Anda di halaman 1dari 16

Coordination Chemistry Reviews 260 (2014) 2136

Contents lists available at ScienceDirect

Coordination Chemistry Reviews


journal homepage: www.elsevier.com/locate/ccr

Review

The electrochemical behavior of cerium(III/IV) complexes: Thermodynamics, kinetics and applications in synthesis
Nicholas A. Piro, Jerome R. Robinson, Patrick J. Walsh, Eric J. Schelter
P. Roy, Diana T. Vagelos Laboratories, Department of Chemistry, University of Pennsylvania, Philadelphia, PA 19104, United States

Contents 1. 2. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Aqueous cerium redox chemistry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.1. Acidic media . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.2. Basic media . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.3. Polyoxometalates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Non-aqueous electrochemistry of cerium complexes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.1. Measurements of cerium electrochemistry in ionic liquid media . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.2. Complexes with oxyacids, neutral nitrogen bases, and crown ether ligands . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.3. Complexes with salen, aryl oxide and acetylacetonate ligands . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.4. Complexes of tetrapyrrole and other tetraaza macrocycles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.5. Organometallic complexes of cerium and endohedral fullerene complexes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.6. Cerium amides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Preparative chemistry of cerium(IV) complexes by oxidation reactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.1. Outer-sphere oxidation, H-atom abstraction, and autooxidation reactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.2. Oxidative functionalization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.3. Oxidation-induced ligand redistribution reactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . The role of kinetics in cerium electrochemistry and oxidation reactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Conclusions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22 22 22 23 23 24 25 25 25 27 30 30 31 31 31 32 33 35 35 35

3.

4.

5. 6.

a r t i c l e

i n f o

a b s t r a c t
A key characteristic of the element cerium is its reversible redox chemistry between trivalent and tetravalent forms, which is central to the application of cerium in synthetic and materials chemistry. Herein we survey the general thermodynamic and kinetic characteristics and reported potentials for molecular cerium redox chemistry. The collected data illustrate that the local electronic environment provided by the coordination sphere around a cerium ion has a great effect on the oxidizing ability of the ion. The

Article history: Received 28 June 2013 Accepted 26 August 2013 Available online 7 September 2013

Abbreviations: acac, acetylacetonate; BINOL, 1,1 -bi-2-napthol; Bu, butyl; BQ, benzoquinone; CAN, ceric ammonium nitrate; CV, cyclic voltammetry/voltammagram; CMPO, n-octyl(phenyl)-N,N-diisobutylcarbamoylmethylphosphine oxide; Cp, cyclopentadienyl; dap, diazaporphyrin; DMSO, dimethylsulfoxide; DPA, dipicolinic acid; DTPA, diethylenetriaminepentaacetic acid; EDTA, ethylenediaminetetraacetic acid; Epa , electrochemical potential at peak anodic current; Epc , electrochemical potential at peak cathodic current; Et, ethyl; Fc, ferrocene; H6TrenSal, tris((2-hydroxybenzyl)aminoethyl)amine; HBP, hexadecahydrotetrabenzoporphyrin; HOPO, hydroxypyridinone; IL, ionic liquid; i Pr, iso-propyl; irr, irreversible; MBP, 2,2 -methylenebis(6-tert-butyl-4-methylphenolate); Me, methyl; Me-3,2-HOPO, 1-methyl-3-hydroxy-2(1H)-pyridinone; NBS, N-bromosuccinimide; Nc, naphthalocyanine; NctBu, tetra-tert-butylnapthalocyanine; NHE, normal hydrogen electrode; NTA, nitrilotriacetic acid; OEP, octaethylporphyrin; omtaa, 5,14-dihydro-2,3,6,8,11,12,15,17-octamethyldibenzo[b,i][1,4,8,11]tetraazacyclotetradecine; OPTAP, octapropyltetrazaporphyrin; Pc, phthalocyanine; POM, polyoxometallate; RE, rare earth; salen, N,N -ethylenebis(salicylimine); salophen, N,N -(1,2-phenylene)bis(salicylimine); SCE, saturated calomel electrode; TAP, tetra(4-methoxyphenyl)porphyrin; TBABF4 , tetrabutylammonium tetrauoroborate; TBAP, tetra-n-butylammonium perchlorate; TBAPF6 , tetra-n-butylammonium hexauorophosphate; TBPP, tetra(4-tert-butylphenyl)porphyrin; i Bu, tert-butyl; TClP, tetra(4-chlorophenyl)porphyrin; TEtAP, tetraethylylammonium perchlorate; THF, tetrahydrofuran; Tf, triuoromethylsulfonyl; tmtaa, dibenzotetramethylaza[14]annulene; TPAB, [N(n-C3 H7 )4 ][B(3,5-(CF3 )2 C6 H3 )4 ]; TPP, tetraphenylporphyrin; TPyP, tetrapyridylporphyrin; TREN, tris(2-aminoethyl)amine; TrenSal, tris((2-hydroxybenzylidene)aminoethyl)amine; TTP, tetratolylporphyrin; XANES, X-ray absorption nearedge structure. Corresponding author. Tel.: +1 215 898 8633; fax: +1 215 573 2112. E-mail address: schelter@sas.upenn.edu (E.J. Schelter). 0010-8545/$ see front matter 2013 Elsevier B.V. All rights reserved. http://dx.doi.org/10.1016/j.ccr.2013.08.034

22

N.A. Piro et al. / Coordination Chemistry Reviews 260 (2014) 2136

Keywords: Lanthanides Cerium Electrochemistry Electron transfer Oxidation

survey also illustrates the ligand types that most effectively stabilize each oxidation state. We expect the collection and comparison of these data will facilitate the development of new cerium(IV) chemistry and applications in oxidation and reduction chemistry.

2013 Elsevier B.V. All rights reserved.

1. Introduction The accessibility of the 4f0 tetravalent oxidation state for molecular cerium complexes in solution is unique among the lanthanide elements. The thermodynamic basis for the isolation of molecular cerium(IV) compounds originates from the enhanced stability afforded by the [Xe] noble gas electronic conguration for the ion [1]. The existence of molecular cerium(IV) compounds and the ability to cycle between the +3 and +4 oxidation states plays a key role in the separation of cerium from lanthanide-containing ore, and more generally in studies of valence-control for nuclear fuel reprocessing [24]. The facile separation of this element from other lanthanides is a major contributing factor to the relatively low cost of cerium. The interconversion of the +3 and +4 oxidation states also underlies the utility of cerium oxides as heterogeneous catalysts and in renewable energy applications [5,6]. Traditionally, it is the potent oxidizing ability of cerium(IV) that is exploited in the redox chemistry of the element, with ceric ammonium nitrate (CAN) oxidation reactions being the primary examples [710]. While the applications of CAN, Fig. 1, and other cerium(IV) reagents in organic synthesis are extensive and have been reviewed previously [710], the ability of cerium(III) to carry out reductive transformations has remained relatively unexplored. This is in contrast to the reductive chemistry of divalent lanthanides, which have formed an area of active investigation [7,11,12]. In this review we survey the thermodynamics of molecular cerium redox chemistry, primarily through assessment of electrochemical measurements, with the goal of gaining an understanding of factors that affect cerium reduction potentials. The tuning of redox potentials through variation of ligand environments for transition metal ions is a widely accepted phenomenon. Perhaps it is the popular dogma that the bonding in lanthanide complexes is

completely ionic that has, in our estimate, prevented a systematic review of ligand effects on cerium reduction potential. However, we show through this survey that the local electronic environment provided by the coordination sphere around a cerium ion can have a great effect on the oxidizing or reducing ability of the cerium ion. The true range of cerium(IV) reduction potentials is remarkably broad, spanning at least 2.3 V in aqueous solution and 2.9 V under non-aqueous conditions (vide infra). Thus, while cerium(IV) is often considered a strong oxidant, variation of the coordination environment can render cerium(III) a strong reductant. From our survey of reported cerium electrochemistry, it is evident that noticeable differences exist between the redox chemistry of cerium and that of transition metal ions. Cerium in both its +3 and +4 oxidation state remains a hard cation in the Pearson terminology [1416]. It is therefore best stabilized by hard anions, which provide the greatest stabilization of cerium(IV) complexes. Thus, certain ligands that may be considered electron-rich and often stabilize high-valent transition metal ions confer relatively smaller stabilization to tetravalent cerium, including amides and cyclopentadienides. Another unique aspect of cerium redox chemistry is the major role of ligand reorganization in oxidation reactions. The change from f1 to f0 conguration upon oxidation from the trivalent to to tetravalent state results in a change in ionic radius from 1.01 A [17]. In contrast to the d-block transition metals, covalent 0.87 A bonding plays only a negligible role in shaping the coordination sphere of cerium complexes and maintaining the coordination sphere throughout an oxidation reaction. Thus, large inner sphere reorganization energies are expected and thermodynamic considerations alone do not necessarily confer predictive power in cerium oxidation reactions [18]. Herein we document these effects with a brief review of aqueous chemistry followed by a focus on cerium complexes that have been characterized electrochemically in non-aqueous conditions. We also discuss the synthetic chemistry of cerium(III/IV) transformations and the kinetics of electron transfer reactions. For ease of comparison in this review we have converted all electrochemical potentials to a common reference, SCE. This was done by performing the following adjustments to the published values when reported versus the following references: Fc/Fc+ in THF, +0.56 V; Fc/Fc+ in MeCN, +0.40 V; Fc/Fc+ in H2 O, +0.16 V; Ag/AgCl (sat KCl), 0.045 V; NHE, 0.24 V [1923].

2. Aqueous cerium redox chemistry 2.1. Acidic media In acidic, aqueous media Ce(IV) is a potent one-electron oxidant. This was recognized as early as 1931, when a series of papers documented the cerium(III/IV) reduction potential under a variety of conditions. Kunz, Noyes and Garner, and Smith and Getz each published reference values for cerium reduction potentials in acidic conditions ranging from +1.04 to +1.63 V, demonstrating a dependence on acid identity and concentration [2426]. Anionand concentration-dependent reduction potentials under several of these conditions are presented in Table 1. The standard potential

Fig. 1. The structure of the [Ce(NO3 )6 ]2 anion in CAN. The coordinates were obtained from Ref. [13].

N.A. Piro et al. / Coordination Chemistry Reviews 260 (2014) 2136 Table 1 Redox potentials of cerium ions under various aqueous conditions. Complexes Ce(ClO4 )3 Ce(ClO4 )3 Ce(ClO4 )3 Ce(ClO4 )3 Ce(NO3 )3 CeCl3 Ce2 (SO4 )3 CeCl3 Ce(EDTA) a (1:1/[Ce]:[EDTA])a Ce(DTPA)2 a (1:2.5/[Ce]:[DTPA])a Ce(NTA)2 a Potential (V) vs. SCE 1.63 1.47 1.46 1.39 1.37 1.22 1.20 1.04 0.805 0.546 0.464 Reported potential (V) 1.87 vs. NHE 1.71 vs. NHE 1.70 vs. NHE 1.63 vs. NHE 1.61 vs. NHE 1.46 vs. NHE 1.44 vs. NHE 1.28 vs. NHE 0.85 vs. Ag/Ag+ 0.581 vs. Ag/Ag+ 0.509 vs. Ag/Ag+ Conditions 8 M HClO4 4.83 M HClO4 1 M HClO4 0.1 M HClO4 1 M HNO3 1 M HCl, 0.2 M NaClO4 1 M H2 SO4 1 M HCl 0.2 M KCl 1 M KNO3 , pH 4.19 (0.002 M Ce(NO3 )3 + 0.05 M NTA) 1 M NaNO3 , pH 6.0) 0.16 M NaOAc pH 4.66 0.1 M KClb with DMF 0.1 M KClb 1.0 M SO4 2 , pH 8.5 5 M [CO3 2 ], pH 12.7 5 M [CO3 2 ], pH 13.6 5 M NaOH/1 M tironc 5 M NaOH/1 M catechol Ref.

23

[26] [28] [26] [28] [25,26] [24,26,29] [24,26] [26] [30] [31] [32]

[Ce(DPA)3 ][(C(NH2 )3 )+ ]a Ce(HOPO-2)2 a Ce(HOPO-3)2 a Ce(HOPO-4)a [Ce(CO3 )5 ]6 [Ce(CO3 )5 ]6 [Ce(tironate)4 ]12 c [Ce(catecholate)4 ]4
a b c

0.44 0.26 0.24 0.13 0.21 0.25 0.49 0.69

0.503 vs. NHE 0.475 vs. NHE 0.175 vs. Ag/Ag+

[33] [34] [34] [35] [36] [36] [37] [38]

Structures of EDTA, DTPA, NTA, DPA, HOPO-2, and HOPO-3 are provided in Fig. 2. Measured using static potentiometry. Tironate = 4,5-dihydroxy-1,3-benzenedisulfonate.

for the ion, at +1.46 V versus SCE (+1.70 V versus NHE), is taken from the oxidation of Ce(ClO4 )3 in 1 M HClO4 [26]. The potentials for cerium oxidation increase markedly as the acids coordinating ability decreases: changing from 1 M HCl to 1 M HClO4 shifts the cerium ion potential by over 400 mV. Furthermore, it has been recognized that addition of uoride to solutions of Ce(IV) gives mixtures that are incapable of oxidizing iodide, further reecting how the Ce(IV) redox potential varies widely with the identity of ligands and counter ions in aqueous solution [27].

2.2. Basic media The observations from acidic media, that lower acid concentration and coordinating ligands stabilize the cerium(IV) ion, suggest that basic conditions would yield stabilization of cerium(IV) reduction potentials. This is true and is attributed to strong coordination of the oxophilic cerium ion by oxygen-donor ligands in solution. Morris and Hobart investigated the redox chemistry of cerium(IV), prepared by air-oxidation of CeCl3 , in strongly basic and carbonaterich solutions [36]. Under these conditions the cerium(III/IV) couple was observed at 250 mV versus SCE. Polyanionic chelating ligands impart stability to high oxidation compounds and also signicantly reduce the cerium potential in aqueous solution, Table 1. This was demonstrated by the work of Raymond and co-workers, where catecholates in basic solution depress the reduction potential of Ce(IV) to 0.69 V, a shift of over 2 V relative to the formal Ce(III/IV) couple measured in aqueous perchloric acid [37,38]. The stabilized cerium(III/IV) oxidation observed in the reversible electrochemistry of cerium complexes with ligands related to Me-3,2-HOPO, Fig. 2, was included in a thermodynamic cycle to assess binding strengths of catechol-type ligands. The authors concluded that cerium(IV) electrochemistry in this family of complexes served as models for the behavior of plutonium(IV) complexes. However, while the cerium(IV) and and 0.86 A, plutonium(IV) ions have similar ionic radii, at 0.87 A respectively [17], subsequent work has shown that the electrochemical behavior of cerium(IV) in other ligand frameworks more

closely resembles that of berkelium(IV) [3942]. Recent work from Abergel et al. has investigated the solution thermodynamic properties of an octadentate hydroxypyridinonate, HOPO-4, Fig. 2, which is a promising candidate as a therapeutic for the removal of Ln and An ions [35]. The incorporation of four hydroxypyridinonate units in HOPO-4 results in a 100 mV shift of the Ce(III/IV) couple to more negative values compared with HOPO-3. Additionally, the HOPO ligands listed in Table 1 and shown in Fig. 2 demonstrate remarkable solution thermodynamic stability, having log () values ranging from 40 to 42 [35]. Multidentate carboxylic acids also signicantly shift the Ce(III/IV) couple to less oxidizing potentials. Electrochemical investigations of cerium complexes from ethylenediaminetetracetic acid (EDTA), diethylenetriaminepentaacetic acid (DTPA), nitrilotriacetic acid (NTA), and dipicolinate (DPA), Fig. 2, demonstrate that the Ce(III/IV) couple can be shifted by 8251190 mV to less oxidizing potentials from those observed at low pH. The E1/2 values for the cerium carboxylates also shift to less oxidizing potentials when more carboxylate groups are coordinated to the Ce(III) center, Table 1.

2.3. Polyoxometalates The concept of stabilizing tetravalent cerium by introduction of polyanions to the coordination sphere has been explored through the use of polyoxometalates (POMs). These species can act as large, polyanionic ligands that coordinate cerium and can serve to stabilize the tetravalent state, and have applications in radiological fuel separations science [4]. POMs ranging in charge from 10 to 20 have been used to form complexes with cerium. The shifts in cerium redox potentials created by the use of these polyanions are shown in Table 2. Detailed studies of the Ce-POMs have revealed that the charge of the POM is not the only consideration for the observed cerium redox potential. Instead, the local coordination environment available to cerium upon complexation to the oxometallate scaffold greatly affects the cerium(III/IV) couple. For example,

24

N.A. Piro et al. / Coordination Chemistry Reviews 260 (2014) 2136

Fig. 2. Polydentate pro-ligands studied in cerium binding [3035].

the Preyssler anion, [Mn+ P5 W30 O110 ]n15 , carries a large negative formal charge; however, the large, rigid coordination sites are not able to effectively approach and donate charge to stabilize a cerium(IV) ion [4345]. Surprisingly, oxidation of the cerium(III) ion is more difcult when complexed by the Preyssler anion than the cerium(III) cation in acidic conditions, and oxidation to cerium(IV) did not reportedly occur even at potentials of +1.9 V versus Ag/Ag+ . In contrast, the exible Wells-Dawson framework, [-2-P2 W17 O61 ]6 , can contract around the cerium ion upon oxidation from cerium(III) to cerium(IV). The by ca. 0.25 A exibility of the coordination environment facilitates the redox
Table 2 Reduction potentials of cerium ions coordinated to polyoxometallate complexes. Complex [Ce(W5 O18 )2 ]6 a [Ce(PW11 O39 )2 ]10 [Ce(SiW11 O39 )2 ]12 [Ce(-2-P2 Wl7 O61 )2 ]16 [Ce(PW11 O39 )2 ]10 [Ce(-2-P2 Wl7 O61 )2 ]16 [Ce(SiW11 O39 )2 ]12
a

change and affords reversible electrochemistry at a modest +0.32 V [43]. 3. Non-aqueous electrochemistry of cerium complexes Studies of the coordination chemistry of cerium ions under nonaqueous conditions allow for a wide range of ligand environments that are not available under aqueous conditions. Also, low dielectric solvents serve to destabilize charged compounds, which in the presence of polyanionic ligands can stabilize cerium(IV) complexes relative to cerium(III). Here we will survey the redox potentials

Potential vs. SCE (V) 0.86 0.64 0.67 0.54 0.38 0.32 0.28

Reported potential (V) 1.10 vs. NHE 0.88 vs. NHE 0.91 vs. NHE 0.78 vs. NHE 0.424 vs. Ag/AgCl 0.365 vs. Ag/AgCl 0.323 vs. Ag/AgCl

Conditions 0.1 M KCl 0.1 M KCl 0.1 M KCl 0.1 M KCl pH 4.5 0.1 M Na2 SO4 pH 4.5, pH 5.7, 0.1 M Na2 SO4 pH 4.5, 0.1 M Na2 SO4

Ref. [44] [44] [44] [44] [45] [45], [43] [45]

In the original report the authors had listed the formulation as Ce[W10 O35 ]6 , however, the correct formulation is [Ce(W5 O18 )2 ]6 [4].

N.A. Piro et al. / Coordination Chemistry Reviews 260 (2014) 2136 Table 3 Reduction potentials of cerium complexes in ionic liquids. Complex [CeCl6 ]
3-/2

25

Potential (V) vs. SCE

Reported potential (V) 0.64 vs Al 0.37 vs Fc/Fc+ a 0.519 vs Pd 0.308 vs Ag/Ag+ 0.25 vs Fc+

Conditions AlCl3 :MeEtimCl (2:1) 40 C bmimCl (70 C) [BuMePyr+ ][NTf2 ] (50 C, 0.095 M CeIII )

Ref. [55] [55,56] [54]

[CeCl6 ] [CeCl6 ]3/2


a

3/2

Fc/Fc+ vs. Al = 0.270 V in AlCl3 :MeEtimCl, 2:1[53,59,20,60].

for reported cerium(III) and cerium(IV) complexes that have been studied in non-aqueous condition, principally in MeCN, DCM or THF solvents. 3.1. Measurements of cerium electrochemistry in ionic liquid media Due to desirable physical properties, such as low vapor pressures, high ion conductivity, and non-ammability, ionic liquids (ILs) have found many applications in chemistry and engineering. Applications related to cerium include synthetic chemistry, separations chemistry, and studies toward the processing of spent nuclear fuel [4652]. Advantages to performing electrochemical measurements in ILs include the large measurement window, the non-coordinating nature of the electrolyte, and mild measurement conditions [53]. For example, measurement of [CeCl6 ]3 in [BuMePyr+ ][NTf2 ] showed no Cl oxidation events and exhibited no stability issues. In contrast, Ce(IV) ions are otherwise unstable in alkali metal chloride molten salts [54]. Despite these potential benets, there are only limited reports on the electrochemistry of cerium complexes in ILs (Table 3). In the case of CeCl3 , the nature of the analyte in the IL is critical for applications. In [BuMePyr+ ][NTf2 ], adjusting the Ce:Cl ratio to 1:6 results in the clean formation and effective electrochemical measurement of the complex ion, [CeCl6 ]3 . However, at a ratio of 1:3, CeCl3 rapidly precipitates, which could be advantageous for separations applications [54]. Generally, comparison of potentials between different ILs is complicated by the absence of common voltammetric reference standards [53]. However, measurement of [CeCl6 ]3 in two ILs [54,55], AlCl3 :MeEtimCl (2:1) and [BuMePyr+ ][NTf2 ], showed 100 mV variation and suggested that the electrochemical behavior of the complex ion does not change dramatically between the two ILs. ILs undoubtedly hold great promise for future applications in cerium redox chemistry. 3.2. Complexes with oxyacids, neutral nitrogen bases, and crown ether ligands Complexes of neutral ligands or oxyanions form a large portion of the known cerium coordination chemistry [16]. Despite this body

of work, electrochemical characterization of the ion in well-dened complexes with such ligands is limited. Complexes of cerium(III) with the three mixed donor ligands shown in Fig. 3 have been characterized electrochemically. The Schiff base ligand L1 was treated with Ce(NO3 )3 hydrate and Ce(L1 )(NO3 )3 was isolated. This complex displays an irreversible oxidation at +0.77 V in MeCN, which is shifted by only a small amount relative to Ce(NO3 )3 hydrate under similar conditions (+0.83 V). A cerium complex of the formylderived ligand L2 , Ce(L2 )(NO3 )3 , also shows irreversible redox behavior, at +0.85 V in DMSO [57,58]. The phosphonate/amine macrocycle L3 has more reversible electrochemistry, but also shows evidence of ligand decomposition upon oxidation in MeCN [59]. The potentials for each of these complexes, Table 4, are not signicantly shifted versus that for cerium(III) nitrate under similar conditions, though they are shifted from the -ate complex [n Bu4 N]2 [Ce(NO3 )6 ] which has been measured at 1.02 V in MeCN at 40 C. Phosphine oxides complexes of cerium have also been characterized electrochemically. A CMPO complex was formed by addition of CMPO to cerium(III) nitrate in MeCN and this mixture characterized electrochemically. Also, (n BuO)3 PO has been used as an extractant for partitioning cerium ions from aqueous solutions and the extracts similarly characterized [60,61]. Each of these ligands forms a complex with cerium that displays reversible electrochemistry, however, the potentials for these cerium(III/IV) redox events remain high at ca. +1.1 V. 3.3. Complexes with salen, aryl oxide and acetylacetonate ligands Complexes of cerium with a variety of anionic oxygen-based ligands such as alkoxides, aryl oxides (including salen-type ligands), and -diketonates have been prepared and studied electrochemically. This body of work demonstrates that such ligands are capable of depressing the cerium(III/IV) potential considerably relative to the standard cerium potential in aqueous solution. Data for the complexes described in this section are compiled in Table 5. The electrochemistry of a series of 8-coordinate bis(salen) complexes with a central cerium(IV) ion has been reported [63]. The potential of the parent compound Ce(salen)2 was reported at 0.676 V in MeCN. The data reveal that even within a series of isostructural ligands, Fig. 4, potentials vary by 250 mV as the

Fig. 3. Assorted ligands for non-aqueous cerium complexes [5760].

26

N.A. Piro et al. / Coordination Chemistry Reviews 260 (2014) 2136

Table 4 Reduction potentials of non-aqueous cerium nitrates supported by ligands. Complex Ce(CMPO)3 (NO3 )3 Potential (V) vs. SCE 1.1 Reported potential (V) 0.74 vs Fc 0.80 vs Ag/Ag+ 1.29 vs NHE 0.62 vs Fc/Fc+ 0.42 vs Fc/Fc+ Conditionsa 0.1 M TBAP MeCN Ref. [60]

[n Bu4 N]2 [Ce(NO)6 ] Ce(L2 )(NO3 )3 Ce(NO3 )3 6H2 O Ce(L1 )(NO3 )3 Ce(L3 )
a

1.02 0.85 (irr) 0.83 0.77 (irr) 0.60

MeCN at 40 C 0.1 M TBAP in DMSO MeCN 0.1 M TEtAP in MeCN 0.05 M TBAPF6 in MeCN

[62] [58] [57] [57] [59]

TBAP = [NBu4 ][ClO4 ], TBAPF6 = [NBu4 ][PF6 ], TEtAP = [NEt4 ][ClO4 ].

Table 5 Reduction potentials of cerium ion supported by anionic oxygen ligands. Complex Ce(acac)4 Ce(HOPO-2)2 Ce(HOPO-1)2 Ce(HOPO-3)2 [Li(THF)2 ][Ce(MBP)2 (THF)2 ] Ce(5,5 -Br2 -Salen)2 [Li3 (THF)4 ][(BINOLate)3 Ce(THF)] Potential (V) vs. SCE 0.02 0.16 0.18 0.21 0.37 0.52 Epc = 0.53 Epa = 0.12 0.53 0.68 Epc = 0.74 Epa = 0.34 0.78 Epc = 0.83 Epa = 0.49 Epc = 1.51 Epa = 0.45 Epc = 1.83 Epa = 1.14 Reported potential (V) 0.22 0.02 vs. NHE 0.11 vs. Ag/AgCl 0.13 vs Ag/AgCl 0.16 vs. Ag/AgCl 0.93 vs. Fc/Fc+ Epc = 1.085 Epa = 0.445 vs. Fc/Fc+ Epc = 1.295 Epa = 0.895 vs. Fc/Fc+ Epc = 1.385 Epa = 1.045 vs. Fc/Fc+ Epc = 2.07 Epa = 1.01 vs. Fc/Fc+ Epc = 2.39 Epa = 1.70 vs. Fc/Fc+ Conditionsa 0.1 M TBAPF6 in MeCN/acetone 0.1 M TBAPF6 in MeCN 0.1 M TBAPF6 in MeCN 0.1 M TBAPF6 in MeCN 0.1 M TPAB in THF 0.1 M TBAP in MeCN 0.1 M TPAB in THF Ref. [70] [34] [34] [34] [66] [63] [67]

Ce(salophen)2 Ce(salen)2 [Na3 (THF)6 ][(BINOLate)3 Ce]

0.1 M TBAP in MeCN 0.1 M TBAP in MeCN 0.1 M TPAB in THF

[63] [63] [67]

Ce(5,5 -(OMe)2 -Salen)2 [K3 (THF)6 ][(BINOLate)3 Ce]

0.1 M TBAP in MeCN 0.1 M TPAB in THF

[63] [67]

Ce(L4 )(Ot Bu)2

0.5 M TPAB in THF

[65]

Ce(L5 )(Ot Bu)2

0.5 M TPAB in THF

[65]

TBAP = [NBu4 ][ClO4 ], TBAPF6 = [NBu4 ][PF6 ], [NEt4 ][ClO4 ], TPAB = [N(n-C3 H7 )4 ][B(3,5-(CF3 )2 C6 H3 )4 ].

electron-donating ability of the ligand is tuned through substitutions on the aromatic ring. The more electron-rich 5-methoxy analog adjusts the reversible cerium(III/IV) potential to 0.776 V, and the less-donating 5-bromosalen and salophen ligands have reversible couples at 0.521 and 0.529 V, respectively. An interesting comparison can be drawn with the electrochemical behavior of 1:1 Cu(II) complexes of the identical series of salen ligands. In the case of the Ce(III) complexes, the shift in potentials upon changing from methoxy to bromo is 255 mV, more than twice

the difference observed for the same ligands when they are used to form 1:1 complexes with Cu(II). In the case of the Cu(II) complexes, the Cu(I/II) redox couple potential shifts by 112 mV upon changing the salen 5,5 -substitution from methoxy to bromo [64]. Thus, the Ce(III/IV) couple is more sensitive to substituent effects in its 2:1 salen complexes than the Cu(I/II) couple in the 1:1 copper complexes. A comparison of the potential ranges for the cerium and copper complexes are depicted in Fig. 5.

Fig. 4. 5,5 -X2 -salen and salophen ligands, X = H, OMe, Br.

N.A. Piro et al. / Coordination Chemistry Reviews 260 (2014) 2136

27

Fig. 5. Ranges in metal ion potentials for series of Cu and Ce 5,5 -substituted salen complexes. Values taken from Refs. [63,64].

A complex of cerium(IV) with a ferrocene-bridged, salen-like ligand has also been prepared with two tert-butoxide ancillary ligands, Fig. 6 [65]. Although the ligand features a redox-active component, XANES measurements supported a cerium(IV) oxidation state assignment while Mssbauer measurements corroborated a divalent ferrocene backbone. This complex has an extraordinary reduction potential in THF of 1.51 V; however, the shape of the voltammetric wave and large peak separation indicate a signicant kinetic barrier and overpotential associated with the electrochemical event, which biases the measured potential. The related ferrocene-linked phosphinimido ligands adjusts the Ce(IV) to Ce(III) reduction event to 1.83 V with similar signs of electrochemical irreversibility, despite the measurements being performed in higher concentrations of weakly-coordinating electrolyte, TPAB. Even with the associated overpotentials, the dramatic shifts in potential suggest that alkoxide and aryloxide ligands have a potent effect on stabilization of cerium(IV) relative to cerium(III). A methylene bisphenolate ligand, 2,2 -methylenebis(6-tertbutyl-4-methylphenolate), has recently been employed to provide 2:1 complexes with cerium that are easily oxidized chemically and electrochemically to give cerium(IV) complexes. The reversible cerium(III/IV) potential for Ce(MBP)2 (THF)2 lies at 0.37 V versus SCE [66]. A series of heterobimetallic ceriumalkali metal BINOLate complexes have been synthesized that demonstrated the effect of ligand reorganization on the redox events of a central cerium ion. A shift of 450 mV is realized from [Li3 (THF)n ][(BINOLate)3 Ce(THF)] to [K3 (THF)n ][(BINOLate)3 Ce], where the alkali metal ions are integral to the structure of the complexes and coordination environment (including coordination number) at the cerium cation (Fig. 6). Secondary coordination sphere effects have large impacts on the reduction potentials of transition metal ions as demonstrated, for example, in proteins. It is remarkable, however, that an effect of this magnitude is also found for a lanthanide ion (see Section 5) [68].

Fig. 7. Structure of Ce(acac)4 . Coordinates from Ref. [70].

Cerium tetrakis(acetylacetonate), Fig. 7, known since 1913, has been studied with regard to the thermodynamics of the cerium redox event, as well as the kinetics of electron transfer [6971]. The 8-coordinate, tetranionic coordination environment depresses the cerium potential by 8001000 mV relative to the non-aqueous cerium nitrates. This stabilization of ceruim(IV) is not as extensive as that provided by the harder alkoxide ligands, which display greater localization of charge. These data suggest that the cerium(IV) ion is stabilized more effectively by ligands with donor atoms of high charge-density. The hydroxypyridinonate ligands, HOPO-1, HOPO-2 and HOPO3, Fig. 2, studied by Raymond and co-workers under aqueous conditions have also been analyzed in acetontrile. Under the nonaqueous conditions the order in which the ligands modulate the Ce(III/IV) potential is different from aqueous solvent and this is attributed to variations in solvation energies between the Ce(III) and Ce(IV) complexes of the three ligands [34]. 3.4. Complexes of tetrapyrrole and other tetraaza macrocycles The cerium(IV) ion is well-accommodated in the eight coordinate environment provided by a sandwich of two equivalents

Fig. 6. Structures of selected alkoxide- and aryloxide-supported cerium complexes with known electrochemistry [6567].

28

N.A. Piro et al. / Coordination Chemistry Reviews 260 (2014) 2136

Fig. 8. Tetraazo macrocycles.

of tetrapyrrole macrocycles such as porphyrins, azaporphyrins, and phthalocyanines, Fig. 8. In these compounds, assigning a distinct Ce(III/IV) redox event is often complicated by the presence ligand redox activity and hole delocalization, in which the formally Ce(IV) complexes have partial Ce(III) character with a radical electron hole in the ligand system, and exist in intermediatevalent ground states [72]. For the formally tetravalent, neutral sandwich compounds a more accurate description of the valence is [CeX (L(n2) )2 ], where X is a valence between III and IV, and n represents the fraction of the electron-hole delocalized over each ligands system. The intermediate valence phenomenon

cannot be assigned on the basis of electrochemistry alone and is typically modeled on spectroscopy data including cerium LIII edge X-ray absorption or photoelectron spectroscopies. In compounds of this type, the rst reduction event is most often ascribed to a formally cerium redox event by comparison to the electrochemical properties of non-redox active lanthanides [73]. The bisporphyrin compounds are regarded as corresponding to true cerium(IV) complexes with minimal oxidation of the porphyrin ligand in the ground state of the neutral compounds. This assignment is supported by the absence of spectroscopic signatures

N.A. Piro et al. / Coordination Chemistry Reviews 260 (2014) 2136 Table 6 Reduction potentials of cerium tetranitrogen macrocycle complexes. Complex Ce2 (OEP)3 Ce(Pc)(TPyP) (X = 3.68) Ce(TPP)Pc Ce(Nc)(TBPP) Ce(TAP)(Pc) Ce(NctBu)2 (trivalent) Ce(TClP) 2 Ce(TPP)(Pc-1) Ce(Pc-5)2 (X = 3.59) Ce(Nc)(OEP) (X = 3.68) Ce(Pc-12)2 Ce(TPP)2 Ce(Pc-8)2 Ce(OEP)(Pc) Ce(TPP)2 Ce(TTP)2 Ce(TAP)2 Ce(Por-Por) Ce(OPTAP)2 Ce(OEP)(TPP) Ce(dap)(OEP) Ce(HBP)2 Ce(OEP)2 Ce(OEP)2 Ce(omtaa)2
a b

29

Potential (V) vs. SCE 0.76, 0.91 0.10 0.05 0.07 0.10 0.11 0.11 0.145 0.18 0.19 0.20 0.21 0.21 0.26 0.27 0.29 0.30 0.30 to 0.31 Epc = 0.68 Epa = 0.18 0.44 0.51 (1st)b 0.87 (2nd)b 0.52 0.56 0.58 1.1

Reported potential (V) 0.85, 0.90 vs Ag/AgCl 0.16 vs. Ag/AgCl, sat KCl 1.24, 0.74 vs. Fc/Fc+ 0.51 vs. Ag/AgCl, sat KCl 1.7 vs. Fc/Fc+

Conditionsa 0.1 M TBAPF6 in DCM 0.1 M TBAP in DCM TBAP in DCM 0.1 M TBAP in DCM TBAP in DCM 0.1 M TBAP in DCM TBAPF6 in DCM TBAP in DCM 0.1 M TBAP in DCM 0.1 M TBAP in DCM 0.1 M TBAP in DCM 0.1 M TBAP in DCM 0.1 M TBAP in DCM TBAP in DCM TBAPF6 in DCM TBAPF6 in DCM TBAPF6 in DCM TBAP in DCM 0.1 M TBAPF6 in DCM TBAPF6 in DCM 0.1 M TBAP in DCM TBAPF6 in DCM 0.1 M TBAP in DCM TBAPF6 in DCM 0.1 M TPAB in THF

Ref. [84] [73] [78] [73] [78] [73] [75] [78] [73] [73] [73] [74] [73] [78] [75] [75] [75] [76] [79] [75] [77] [75] [74] [75] [87]

TBAP = [NBu4 ][ClO4 ], TBAPF6 = [NBu4 ][PF6 ], = [NEt4 ][ClO4 ], TPAB = [N(n-C3 H7 )4 ][B(3,5-(CF3 )2 C6 H3 )4 ]. The authors of assign the second reduction event to cerium, but the rst observed reduction is more consistent with potentials observed for the related compounds in this table.

for the porphyrin radical monoanion [7476]. Electrochemical data for the one-electron reduction of several of these complexes are presented in Table 6. Mixed porphyrin/diazaporphyrin, porphyrin/phthalocyanine, and bisporphyrazine complexes are also typically regarded as containing Ce(IV) ions, and the one-electron reduction events of these complexes are also included in Table 6 [7779]. From the data in Table 6 it is evident that the tetranionic nitrogen-rich coordination environments do not provide

comparable stabilization of cerium(IV) to the basic anionic oxygen donors of Section 3.3. In contrast to the porphyrin sandwich complexes, bisphthalocyanine and naphthalocyanine complexes, Fig. 9, are described as having appreciable cerium(III) character [73,77,8083]. These complex are nevertheless included in the Table 6, and a value for the valence, X, is given in cases where it was determined experimentally by XANES spectroscopy [73].

Fig. 9. Structure of Ce(Pc)2 . Coordinates from Ref. [86].

30

N.A. Piro et al. / Coordination Chemistry Reviews 260 (2014) 2136

The triple decker sandwich Ce2 (OEP)3 has also been characterized electrochemically, and in comparison to the europium analog, cerium oxidation events for each ion were identied at +0.76 and +0.91 V [84]. Related cerium triple-decker complexes could also be oxidized, but the metal centered events were not identied [85]. The presence of only 1.5 porphyrin rings per metal ion in these complexes and the associated increase in charge at the metal centers have a signicant effect on the ease with which the metal is oxidized, as these potentials are greater than 1.3 V more oxidizing than that for Ce(OEP)2 . The triple decker sandwiches Ce2 (TPP)(Pc)2 and Ce2 Pc(TPP)2 have also been studied electrochemically and show behavior that is complicated by phthalocyanine oxidation processes [78]. Cerium complexes of other nitrogen macrocycles such as dibenzotetramethylaza[14]annulene (tmtaa) and 5,14-dihydro2,3,6,8,11,12,15,17-octamethyldibenzo[b,i][1,4,8,11]tetraazacyclotetradecine (omtaa) have also been isolated. The bis(omtaa) complex, Ce(omtaa)2 , has been characterized electrochemically and displays, to our knowledge, the most negative reversible cerium(III/IV) redox event reported to date, at 1.1 V versus SCE. The omtaa ligand with its anionic charges more localized than in the tetrapyrrole complexes (due to the saddle-shaped geometry of the non-planar omtaa macrocycle) is evidently a very effective environment for stabilizing the cerium(IV) ion. 3.5. Organometallic complexes of cerium and endohedral fullerene complexes The cyclopentadienide ligand plays a prominent role in organo-transition metal chemistry, where it is considered an electron-rich donor to a variety of metals. Cyclopentadienide also supports f-block metals, where more than two of the units often coordinate to the large, highly charged cations. The sodium tetrakis(cyclopentadienyl)cerium(III) ate-complex is reportedly accessible, but shows irregular electrochemistry [88]. The cerium alkoxide complex Cp3 Ce(Oi Pr) has been prepared and fully characterized and shows a reversible reduction event at +0.31 V in THF/TBABF4 , Table 7 [89]. This result is particularly telling about the electronics of the Cp ligand within the Ce framework; the more delocalized charge across the Cp ring seems to less efciently transfer charge to the cerium center, and results in Cp3 Ce(Oi Pr) behaving as a mild oxidant with potentials comparable to that of Fc+ . Cerocene, Ce(C8 H8 )2 , Fig. 10, is a particularly interesting organocerium molecule. Cerocene has been studied extensively since its rst reliable report in 1976 [9497]. While formally a Ce(IV) ion sandwiched between two cyclooctatetraene dianions, the true electronic conguration is complex and best described as an intermediate valent cerium ion: an electron hole is shared jointly by the cerium ion and the two C8 H8 rings. The ground state wavefunction of cerocene is best described as an admixture of a cerium(III) and cerium(IV) electronic congurations [72,93,98]. As such, any description of a Ce(III/IV) redox couple for this molecule is particularly imprecise. Nevertheless, reversible redox chemistry is observed for cerocene; the rst reduction event is observed at 0.8 V, which comes at a signicantly negative potential relative
Table 7 Reduction potentials of organometallic cerium complexes. Complex CeSc2 N@C80 CeLu2 N@C80 CeY2 N@C80 Cp3 Ce(Oi Pr) Ce(C8 Me6 )2 Ce(C8 H8 )2
a

Fig. 10. Structure of Ce(C8 H8 )2 . Coordinates from Ref. [100].

to the formal Ce(III/IV) couple. An electronically related complex, bis(permethylpentalene)cerium is also best described with an intermediate valent ground state, however, the reversible electrochemical reduction occurs more readily by 530 mV compared with cerocene, Table 6 [92,99]. Endohedral fullerene complexes that contain cerium ions have also been prepared and thoroughly characterized electrochemically. Interestingly, a series of trimetallic compounds, M2 CeN@C80 (M = Sc, Y, Lu), displayed reversible redox events that were assigned as Ce(III/IV) oxidations (Table 7) [90,91]. The trimetallic nitride cluster shows a unique strain-driven tunability of the redox couple; the identity of the redox-inactive metal results in a 400 mV potential range of the cerium oxidation and trends with their ionic radii. Larger M cations show greater pyramidalization of the CeM2 N unit and shorter Ce N bond distances, which results in a more electron-rich cerium center and more negative oxidation potentials. Notably, other cerium containing endohedral fullerenes, including Ce3 N@C80 , only display cage oxidation events up to the solvent window and do not show assignable Ce(III/IV) redox events [101106]. 3.6. Cerium amides Cerium amides, in particular Ce[N(SiMe3 )2 ]3 , have a welldeveloped synthetic chemistry, but studies of their electrochemistry are rare [107110]. To our knowledge, the reduction potential of only one amide-supported cerium complex has been reported, CeCl[N(SiMe3 )2 ]3 . This lone data point, presented in Table 8, suggests that amides are not as well-suited to stabilizing cerium(IV) as

Potential (V) vs. SCE 0.89 0.57 0.49 0.31 0.27 0.8

Reported potential (V) 0.33 vs. Fc/Fc+ 0.01 vs. Fc/Fc+ 0.07 vs. Fc/Fc+ 0.32 vs. SCE0.01 vs Fc/Fc+ 0.83 vs. Fc/Fc+ 1.4 vs. Fc/Fc+

Conditionsa TBABF4 in o-DCB TBABF4 in o-DCB TBABF4 in o-DCB 0.25 M TBABF4 in THF unspecied 0.1 M TBAPF6 in THF

Ref. [90] [90] [91] [89] [92] [93]

TBABF4 = [NBu4 ][BF4 ], TBAPF6 = [NBu4 ][PF6 ].

N.A. Piro et al. / Coordination Chemistry Reviews 260 (2014) 2136 Table 8 Reduction potentials of an amide supported cerium complex. Complex CeCl[N(SiMe3 )2 ]3
a

31

Potential (V) vs. SCE 0.26

Reported potential (V) 0.30 vs. Fc/Fc


+

Conditionsa 0.1 M TPAB in THF

Ref. [67]

TPAB = [N(n-C3 H7 )4 ][B(3,5-(CF3 )2 C6 H3 )4 ].

are alkoxide or aryloxide ligands, although studies with complexes of directly comparable coordination numbers at the cerium cation are currently lacking. 4. Preparative chemistry of cerium(IV) complexes by oxidation reactions The development of the chemistry of cerium(IV) is an active area of research [27,110]. Two synthetic strategies have been used to access complexes of cerium(IV). The rst is to start with a cerium(IV) precursor, such as CeO2 , [CeCl6 ]2 or CAN, and proceed through ligand substitution reactions. This strategy has been applied successfully, however, preservation of the tetravalent oxidation state can be challenging due to the reducing nature of many deprotonated ligands [89,111,112]. The successes of this approach have been reviewed recently and will not be discussed in detail [27]. A second method is to access cerium(IV) complexes through one-electron oxidation of cerium(III) complexes. As illustrated over the previous sections, the cerium(III/IV) potentials are highly ligand set dependent; the strategy of oxidizing cerium(III) complexes to stable cerium(IV) complexes becomes applicable with ligands that are good donors and depress the cerium potential into the range of common synthetic laboratory oxidants. Below, we survey the reported cerium(III) oxidations by reaction type: (1) outer-sphere oxidation, H-atom abstraction, and autooxidation, (2) oxidative functionalization, and (3) oxidation-induced ligand redistribution reactions. 4.1. Outer-sphere oxidation, H-atom abstraction, and autooxidation reactions The cerium(III) to cerium(IV) transformation can occur spontaneously with oxidants as simple as dioxygen. From a preparative perspective, this oxidation occurs during the introduction of ligands in order to isolate cerium(IV) products from cerium(III) starting materials. A derivative of TrenSal is oxidized to cerium(IV) by simple stirring in the presence of O2 , Scheme 1 [113]. A Zn complex of the fully hydrogenated ligand H6TrenSal, when treated with cerium(III) nitrate in the presence of air affords a bimetallic cerium complex with a bridging 2 ;2 , 2 -O2 unit. It has been proposed that this complex is mixed-valent and contains a superoxo bridge, though this has not been conrmed [114]. One peculiar example involves a silesquioxane ligand, which spontaneously yielded cerium(IV) products from cerium(III) precursors. The authors contest that adventitious O2 or peroxides from ethereal solvents do not

serve as the oxidant, although, the identity of the oxidant remains unclear [115]. Many of the formally cerium(IV) complexes of porphyrin and porphyrin-like ligands of Section 3.4 are synthesized from cerium(III) acetylacetonate complexes by treatment with the ligand, its dianion, or its condensation precursors, and air. This strategy is effective for isolation of numerous compounds of this class [116]. The cerium(III) complex Ce(Htmtaa)(tmtaa) undergoes oxidation with O2 , benzoquinone, or ferrocenium with loss of a proton to afford the Ce(tmtaa)2 complex [117]. In the case of the related Ce(Homtaa)(omtaa) complex, the reaction with O2 occurs in a single-crystal to single-crystal transformation to afford the cerium(IV) product [87]. In solution, TEMPO also oxidizes Ce(Homtaa)(omtaa) through a formal H-atom transfer reaction. Net outer-sphere oxidations of reducing cerium(III) compounds will also generate cerium(IV) complexes. Cerocene and its derivatives are often accessed from the oxidation of the cerium(III) bis(cyclooctatetraene) anions with oxidants such as AgI or allyl bromide [93,118,119]. The phenolate complex [Li(THF)2 CeIII (MBP)2 (THF)2 ] is oxidized with a range of oxidants CuCl2 , CuBr2 , CuCl, CuI, I2 , NBS, Fc+ forming a mixture of the neutral Ce(MBP)2 (THF)2 and the -ate complexes [Li(THF)CeX(MBP)2 (THF)], X = Cl, Br, I [66]. 4.2. Oxidative functionalization Oxidative functionalization is the simultaneous oxidation of a metal center and introduction of new ligands to the coordination sphere. In the case of cerium(III) complexes, oxidative functionalization is another method of preparing cerium(IV) complexes. For example, Ce[N(SiMe3 )2 ]3 reacts with chlorine atom donors such as TeCl4 , PhICl2 , or Ph3 CCl to afford ClCe[N(SiMe3 )2 ]3 , Fig. 11 [109,120,121,115,126128]. Inner-sphere oxidations that form a strong Ce X bond concomitant with electron-transfer, and do not require either release of high-energy radicals (e.g. Cl ) or termolecular reactions, are crucial to successful synthetic schemes. That is to say that the Ce X bond should form during the initial oxidation, and that the electron be deposited into a low energy intermediate. The use of PhICl2 , which ts these requirements, was shown to be a general reagent for the synthesis of a variety of Ce(IV) complexes, Fig. 12 [109]. By conformationally constraining cerium(III) employing a TREN-based ligand, Ce[N(C2 H4 NSiMe2 t Bu)3 ], Ce(III) can be

Scheme 1. Synthesis of CeCl(TrenSal). Scheme adapted from [113].

32

N.A. Piro et al. / Coordination Chemistry Reviews 260 (2014) 2136

Fig. 11. A comparison of oxidants for the oxidative halogenation of Ce[N(SiMe3 )2 ]3 [109,120,121].

oxidized by I2 to afford the only known Ce(IV) iodide complex. In contrast, bulky monodentate Ce(III) amides such as Ce[N(SiMe3 )2 ]3 and Ce[TMP]3 (TMP = 2,2,6,6-tetramethylpiperidinide) are not oxidized with iodine [122,123]. Additionally, neither Cl2 , Br2 , I2 nor Ag+ is capable of oxidizing Ce[N(SiMe3 )2 ]3 , despite several of these being thermodynamically better oxidants than reagents that have been successfully applied. These results suggest a kinetic barrier to accessing Ce(IV) complexes and are discussed in further detail in Section 5. Interestingly, oxidation of Ce[N(C2 H4 NSiMe2 t Bu)3 ] in pentane with either Cl2 or Br2 in place of I2 does not facilitate the isolation of monomeric [CeIV ]X complexes, instead the mixed-valent species [CeIV ] X [CeIII ] were obtained, where [Ce] = CeN(C2 H4 NSiMe2 t Bu)3 and X = Cl , Br [122]. Scott and coworkers proposed that disproportionation of [CeIV ] X to [CeIV ] X [CeIII ] and X2 was favored for the harder Lewis bases Cl and Br , and suggested that stronger dative interactions for CeIV X CeIII offset the loss of bond enthalpy associated with formation of CeIV X. Recently, trityl chloride has been reported as an effective chloride group transfer oxidant, with amide, alkoxide, and aryloxide supporting ligand frameworks [67,120]. In these cases, homoleptic and heteroleptic complexes are oxidized to their corresponding Ce(IV) chloro complexes without any ligand redistribution being observed. Typically, different oxidants have been investigated in cerium(III) oxidation chemistry in order to change reactivity preferences; however, three of us have recently reported that the choice of alkali metal can result in different oxidation reactivity within a conserved ligand framework [67]. Oxidation of [M3 (THF)n ][(BINOLate)3 Ce], M = Li, Na, K, with trityl chloride yields either the product of oxidative functionalization [Li3 (THF)5 ][(BINOLate)3 CeCl] or the salt-eliminated products [M2 (THF)n ][(BINOLate)3 Ce], M = Na, K [67].

Oxidants that form Ce O bonds are also amenable to the synthesis of well-dened Ce(IV) products. Ce(OCt Bu3 )3 is oxidized by a variety of organic peroxides as well as quinones to form tetravalent products with new Ce O bonds, Fig. 13 [124]. Benzoquinone, BQ, results in the formation of a dimeric cerium(IV) product, whereas use of a bulky quinone results in monomeric Ce(IV) semiquinolates. The use of BQ as an oxidant for Ce(III) is surprising due to Ce(III) being a mild reductant and benzoquinone a mild oxidant. However, coordination of a Lewis basic oxidant to a strong Lewis acid can result in large Lewis Acid Promoted Potential Shifts (LAPPS). Recently, BQ has been used as an oxidant for [Li3 (THF)5 ][(BINOLate)3 Ce(THF)], which forms an authenticated cerium(IV) dimer bridged by the doubly reduced hydroquinolate [125]. The presence of the fully reduced hydroquinone represents an overall shift of BQs oxidizing potential by 1.6 V, demonstrating the potential utility of Lewis acid coordination in RE oxidation chemistry. 4.3. Oxidation-induced ligand redistribution reactions Under oxidizing conditions, cerium complexes will often undergo ligand redistribution reactions to afford rearranged products. -Diketonate complexes of cerium have been synthesized by metallating with cerium(III) precursors to give tris(ligand) complexes, which under aerobic conditions redistribute to afford cerium(IV) products [126128]. That these conditions afford the cerium(IV) products in good yields is a testament to the stability afforded the cerium(IV) ion by the -diketonate ligand eld. The aerobic oxidation of Ce(NCy2 )3 (THF) is also used to afford homoleptic Ce(NCy2 )4 in 35% yield in a ligand redistribution [129]. The homoleptic Ce(IV) complex Ce[N(SiHMe2 )2 ]4 is similarly formed through a redistribution process in the reaction of Ce[N(SiHMe2 )2 ]3 (THF)x (x = 0, 2) with PhICl2 ,

Fig. 12. Use of PhICl2 for the formation of Ce(IV) Cl bonds. Scheme adapted from Ref. [109].

N.A. Piro et al. / Coordination Chemistry Reviews 260 (2014) 2136

33

Fig. 13. Oxygencerium bond-forming oxidation reactions [124,125].

5. The role of kinetics in cerium electrochemistry and oxidation reactions The existence and stability of strongly oxidizing, molecular cerium(IV) complexes such as ceric ammonium nitrate (CAN) in the presence of moisture has been ascribed not only to the large overpotential for water oxidation but also to kinetic inertness of the cerium(IV) compounds in the solid state [133]. For example, CAN features a saturated CeIV coordination sphere and highly hydrogen bonded secondary coordination sphere, which leads to a kinetically stable and robust reagent. Ceric ammonium uorides also demonstrate a unique kinetic inertness; (NH4 )2 CeF6 is stable to heat and moisture, whereas M2 CeF6 and M3 CeF7 are only isolable in the absence of water [133]. Slow electron transfer rates have also been implicated as a detriment to the preparation of cerium(IV) compounds [131]. An interesting example is the oxidation chemistry of the Ce[N(SiMe3 )2 ]3 (Fig. 16). Oxidation reactions of this homoleptic amide have proven challenging; oxidants that might be predicted to oxidize the amide on the basis of thermodynamics have otherwise shown no reaction [121,122]. This disagreement between thermodynamic predictions and observed reactivity suggests a kinetic aspect, such as ligand reorganization, might be a mitigating factor in oxidation reactions of cerium(III) complexes. This is further supported by the observation that a conformationally-constrained

Fig. 14. Oxidationrearrangement reactions to afford homoleptic amide complexes [108,129,130].

trityl chloride, or hexachloroethane [108]. A similar redistribution process is observed in the oxidation of the tris(dithiocarbamate) complex Ce(2 -S2 CNEt2 )3 with O2 , which afforded the cerium(IV) tetrakis(dithiocarbamate) complex, Fig. 14 [130]. Benzoquinone can also act as a net outer-sphere oxidant. The cerium tris(alkoxide-carbene) complex shown in Fig. 15 undergoes oxidation with benzoquinone, XeF2 , or ferrocenium to afford the ligand distributed product [131]. Notably, this oxidation is not afforded by TeCl4 , PBr2 Ph3 or I2 . Oxygen has also been used as an oxidant to give various ceriumoxo clusters originating from tris(amide) complexes [132]. In these clusters, oxygen incorporation is coupled with ligand redistribution to give a variety of products that have been identied by X-ray crystallography.

Fig. 15. Oxidative ligand redistribution of an alkoxide/carbene complex. [Ox] = benzoquinone, XeF2 , or [Fc][OTf]. Scheme adapted from [131].

34

N.A. Piro et al. / Coordination Chemistry Reviews 260 (2014) 2136

Fig. 16. Oxidants used to oxidize Ce[N(SiMe3 )2 ]3 grouped by thermodynamically favorable and unfavorable couples relative to that observed for ClCe[N(SiMe3 )2 ]3 . Oxidants with accessible potentials that are unreactive are highlighted (red boxes).

cerium tris(amide) can be oxidized by halogens, whereas the monodentate tris(amides) are not [122,123]. Investigations of the kinetics of the CeIII/IV couple have focused largely on CeIV in acidic media, due to the application of cerium(IV) compounds in batteries, fuel cells, and separations chemistry [134144]. In non-aqueous media there have been few studies of the kinetics of the CeIII/IV couple [67,71]. Matsumoto and coworkers investigated the electrochemical and chemical kinetics associated with the CeIII/IV couple for Ce(acac)4 0/ by cyclic voltammetry and cross reactions of outer sphere electron transfer (ET) reagents. These studies determined the CeIII/IV self-exchange rate (kex ) using the Marcus cross relation. As expected, values of kex depend on the driving force ( G ) of the cross-reaction partner, where larger free energy differences resulted in closer agreement to calculated values of kex . Similar behavior has been observed for EuII/III under aqueous conditions, and has been attributed to poor electronic coupling (HAB ) between donor and acceptor for ET [145]. Physically, the origin for this poor electronic coupling is the shielding of the radially contracted 4f orbitals by outer shell orbitals, resulting in slower rates of electron transfer. The authors suggest that contributions from the product excited state with larger driving forces serve to overcome the poor electronic coupling; values for kex are in agreement with those calculated using the Marcus cross relation, and suggest a change from a concerted to a directional ET process (Fig. 17). More recently, three of us have investigated the role of ligand reorganization in CeIII oxidation chemistry [67]. We studied the electrochemical, Fig. 18, and chemical oxidation behavior for a

Fig. 18. Normalized cyclic voltammograms of [M3 (THF)n ][(BINOLate)3 Ce], taken at 100 mV/s in 0.1 M [N(n-Pr)4 ][BArF 4 ]THF. Adapted from Ref. [67].

heterobimetallic framework; we found that the rates of heterogeneous electron transfer (ks ) run counter to rates of the chemical oxidations with trityl chloride (kobs ). An increase in the driving force, G , and associated increase in the rate constant ks were observed as the Lewis acidity of the secondary cation M+ decreased (K+ > Na+ > Li+ ), whereas the rate of chemical oxidation kobs followed an opposite trend corresponding to the accessibility of the CeIII ion to Lewis base coordination (Li+ > > Na+ > K+ ), Table 9. The opposing trends in ks and kobs support different mechanisms for the electrochemical and chemical processes, and indicate an inner sphere mechanism for the chemical oxidation of the complexes with trityl chloride. These ndings highlight the importance of ligand reorganization; the Ce/Li framework readily reorganizes to allow access to both six and seven-coordinate geometries, which are essential for coordination of the oxidant and oxidation reaction. When reorganization between six- and seven-coordinate
Table 9 Kinetics Data for [M3 (THF)n ][(BINOLate)3 Ce] (M = Li, Na, K) [67]. Entry 1 2 3 Alkali metal, M K Na Li ks [ 104 cm s1 ]a 9.45 5.97 4.84 kobs [ 104 s1 ]b 0.216 1.93 30.0

Fig. 17. Depiction of non-adiabatic direct ET and directional ET processes following Marcus theory. Scheme adapted from Ref. [71].

a Obtained by cyclic voltammetry measurements in THF using 0.1 M TPAB at 100 mV/s. b Obtained under pseudo-rst order conditions using UVvis absorption spectroscopy;[Ce]/[Ph3 CCl] = 1:10.

N.A. Piro et al. / Coordination Chemistry Reviews 260 (2014) 2136

35

geometries becomes high in energy, sluggish reaction kinetics are observed and alternate lower energy pathways become available. For heterobimetallic or -ate complexes, salt elimination facilitates a non-destructive pathway for oxidation reactions; however, in bulky or non-chelating ligand frameworks, ligand redistribution can occur leading to unpredictable products and/or low yields of cerium(IV) compounds of interest. 6. Conclusions. As evident from the data reported in this review, the Ce(III/IV) redox couple is thermodynamically and kinetically highly sensitive to its ligand environment. Under aqueous, acidic conditions with poor donor ligands the cerium(III) form is strongly favored, an aqueous oxidation potential of +1.63 V was reported for Ce(IV) ions in 8 M HClO4 . In general, the cerium(IV) state is favored by basic conditions with higher coordination numbers of effective donors, especially anionic oxygen donor ligands. This is exemplied by Ce(L5 )(Ot Bu)2 , Fig. 6, a cerium(IV) compound that is reduced, albeit irreversibly, at 1.83 V in non-aqueous conditions. Thus, the reported cerium redox events span a nominal range of 3.5 V. Importantly, ligands that are traditionally associated with stabilizing higher oxidation state transition metals and actinides are not necessarily as well suited for stabilizing cerium(IV); the predominantly ionic bonding eliminates signicant contributions from bonding or other covalent interactions. As such, cyclopentadienyl and amide frameworks do not transfer charge to the cerium ion as effectively as the more localized charge of alkoxide ligands. We expect these data will be useful for the expansion of molecular chemistry of cerium(IV), the development of a new reductive chemistry of cerium(III), and the rational design of cerium complexes for tailored redox applications. It is clear that cerium redox chemistry has great potential for new fundamental and applied chemistry. Acknowledgments E.J.S. gratefully acknowledges the U.S. Department of Energy, Ofce of Science, Early Career Research Program (DE-SC0006518) for support. E.J.S. and P.J.W. also thank the University of Pennsylvania and the NSF (CHE-1026553) for support. We are also grateful to the reviewers for their highly constructive assessment and helpful suggestions for this review. References
[1] F.A. Cotton, G. Wilkinson, Advanced Inorganic Chemistry, John Wiley & Sons, 1988, pp. 957. [2] R.J. Ellis, M.R. Antonio, ChemPlusChem 77 (2012) 4147. [3] Y. Wei, M. Kumagai, Y. Takashima, M. Asou, T. Namba, K. Suzuki, A. Maekawa, S. Ohe, J. Nucl. Sci. Technol. 35 (1998) 357364. [4] M.-H. Chiang, W.W. Clayton, L. Soderholm, M.R. Antonio, Eur. J. Inorg. Chem 2003 (2003) 26632669. [5] W.C. Chueh, C. Falter, M. Abbott, D. Scipio, P. Furler, S.M. Haile, A. Steinfeld, Science 330 (2010) 17971801. [6] A. Trovareli, in: G.J. Huthings (Ed.), Catalytic Science, Imperial College Press, London, 2002. [7] G.A. Molander, Chem. Rev. 92 (1992) 2968. [8] V. Nair, L. Balagopal, R. Rajan, J. Mathew, Acc. Chem. Res. 37 (2004) 2130. [9] K. Binnemans, Applications of Tetravalent Cerium Compounds, in: K.A. Gscheidner Jr., J.-C.G. Bnzli, V.K. Pecharsky (Eds.), Handbook on the Physics and Chemistry of Rare Earths, Elsevier, 2006, pp. 281392. [10] J.R. Hwu, K.-Y. King, Current Sci. 81 (2001) 10431053. [11] W.J. Evans, Coord. Chem. Rev. 206 (2000) 263283. [12] M. Szostak, D.J. Procter, Angew. Chem. Int. Ed. 51 (2012) 92389256. [13] T.A. Beineke, J. Delgaudio, Inorg. Chem. 7 (1968) 715721. [14] R.G. Pearson, J. Am. Chem. Soc. 85 (1963) 35333539. [15] S. Ahrland, J. Chatt, N.R. Davies, Q Rev. Chem. Soc. 12 (1958) 265276. [16] D.A. Atwood, The Rare Earth Elements: Fundamentals and Applications, in: R.A. Scott (Ed.), Encyclopedia of Inorganic and Bioinorganic Chemistry, John Wiley & Sons, Chichester, 2012. [17] R.D. Shannon, Acta Cryst. A32 (1976) 751767.

[18] P.B. Hitchcock, A.G. Hulkes, M.F. Lappert, Inorg. Chem. 43 (2004) 10311038. [19] D. Chang, T. Malinski, A. Ulman, K.M. Kadish, Inorg. Chem. 23 (1984) 817824. [20] A.J. Bard, L.R. Faulkner, Electrochemical Methods: Fundamentals and Applications, John Wiley & Sons, Inc., New York, 2001. [21] N. Connelly, W. Geiger, Chem. Rev. 96 (1996) 877910. [22] A.M. Bond, E.A. McLennan, R.S. Stojanovic, F.G. Thomas, Anal. Chem. 59 (1987) 28532860. [23] D.J.G. Ives, G.J. Janz, Reference Electrodes: Theory and Practice, Academic Press, New York, 1961. [24] A.H. Kunz, J. Am. Chem. Soc. 53 (1931) 98102. [25] A.A. Noyes, C.S. Garner, J. Am. Chem. Soc. 58 (1936) 12651268. [26] G.F. Smith, C.A. Getz, Ind. Eng. Chem. Anal. Ed. 10 (1938) 191195. [27] F.M.A. Sroor, F.T. Edelmann, Tetravalent Chemistry Inorganic, in: D.A. Atwood (Ed.), The Rare Earth Elements: Fundamentals and Applications, John Wiley & Sons, Chichester, 2012, pp. 313320. [28] B.A. Bilal, E. Muller, Z. Naturforsch. A. 47 (1992) 974984. [29] A.W. Maverick, Q. Yao, Inorg. Chem. 32 (1993) 56265628. [30] A. Abbaspour, M. Mehrgardi, Talanta 67 (2005) 579584. [31] M.A. Brown, A. Paulenova, A.V. Gelis, Inorg. Chem. 51 (2012) 77417748. [32] Y. Suzuki, T. Nankawa, A.J. Francis, T. Ohunki, Radiochim. Acta 98 (2010) 397402. [33] A. DAlo, L. Toupet, S. Rigaut, C. Andraud, O. Maury, Opt. Mater. 30 (2008) 16821688. [34] J. Xu, E. Radkov, M. Ziegler, K.N. Raymond, Inorg. Chem. 39 (2000) 41564164. [35] G.J.P. Deblonde, M. Sturzbecher-Hoehne, R.J. Abergel, Inorg. Chem. 52 (2013) 88058811. [36] D.E. Morris, D.E. Hobart, Lanthanide Actinide Res. 2 (1987) 91103. [37] S.F. Haddad, K.N. Raymond, Inorg. Chim. Acta 122 (1986) 111118. [38] S.R. Sofen, S.R. Cooper, K.N. Raymond, Inorg. Chem. 18 (1979) 16111616. [39] M.R. Antonio, C.W. Williams, L. Soderholm, Radiochim, Acta 90 (2002) 851856. [40] G. Vasquez, D.G. Kolman, J. Electrochem. Soc. 155 (2008) C571C577. [41] L. De Almeida, S. Grandjean, N. Vigier, F. Patisson, Eur. J. Inorg. Chem. 2012 (2012) 49864999. [42] G. Szigethy, J. Xu, A.E.V. Gorden, S.J. Teat, D.K. Shuh, K.N. Raymond, Eur. J. Inorg. Chem. 2008 (2008) 21432147. [43] M.R. Antonio, L. Soderholm, C.W. Williams, N. Ullah, L.C. Francesconi, J. Chem. Soc. Dalton Trans. (1999) 38253830. [44] R.D. Peacock, T.J.R. Weakley, Chem. Commun. (1971) 1836. [45] N. Haraguchi, Y. Okaue, T. Isobe, Y. Matsuda, Inorg. Chem. 33 (1994) 10151020. [46] K. Binnemans, Chem. Rev. 107 (2007) 25922614. [47] M. Armand, F. Endres, D.R. MacFarlane, H. Ohno, B. Scrosati, Nat. Mater. 8 (2009) 621629. [48] N.V. Plechkova, K.R. Seddon, Chem. Soc. Rev. 37 (2008) 123150. [49] F. van Rantwijk, R.A. Sheldon, Chem. Rev. 107 (2007) 27572785. [50] J.H. Davis, Chem. Lett. 33 (2004) 10721077. [51] R. Sheldon, Chem. Commun. (2001) 23992407. [52] T. Welton, Chem. Rev. 99 (1999) 20712083. [53] H. Ohno, Electrochemical Aspects of Ionic Liquids, Wiley & Sons Inc., Hoboken, NJ, 2005. [54] L.-H. Chou, W.E. Cleland Jr., C.L. Hussey, Inorg. Chem. 51 (2012) 1145011457. [55] F.M. Lin, C.L. Hussey, J. Electrochem. Soc. 140 (1993) 30933096. [56] C.J. Rao, K.A. Venkatesan, K. Nagarajan, T.G. Srinivasan, P.R.V. Rao, J. Nucl. Mater. 399 (2010) 8186. [57] A.M. Arif, C.J. Gray, F. Alan Hart, M.B. Hursthouse, Inorg. Chim. Acta 109 (1985) 179183. [58] R. Bandin, R. Bastida, A. de Blas, P. Castro, D.E. Fenton, A. Macias, A. Rodriguez, T. Rodriguez, Chem. Commun. (1994) 11851188. [59] J. Rohovec, P. Vojt sek, P. Hermann, J. Ludvk, I. Luke s, J. Chem. Soc. Dalton Trans. (2000) 141148. [60] P.-Y. Jiang, Y. Ikeda, M. Kumagai, J. Nucl. Sci. Technol. 31 (1994) 491493. [61] D. Pletcher, E.M. Valdes, Electrochim. Acta 33 (1988) 499507. [62] H. Zheng, S.J. Yoo, E. Mnck, L. Que, J. Am. Chem. Soc. 122 (2000) 37893790. [63] D.W. Wester, G.J. Palenik, R.C. Palenik, Inorg. Chem. 24 (1985) 44354437. [64] S. Zolezzi, E. Spodine, A. Decinti, Polyhedron 21 (2002) 5559. [65] E.M. Broderick, P.S. Thuy-Boun, N. Guo, C.S. Vogel, J. Sutter, J.T. Miller, K. Meyer, P.L. Diaconescu, Inorg. Chem. 50 (2011) 28702877. [66] B.D. Mahoney, N.A. Piro, P.J. Carroll, E.J. Schelter, Inorg. Chem. (2013). [67] J.R. Robinson, P.J. Carroll, P.J. Walsh, E.J. Schelter, Angew. Chem. Int. Ed. 51 (2012) 1015910163. [68] B.L. Vallee, R.J. Williams, Proc. Natl Acad. Sci. USA 59 (1968) 498505. [69] A. Job, P. Goissedet, Comptes Rendus 157 (1913) 5052. [70] T. Behrsing, A.M. Bond, G.B. Deacon, C.M. Forsyth, M. Forsyth, K.J. Kamble, B.W. Skelton, A.H. White, Inorg. Chim. Acta 352 (2003) 229237. [71] M. Matsumoto, H. Kodama, S. Funahashi, H.D. Takagi, Inorg. React. Mech. 2 (2000) 1931. [72] C.H. Booth, M.D. Walter, M. Daniel, W.W. Lukens, R.A. Andersen, Phys. Rev. Lett. 95 (2005) 267202.

36

N.A. Piro et al. / Coordination Chemistry Reviews 260 (2014) 2136 [107] D. Bradley, J. Ghotra, F. Hart, J. Chem. Soc. Dalton Trans. (1973) 10211027. [108] A.R. Crozier, A.M. Bienfait, C. Maichle-Mssmer, K.W. Trnroos, R. Anwander, Chem. Commun (2013). [109] P.P. Drse, A.R.A. Crozier, S.S. Lashkari, J.J. Gottfriedsen, S.S. Blaurock, C.G.C. Hrib, C.C. Maichle-Mssmer, C.C. Schdle, R.R. Anwander, F.T.F. Edelmann, J. Am. Chem. Soc. 132 (2010) 1404614047. [110] F.M.A. Sroor, F.T. Edelmann, Tetravalent Chemistry Organometallic, in: D.A. Atwood (Ed.), The Rare Earth Elements: Fundamentals and Applications, John Wiley & Sons, Chichester, 2012, pp. 321334. [111] P.L. Arnold, I.J. Casely, S. Zlatogorsky, C. Wilson, Helvet. Chim. Acta 92 (2009) 22912303. [112] A.F. England, Chemistry, Massachusetts University of Technology, 1995, PhD Thesis. [113] P.P. Drse, J. Gottfriedsen, Z. Anorg. Allg. Chem. 634 (2008) 87. [114] A. Mustapha, J. Reglinski, A.R. Kennedy, Inorg. Chim. Acta 362 (2009) 12671274. [115] Y.K. Gunko, R. Reilly, F.T. Edelmann, H.-G. Schmidt, Angew. Chem. Int. Ed. 40 (2001) 12791281. [116] J. Jiang, D.K. Ng, Acc. Chem. Res. 42 (2009) 7988. [117] M.D. Walter, R. Fandos, R.A. Andersen, New J. Chem. 30 (2006) 1065. [118] U. Kilimann, R. Herbst-Irmer, D. Stalke, F.T. Edelmann, Angew. Chem. Int. Ed. 33 (1994) 16181621. [119] V. Lorenz, B.M. Schmiege, C.G. Hrib, J.W. Ziller, A. Edelmann, S. Blaurock, W.J. Evans, F.T. Edelmann, J. Am. Chem. Soc. 133 (2011) 12571259. [120] P.L. Arnold, Z.R. Turner, N. Kaltsoyannis, P. Pelekanaki, R.M. Bellabarba, R.P. Tooze, Chem. Eur. J. 16 (2010) 96239629. [121] O. Eisenstein, P.B. Hitchcock, A.G. Hulkes, M.F. Lappert, L. Maron, Chem. Commun. (2001) 15601561. [122] C. Morton, N.W. Alcock, M.R. Lees, I.J. Munslow, C.J. Sanders, P. Scott, J. Am. Chem. Soc. 121 (1999) 1125511256. [123] P.B. Hitchcock, Q.-G. Huang, M.F. Lappert, X.-H. Wei, J. Mater. Chem. 14 (2004) 3266. [124] A. Sen, H.A. Stecher, A.L. Rheingold, Inorg. Chem. 31 (1992) 473479. [125] J.R. Robinson, C.H. Booth, P.J. Carroll, P.J. Walsh, E.J. Schelter, Chem. Eur. J. 19 (2013) 59966004. [126] T.J. Pinnavaia, R.C. Fay, Inorg. Synth. 12 (1970) 7780. [127] M. Ciampolini, F. Mani, N. Nardi, J. Chem. Soc. Dalton Trans. (1977) 1325. [128] M. DelaRosa, K. Bousman, J. Welch, P. Toscano, J. Coord. Chem. 55 (2002) 781793. [129] P.B. Hitchcock, M.F. Lappert, A.V. Protchenko, Chem. Commun. (2006) 3546. [130] P.B. Hitchcock, A.G. Hulkes, M.F. Lappert, Z. Li, Dalton Trans (2004) 129136. [131] I.J. Casely, S.T. Liddle, A.J. Blake, C. Wilson, P.L. Arnold, Chem. Commun. (2007) 50375039. [132] M.P. Coles, P.B. Hitchcock, A.V. Khvostov, M.F. Lappert, Z. Li, A.V. Protchenko, Dalton Trans. 39 (2010) 6780. [133] S. Cotton, Lanthanide and Actinide Chemistry, John Wiley & Sons Ltd, West Sussex, England, 2006. [134] T.S. Chen, K.L. Huang, K.J.C. Yeh, J. Electrochem. Soc. 156 (2009) E69E74. [135] P.K. Leung, C.P. de Leon, C.T.J. Low, F.C. Walsh, Electrochim. Acta 56 (2011) 21452153. [136] Y. Liu, X. Xia, H. Liu, J. Power Sources 130 (2004) 299305. [137] P. Modiba, M. Matoetoe, A.M. Crouch, J. Power Sources 205 (2012) 19. [138] P. Modiba, A.M. Crouch, J. Appl. Electrochem. 38 (2008) 12931299. [139] F.J. Xiong, D.B. Zhou, Z.P. Xie, Y.Y. Chen, Appl. Energ. 99 (2012) 291296. [140] Z.P. Xie, D.B. Zhou, F.J. Xiong, S.M. Zhang, K.L. Huang, J. Rare Earths 29 (2011) 567573. [141] A. Paulenova, S.E. Creager, J.D. Navratil, Y. Wei, J. Power Sources 109 (2002) 431438. [142] M.H. Chakrabarti, S.A. Hajimolana, F.S. Mjalli, M. Saleem, I. Mustafa, Arabian. J. Sci. Eng. 38 (2013) 723739. [143] M. Skyllas-Kazacos, M.H. Chakrabarti, S.A. Hajimolana, F.S. Mjalli, M. Saleem, J. Electrochem. Soc. 158 (2011) R55R79. [144] P. Leung, X.H. Li, C.P. de Leon, L. Berlouis, C.T.J. Low, F.C. Walsh, R. Soc. Chem. Adv. 2 (2012) 1012510156. [145] V. Balzani, F. Scandola, G. Orlandi, N. Sabbatini, M.T. Indelli, J. Am. Chem. Soc. 103 (1981) 33703378.

[73] Y. Bian, J. Jiang, Y. Tao, M.T.M. Choi, R. Li, A.C.H. Ng, P. Zhu, N. Pan, X. Sun, D.P. Arnold, Z.-Y. Zhou, H.-W. Li, T.C.W. Mak, D.K.P. Ng, J. Am. Chem. Soc. 125 (2003) 1225712267. [74] R.J. Donohoe, J.K. Duchowski, D.F. Bocian, J. Am. Chem. Soc. 110 (1988) 61196124. [75] J.W. Buchler, A. De Cian, J. Fischer, P. Hammerschmitt, J. Lfer, B. Scharbert, R. Weiss, Chem. Ber. 122 (1989) 22192228. [76] J.W. Buchler, T. Dippell, Eur. J. Inorg. Chem. 1998 (1998) 445449. [77] N. Pan, Y. Bian, M. Yokoyama, R. Li, T. Fukuda, S. Neya, J. Jiang, N. Kobayashi, Chem. Ber. 2008 (2008) 55195523. [78] T.-H. Tran-Thi, T.A. Mattioli, D. Chabach, A. De Cian, R. Weiss, J. Phys. Chem. 98 (1994) 82798288. [79] A. Garrido Montalban, S.L.J. Michel, S.M. Baum, B.J. Vesper, A.J.P. White, D.J. Williams, A.G.M. Barrett, B.M. Hoffman, J. Chem. Soc. Dalton Trans. (2001) 32693273. [80] H. Isago, M. Shimoda, Chem. Lett. (1992) 147150. [81] P. Zhu, F. Lu, N. Pan, D.P. Arnold, S. Zhang, J. Jiang, Eur. J. Inorg. Chem. (2004) 510517. [82] S.L. Selector, V.V. Arslanov, Y.G. Gorbunova, O.A. Raitman, L.S. Sheinina, K.P. Birin, A.Y. Tsivadze, J. Porphyrins Phthalocyanines 12 (2008) 11541162. [83] J. Jiang, D.K. Ng, Acc. Chem. Res. 42 (2008) 7988. [84] J.K. Duchowski, D.F. Bocian, J. Am. Chem. Soc. 112 (1990) 88078811. [85] H. Miwa, N. Kobayashi, K. Ban, K. Ohta, Bull. Chem. Soc. Jpn. 72 (1999) 27192728. [86] M.S. Haghighi, C.R.L. Teske, H. Honborg, Z. Anorg. Allg. Chem. 608 (1992) 7380. [87] U.J. Williams, B.D. Mahoney, A.J. Lewis, P.T. DeGregorio, P.J. Carroll, E.J. Schelter, Inorg. Chem. 52 (2013) 41424144. [88] K. Jacob, M. Glanz, K. Tittes, K.H. Thiele, Z. Anorg. Allg. Chem. 556 (1988) 170178. [89] A. Gulino, M. Casarin, V.P. Conticello, J.G. Gaudiello, H. Mauermann, I. Fragala, T.J. Marks, Organometallics 7 (1988) 23602364. [90] Y. Zhang, S. Schiemenz, A.A. Popov, L. Dunsch, J. Phys. Chem. Lett. 4 (2013) 24042409. [91] L. Zhang, A.A. Popov, S. Yang, S. Klod, P. Rapta, L. Dunsch, Phys. Chem. Chem. Phys. 12 (2010) 7840. [92] A. Ashley, G. Balzs, A. Cowley, J. Green, C.H. Booth, D. OHare, Chem. Commun. (2007) 15151517. [93] A. Streitwieser, S.A. Kinsley, C.H. Jenson, J.T. Rigsbee, Organometallics 23 (2004) 51695175. [94] A. Greco, S. Cesca, W. Bertolini, J. Organomet. Chem. 113 (1976) 321330. [95] N.M. Edelstein, P.G. Allen, J.J. Bucher, D.K. Shuh, C.D. Soeld, N. Kaltsoyannis, G.H. Maunder, M.R. Russo, A. Sella, J. Am. Chem. Soc. 118 (1996) 1311513116. [96] M.D. Walter, C.H. Booth, W.W. Lukens, R.A. Andersen, Organometallics 28 (2009) 698707. [97] R. Coates, PhD Thesis, University College London, 2010. [98] P. Fulde, Electron Correlations in Molecules and Solids, 3 ed., Springer, New York, 1995. [99] The claim by Ashley et al. that Ce(C8 Me6 )2 and Ce(C8 H8 )2 have nearly identical reduction potentials was a misstament due to a confusion regarding the reference with respect to which the cerocene potential was reported. Streitwieser et al. reports a value of 1.4 V vs. Fc/Fc+, not 800 mV. [100] E.Kurras, C.Kruger, Private Communication to the Cambridge Structure Database. Identier YARLEJ. (2004). [101] M.N. Chaur, F. Melin, B. Elliott, A. Kumbhar, A.J. Athans, L. Echegoyen, Chem. Eur. J. 14 (2008) 45944599. [102] M.N. Chaur, R. Valencia, A. Rodrguez-Fortea, J.M. Poblet, L. Echegoyen, Angew. Chem. Int. Ed. 48 (2009) 14251428. [103] T. Suzuki, K. Kikuchi, F. Oguri, Y. Nakao, S. Suzuki, Y. Achiba, K. Yamamoto, H. Funasaka, T. Takahashi, J. Molec. Cat. A 52 (1996) 49734982. [104] M. Yamada, T. Nakahodo, T. Wakahara, T. Tsuchiya, Y. Maeda, T. Akasaka, M. Kako, K. Yoza, E. Horn, N. Mizorogi, K. Kobayashi, S. Nagase, J. Am. Chem. Soc. 127 (2005) 1457014571. [105] M. Yamada, T. Wakahara, T. Tsuchiya, Y. Maeda, T. Akasaka, N. Mizorogi, S. Nagase, J. Phys. Chem. A 112 (2008) 76277631. [106] M. Yamada, N. Mizorogi, T. Tsuchiya, T. Akasaka, S. Nagase, Chem. Eur. J 15 (2009) 94869493.

Anda mungkin juga menyukai