MMA.
In general, Eu2+-containing complexes that have fluorenyl ligands are more effective as catalysts for both MMA and ε-caprolactone polymerizations than indenyl ligands with similar substituents. While Eu2+-containing complexes 64–82 initiate the polymerization of MMA or ε-caprolactone, it would be useful to investigate the mechanistic role of the different ligands in the initiation. Also, exploration of a broader ligand system would enable an in-depth understanding of how to control better the properties of the synthesized polymers. Although a Eu2+ complex has been used in the cyclotrimerization of isocyanates, the scope of the ligands used is still limited.[83] These ligand-centered studies would expand our understanding of the reduction chemistry of Eu2+ and its application to synthetic and polymer chemistry.
Luminescent Complexes
Aside from the rich reduction properties of Eu2+, this ion produces complexes that are luminescent both in solution and in the solid state. Current research efforts are directed towards developing Eu2+-containing complexes that have high luminescence efficiency.
The luminescence of uncomplexed Eu2+ and Eu3+ in protic solvents is quenched by the O–H oscillators of coordinated solvent molecules.[84] Thus, macrocyclic ligands including crown ethers and cryptands are used to encapsulate Eu2+ for protecting this ion from luminescence quenching by solvent molecules. The luminescence properties of Eu2+-containing crown ethers, azacrown ethers, cryptands, and polymers were reviewed by Adachi and co-workers in 1998.[6] Since then, efforts have focused on developing luminescent complexes with high quantum yields by introducing functional groups into macrocyclic ethers or by using other ligands that better encapsulate Eu2+.[85–88] There is also a growing interest in the comparison of spectroscopic properties of Eu2+ in the solid state to Eu2+-containing complexes in solution.
A variety of functional groups have been used to study the contribution of antenna effects and conjugated π systems on the luminescence efficiency of Eu2+-containing complexes.[86–87] Pyridine groups, known for exhibiting an antenna effect in Eu3+-containing complexes, were introduced into 18-crown-6. When bis(pyridino)-18-crown-6 was complexed with Eu2+, the emission of the complex in the solid state was quenched at room temperature, and only a weak emission at 430 nm was observed at 77 K, which is typical for Eu2+-containing crown ethers.[86] This quenching is likely due to the energy transfer from the excited-state, quasi-5d energy level of Eu2+ to the π* levels of pyridine, and this quenching was not observed when unfunctionalized 18-crown-6 ligand 83 was used.[86] However, a blue luminescence was observed when benzo-15-crown-5 84 and benzo-18-crown-6 85 were used as ligands.[54,87] The Eu2+-containing complex with benzo-15-crown-5 is characterized by a Eu2+ metal center that is sandwiched by two benzo-15-crown-5 ligands.[86] In the solid state, Eu2+–85 shows an emission peak that is similar to that observed in methanol at room temperature. This observation is also true for Eu2+-containing 15-crown-5 compound Eu–86. These observations suggest that similar structures exist in the solid state and in solution.[88] The mean lifetime of the excited state of Eu2+–84 at room temperature is 0.65 μs in the solid state and 0.14 μs in methanol (Table 1).[87] This difference in excited state lifetime is likely due to the additional relaxation processes caused by the methanol O–H oscillators.[87] Furthermore, a phenyl moiety in the crown ether might contribute to a reduction in the lifetime, because this moiety can rigidify the crown ether and lead to structural changes to make Eu2+ more exposed to luminescence quenchers relative to Eu2+ bound to unmodified crown ethers.[6] Also, the cavity size of the crown ether affects the luminescence lifetime of the excited state. Among the three crown ethers studied, the cavity size of 86 is most efficient in reducing the number of coordinating anions or solvent molecules that decrease luminescence lifetime.[88]
Table 1.
Luminescence properties of Eu2+-containing complexes in methanol solution and in the solid state (room temperature and 77 K).[6,54,87,88]
| Complex | Maximum emission band (nm) (solid) | Maximum emission band (nm) (methanol) | Luminescence lifetime (μs) (solid) | Luminescence lifetime (μs) (methanol) |
|---|---|---|---|---|
| Eu2+–83 | ND (r.t.[a]), 411.5 (77 K) | ND[a] | ND, ND | ND |
| Eu2+–84 | 422 (r.t.), 427 (77 K) | 417 | 0.65 (r.t.), 0.59 (77 K) | 0.14 |
| Eu2+–85 | 425 (r.t.), 430 (77 K) | 447 | 8.34 (r.t.), 8.01 (77 K) | 0.028 |
| Eu2+–86 | 433 (r.t.), 417 (77 K) | 432 | 0.922 (r.t.), 0.745 (77 K) | 0.800 |
Legend: r.t. is room temperature, ND is no data available.
The search for ligands that could provide more complete encapsulation remains an active area of research in investigating the luminescence properties of Eu2+. Along these lines, a nonmacrocyclic ligand, nitrotriacetate [N(CH2COO)33−], was tested for its luminescence properties when chelated to Eu2+.[85] An emission at 483 nm at 77 K was observed for the Eu2+-containing complex with nitro-triacetate in the solid state, and this emission is in the wavelength range observed for Eu2+-containing complexes of crown ethers that contain tertiary amines.[85] The luminescence lifetime was shorter (< 0.1 μs) relative to those of Eu2+-containing crown ethers.[85]
In general, macrocyclic ligands with cavity sizes that closely match the ionic radius of Eu2+ are the best ligands to insulate Eu2+ from solvent-based luminescence quenchers. Rigidifying ligands by adding phenyl groups and the addition of substituents like pyridine lead to a decrease in luminescence emission. These factors should be taken into consideration when designing Eu2+-based luminescent materials.
MRI Contrast Agents
In addition to the reducing and luminescence properties of Eu2+, the magnetic properties of this ion are important in the development of Eu2+-based contrast agents for magnetic resonance imaging (MRI). In the last two decades, research on the properties of Eu2+ that are relevant to MRI has been reported. A fast water-exchange rate and a ground state isoelectronic to Gd3+ are properties that make Eu2+ an excellent candidate for use as a contrast agent for MRI.[89] This area started when Merbach and co-workers reported the fast water-exchange rate of the Eu2+ aqua ion, which is in the order of 109 s−1. After detailed mechanistic studies on the water exchange of the Eu2+ aqua ion,[90] investigations of the nuclear and electronic relaxation processes of Eu2+ chelated by polyaminopolycarboxylates, commonly used with Gd3+, were performed.[91] When chelated to diethylenetriamine pentaacetate (DTPA), Eu2+ has a water-exchange rate that is three orders of magnitude faster than that of the clinically approved contrast agent [GdDTPA]2− (87) (Figure 9).[91] In addition to the water-exchange rate, other molecular parameters, including rotational correlation time and electronic spin relaxation time, are important factors that influence the efficiency of contrast agents, known as relaxivity. For example, despite the fast water-exchange of [EuDTPA]3− (88), the reported relaxivity of 88 at 20 MHz is 20% lower than that of 87, because relaxivity is influenced by fast rotation and by fast electronic spin relaxation.[91] The same observation was implicated for the higher relaxivity of Eu2+-containing 1,4,7,10-tetraazacyclododecane-1,4,7,10-tetraacetate [Eu-DOTA]2− (4.74 mM−1 s−1) than that of [EuDTPA]3− (3.49 mM−1 s−1) at 20 MHz and 298 K.[92] Results with other Eu2+-containing complexes indicate that the unfavorable fast electronic spin relaxation rate is not observed with all Eu2+ complexes.[91,93] Tóth and co-workers reported that the electron paramagnetic resonance line widths of 10 are narrower by a factor of 8–10 relative to those of 88, which indicates a slower electronic spin relaxation of Eu2+ in cryptate 10 relative to that in 88.[94] Consequently, the observed electronic spin relaxation for Eu2+-containing cryptates does not limit relaxivity. In addition to slow electronic spin relaxation, this Eu2+-containing cryptate has a fast water-exchange rate (108 s−1) and contains two inner-sphere water molecules in a ten-coordinate complex.[94] These properties are promising because they present an opportunity to develop Eu2+-containing cryptates as effective contrast agents for MRI. A review of the similarities and differences in terms of these molecular parameters for several Eu2+- and Gd3+-containing complexes and their implications on research in MRI contrast agents was written by Merbach and co-workers in 2001.[93]
Figure 9.
Structures of complexes 87 and 88.
While Eu2+ has properties that make it a candidate for use as a contrast agent for MRI, its propensity to get oxidized prevents in vivo applications.[57] We have demonstrated that Eu2+ can be oxidatively stabilized by using coordination chemistry principles, including hard–soft acid–base theory, to the extent that it is more oxidatively stable than Fe2+ in hemoglobin.[57] These studies raised the possibility of using Eu2+-containing cryptates for in vivo applications.
In addition to using modified cryptands as ligands to stabilize Eu2+, we investigated the relaxivity of several Eu2+-containing cryptates 10–11 at different field strengths (1.4,3, 7, 9.4, and 11.7 T).[25,58] These cryptates showed higher relaxivity at higher fields (7 and 9.4 T) than at lower fields (1.4 and 3 T), unlike common Gd3+-containing contrast agents.[25,58] While high relaxivity is a requirement for effective contrast agents, other properties including kinetic stability need to be considered for utility in vivo. We demonstrated that Eu2+-containing cryptates 10 and 11 are stable to transmetallation in the presence of endogenous ions such as Ca2+, Mg2+, and Zn2+,[25] and we are currently investigating the thermodynamic stability and toxicity of several other Eu2+-containing cryptates.
Conclusions
The rapid increase in the number of Eu2+-containing complexes reported in the past few decades has paved the way to making these complexes commonplace in coordination chemistry. New information regarding the stability and utility of Eu2+-containing complexes with a variety of ligands has opened new frontiers in lanthanide chemistry. Careful control of ligand properties produces Eu2+-containing complexes that are useful in synthetic, materials, and medicinal applications.
Acknowledgments
M. J. A. acknowledges Wayne State University (WSU) for a Paul and Carol Schaap Faculty Scholar Award and the National Institute of Biomedical Imaging and Bioengineering of the National Institutes of Health for a Pathway to Independence Career Transition Award (R00EB007129). J. G. was supported by a Paul and Carol Schaap Graduate Fellowship and a Rumble Fellowship.
Biographies
Joel Garcia received his B.S. in Chemistry from the University of the Philippines in Diliman in 2004. During his undergraduate years, he studied the electrocatalytic activity of rhenium(I)-containing 2,3-bis(2′-pyridyl)benzoquinoxaline, 2,3-bis(2′-pyridyl)pyrazine and 2,3-bis(2′-pyridyl)quinoxaline complexes under the supervision of Dr. Girlie Sison. He moved to Wayne State University to join the Allen group for his Ph.D. studies and is studying the properties of europium(II)-containing cryptates and their relevance to magnetic resonance imaging.
Matthew J. Allen, received his B.S. in Chemistry from Purdue University in 1998, where he did undergraduate research with Jillian Buriak, and his Ph.D. from the California Institute of Technology in 2004, where he studied gadolinium-containing complexes with Thomas Meade. Matt then moved to the University of Wisconsin–Madison where he was an NIH postdoctoral fellow in the laboratories of Laura Kiessling and Ronald Raines. In 2008, he started as an assistant professor of chemistry at Wayne State University, where his group studies the chemistry of lanthanides in aqueous solution.
References
- 1.Huheey JE, Keiter EA, Keiter RL. Inorganic Chemistry: Principles of Structure & Reactivity. 4. HarperCollins College Publishers; New York: 1993. p. 601. [Google Scholar]
- 2.Evans WJ. Coord Chem Rev. 2000;206–207:263–283. [Google Scholar]
- 3.Tilley TD, Zalkin A, Andersen RA, Templeton DH. Inorg Chem. 1981;20:551–554. [Google Scholar]
- 4.Adin A, Sykes AG. Nature. 1966;209:804–804. [Google Scholar]
- 5.Sabbatini N, Ciano M, Dellonte S, Bonazzi A, Bolletta F, Balzani V. J Phys Chem. 1984;88:1534–1537. [Google Scholar]
- 6.Jiang J, Higashiyama N, Machida K, Adachi G. Coord Chem Rev. 1998;170:1–29. [Google Scholar]
- 7.Hoffman MV. J Electrochem Soc. 1971;118:933–936. [Google Scholar]
- 8.Ryan FM, Lehmann W, Feldman DW, Murphy J. J Electrochem Soc. 1974;121:1475–1481. [Google Scholar]
- 9.Tilley TD, Andersen RA, Spencer B, Ruben H, Zalkin A, Templeton DH. Inorg Chem. 1980;19:2999–3003. [Google Scholar]
- 10.Evans WJ. J Organomet Chem. 2002;652:61–68. [Google Scholar]
- 11.Wang S, Zhou S, Sheng E, Xie M, Zhang K, Zheng L, Feng Y, Mao L, Huang Z. Organometallics. 2003;22:3546–3552. [Google Scholar]
- 12.Zhou S, Wang S, Sheng E, Zhang L, Yu Z, Xi X, Chen G, Luo W, Li Y. Eur J Inorg Chem. 2007:1519–1528. [Google Scholar]
- 13.Fischer EO, Fischer H. Angew Chem. 1964;76:52. [Google Scholar]; Angew Chem Int Ed Engl. 1964;3:132–133. [Google Scholar]
- 14.Nief F. Dalton Trans. 2010;39:6589–6598. doi: 10.1039/c001280g. [DOI] [PubMed] [Google Scholar]
- 15.Tilley TD, Andersen RA, Zalkin A. Inorg Chem. 1984;23:2271–2276. [Google Scholar]
- 16.Williams AF, Grandjean F, Long GJ, Ulibarri TA, Evans WJ. Inorg Chem. 1989;28:4584–4588. [Google Scholar]
- 17.Rabe GW, Yap GPA, Rheingold AL. Inorg Chim Acta. 1998;267:309–311. [Google Scholar]
- 18.Rabe GW, Yap GPA, Rheingold AL. Inorg Chem. 1997;36:3212–3215. doi: 10.1021/ic9607954. [DOI] [PubMed] [Google Scholar]
- 19.Tilley TD, Andersen RA, Zalkin A. J Am Chem Soc. 1982;104:3725–3727. [Google Scholar]
- 20.White JP, III, Shore SG. Inorg Chem. 1992;31:2756–2761. [Google Scholar]
- 21.Tilley TD, Andersen RA, Zalkin A. Inorg Chem. 1983;22:856–859. [Google Scholar]
- 22.Gansow OA, Kausar AR, Triplett KM, Weaver MJ, Yee EL. J Am Chem Soc. 1977;99:7087–7089. [Google Scholar]
- 23.Burns JH, Baes CF. Inorg Chem. 1981;20:616–619. [Google Scholar]
- 24.Yee EL, Gansow OA, Weaver MJ. J Am Chem Soc. 1980;102:2278–2285. [Google Scholar]
- 25.Garcia J, Kuda-Wedagedara ANW, Allen MJ. Eur J Inorg Chem. 2012:2135–2140. doi: 10.1002/ejic.201101166. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 26.Adachi G-Y, Tomokiyo K, Sorita K, Shiokawa J. J Chem Soc, Chem Commun. 1980:914–915. [Google Scholar]
- 27.White JP, III, Deng H, Shore SG. Inorg Chem. 1991;30:2337–2342. [Google Scholar]
- 28.Cary DR, Arnold J. Inorg Chem. 1994;33:1791–1796. [Google Scholar]
- 29.Maudez W, Meuwly M, Fromm KM. Chem Eur J. 2007;13:8302–8316. doi: 10.1002/chem.200700597. [DOI] [PubMed] [Google Scholar]
- 30.Plečnik CE, Liu S, Chen X, Meyers EA, Shore SG. J Am Chem Soc. 2004;126:204–213. doi: 10.1021/ja0304852. [DOI] [PubMed] [Google Scholar]
- 31.Rybak JC, Müller-Bushbaum K. Z Anorg Allg Chem. 2010;636:126–131. [Google Scholar]
- 32.Janicki R, Mondry A, Starynowicz P. Z Anorg Allg Chem. 2005;631:2475–2477. [Google Scholar]
- 33.Hou Z, Zhang Y, Nishiura M, Wakatsuki Y. Organometallics. 2003;22:129–135. [Google Scholar]
- 34.Deacon GB, Gitlits A, Roesky PW, Bürgstein MR, Lim KC, Skelton BW, White AH. Chem Eur J. 2001;7:127–138. doi: 10.1002/1521-3765(20010105)7:1<127::aid-chem127>3.0.co;2-1. [DOI] [PubMed] [Google Scholar]
- 35.Hamidi S, Deacon GB, Junk PC, Neumann P. Dalton Trans. 2012;41:3541–3552. doi: 10.1039/c2dt11752e. [DOI] [PubMed] [Google Scholar]
- 36.Carretas J, Branco J, Marçalo J, Domingos Â, Pires de Matos A. Polyhedron. 2003;22:1425–1429. [Google Scholar]
- 37.Carretas J, Branco J, Marčalo J, Isolani P, Domingos A, de Matos AP. J Alloys Compd. 2001;323–324:169–172. [Google Scholar]
- 38.Carretas J, Branco J, Marčalo J, Valente N, Waerenborgh JC, Carvalho A, Marques N, Domingos Â, de Matos AP. J Alloys Compd. 2004;374:289–292. [Google Scholar]
- 39.Hitzbleck J, O’Brien AY, Deacon GB, Ruhlandt-Senge K. Inorg Chem. 2006;45:10329–10337. doi: 10.1021/ic061294d. [DOI] [PubMed] [Google Scholar]
- 40.Chen X, Lim S, Plečnik CE, Liu S, Du B, Meyers EA, Shore SG. Inorg Chem. 2004;43:692–698. doi: 10.1021/ic030249s. [DOI] [PubMed] [Google Scholar]
- 41.Forsyth CM, Deacon GB. Organometallics. 2000;19:1205–1207. [Google Scholar]
- 42.Hitzbleck J, Deacon GB, Ruhlandt-Senge K. Eur J Inorg Chem. 2007:592–601. [Google Scholar]
- 43.Hauber SO, Niemeyer M. Inorg Chem. 2005;44:8644–8646. doi: 10.1021/ic0514602. [DOI] [PubMed] [Google Scholar]
- 44.Lee HS, Niemeyer M. Inorg Chem. 2010;49:730–735. doi: 10.1021/ic902055h. [DOI] [PubMed] [Google Scholar]
- 45.Bakker JM, Deacon GB, Forsyth CM, Junk PC, Wiecko M. Eur J Inorg Chem. 2010:2813–2825. [Google Scholar]
- 46.Beaini S, Deacon GB, Hilder M, Junk PC, Turner DR. Eur J Inorg Chem. 2006:3434–3441. [Google Scholar]
- 47.Beaini S, Deacon GB, Delbridge EE, Junk PC, Skelton BW, White AH. Eur J Inorg Chem. 2008:4586–4596. [Google Scholar]
- 48.Niemeyer M. Eur J Inorg Chem. 2001:1969–1981. [Google Scholar]
- 49.Heckmann G, Niemeyer M. J Am Chem Soc. 2000;122:4227–4228. [Google Scholar]
- 50.Deacon GB, Junk PC, Moxey GJ, Ruhlandt-Senge K, StPrix C, Zuniga MF. Chem Eur J. 2009;15:5503–5519. doi: 10.1002/chem.200900229. [DOI] [PubMed] [Google Scholar]
- 51.Deacon GB, Junk PC, Moexy GJ. Chem Asian J. 2009;4:1309–1317. doi: 10.1002/asia.200900176. [DOI] [PubMed] [Google Scholar]
- 52.Daly SR, Girolami GS. Inorg Chem. 2010;49:4578–4585. doi: 10.1021/ic100292u. [DOI] [PubMed] [Google Scholar]
- 53.Bao L, Yingming Y, Mingyu D, Yong Z, Qi S. J Rare Earths. 2006;24:264–267. [Google Scholar]
- 54.Shinoda S, Nishioka M, Tsukube H. J Alloys Compd. 2009;488:603–605. [Google Scholar]
- 55.Starynowicz P, Bukietyńska K, Goląb S, Ryba-Romanowski W, Sokolnicki J. Eur J Inorg Chem. 2002:2344–2347. [Google Scholar]
- 56.Starynowicz P. J Alloys Compd. 2001;323–324:159–163. [Google Scholar]
- 57.Gamage NDH, Mei Y, Garcia J, Allen MJ. Angew Chem Int Ed. 2010;49:8923–8925. doi: 10.1002/anie.201002789. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 58.Garcia J, Neelavalli J, Haacke EM, Allen MJ. Chem Commun. 2011;47:12858–12860. doi: 10.1039/c1cc15219j. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 59.Dietel AM, Döring C, Glatz G, Butovskii MV, Tok O, Schappacher FM, Pöttgen R, Kempe R. Eur J Inorg Chem. 2009:1051–1059. [Google Scholar]
- 60.Dick AKJ, Frey ASP, Gardiner MG, Hilder M, James AN, Junk PC, Powanosorn S, Skelton BW, Wang J, White AH. J Organomet Chem. 2010;695:2761–2767. [Google Scholar]
- 61.Yao S, Chan HS, Lam CK, Lee HK. Inorg Chem. 2009;48:9936–9946. doi: 10.1021/ic901327m. [DOI] [PubMed] [Google Scholar]
- 62.Evans WJ, Perotti JM, Brady JC, Ziller JW. J Am Chem Soc. 2003;125:5204–5212. doi: 10.1021/ja020957x. [DOI] [PubMed] [Google Scholar]
- 63.Roesky PW. Inorg Chem. 2006;45:798–802. doi: 10.1021/ic051738q. [DOI] [PubMed] [Google Scholar]
- 64.Panda TK, Zulys A, Gamer MT, Roesky PW. J Organomet Chem. 2005;690:5078–5089. [Google Scholar]
- 65.Datta S, Gamer MT, Roesky PW. Organometallics. 2008;27:1207–1213. [Google Scholar]
- 66.Jenter J, Gamer MT, Roesky PW. Organometallics. 2010;29:4410–4413. [Google Scholar]
- 67.Heitmann D, Jones C, Mills DP, Stasch A. Dalton Trans. 2010;39:1877–1882. doi: 10.1039/b924367d. [DOI] [PubMed] [Google Scholar]
- 68.Wiecko M, Roesky PW. Organometallics. 2009;28:1266–1269. [Google Scholar]
- 69.Evans WJ, Walensky JR, Furche F, DiPasquale AG, Rheingold AL. Organometallics. 2009;28:6073–6078. [Google Scholar]
- 70.Jones C, Stasch A, Woodul WD. Chem Commun. 2009:113–115. doi: 10.1039/b817933f. [DOI] [PubMed] [Google Scholar]
- 71.Wiecko M, Roesky PW. Organometallics. 2007;26:4846–4848. [Google Scholar]
- 72.Gamer MT, Roesky PW, Konchenko SN, Nava P, Ahlrichs R. Angew Chem. 2006;118:4558. doi: 10.1002/anie.200600423. [DOI] [PubMed] [Google Scholar]; Angew Chem Int Ed. 2006;45:4447–4451. doi: 10.1002/anie.200600423. [DOI] [PubMed] [Google Scholar]
- 73.Evans WJ, Clark RD, Ansari MA, Ziller JW. J Am Chem Soc. 1998;120:9555–9563. [Google Scholar]
- 74.Moore JA, Cowley AH, Gordon JC. Organometallics. 2006;25:5207–5209. [Google Scholar]
- 75.Sheng E, Zhou S, Wang S, Yang G, Wu Y, Feng Y, Mao L, Huang Z. Eur J Inorg Chem. 2004:2923–2932. [Google Scholar]
- 76.Zhang K, Zhang W, Wang S, Sheng E, Yang G, Xie M, Zhou S, Feng Y, Mao L, Huang Z. Dalton Trans. 2004:1029–1037. doi: 10.1039/b314115b. [DOI] [PubMed] [Google Scholar]
- 77.Wang S, Feng Y, Mao L, Sheng E, Yang G, Xie M, Wang S, Wei Y, Huang Z. J Organomet Chem. 2006;691:1265–1274. [Google Scholar]
- 78.Wu Y, Wang S, Qian C, Sheng E, Xie M, Yang G, Feng Q, Zhang L, Tang X. J Organomet Chem. 2005;690:4139–4149. [Google Scholar]
- 79.Miao H, Wang S, Zhou S, Wei Y, Zhou Z, Zhu H, Wu S, Wang H. Inorg Chim Acta. 2010;363:1325–1331. [Google Scholar]
- 80.Zhou S, Yin C, Wang H, Zhu X, Yang G, Wang S. Inorg Chem Commun. 2011;14:1196–1200. [Google Scholar]
- 81.Wang S, Wang S, Zhou S, Yang G, Luo W, Hu N, Zhou Z, Song HB. J Organomet Chem. 2007;692:2099–2106. [Google Scholar]
- 82.Wei Y, Yu Z, Wang S, Zhou S, Yang G, Zhang L, Chen G, Qian H, Fan J. J Organomet Chem. 2008;693:2263–2270. [Google Scholar]
- 83.Zhu X, Fan J, Wu Y, Wang S, Zhang L, Yang G, Wei Y, Yin C, Zhu H, Wu S, Zhang H. Organometallics. 2009;28:3882–3888. [Google Scholar]
- 84.Kropp JL, Windsor MW. J Chem Phys. 1963;39:2769–2770. [Google Scholar]
- 85.Starynowicz P. Polyhedron. 2003;22:2761–2765. [Google Scholar]
- 86.Christoffers J, Starynowicz P. Polyhedron. 2008;27:2688–2692. [Google Scholar]
- 87.Starynowicz P, Bukietyńska K. Eur J Inorg Chem. 2002:1835–1838. [Google Scholar]
- 88.Starynowicz P. Polyhedron. 2003;22:337–345. [Google Scholar]
- 89.Caravan P, Merbach AE. Chem Commun. 1997:2147–2148. [Google Scholar]
- 90.Caravan P, Tóth É, Rockenbauer A, Merbach AE. J Am Chem Soc. 1999;121:10403–10409. [Google Scholar]
- 91.Seibig S, Tóth É, Merbach AE. J Am Chem Soc. 2000;122:5822–5830. [Google Scholar]
- 92.Burai L, Tóth É, Moreau G, Sour A, Scopelliti R, Merbach AE. Chem Eur J. 2003;9:1394–1404. doi: 10.1002/chem.200390159. [DOI] [PubMed] [Google Scholar]
- 93.Tóth É, Burai L, Merbach AE. Coord Chem Rev. 2001;216–217:363–382. [Google Scholar]
- 94.Burai L, Scopelliti R, Tóth É. Chem Commun. 2002:2366–2367. doi: 10.1039/b206709a. [DOI] [PubMed] [Google Scholar]











