a Possible Electrode - American Chemical Society

Mar 25, 2015 - and Jean-Marie Tarascon*. ,†,‡,§. †. FRE 3677 “Chimie du Solide et Energie”, Collège de France, 11 Place Marcelin Berthelot...
0 downloads 0 Views 4MB Size
Subscriber access provided by SUNY DOWNSTATE

Article

Li2Cu2O(SO4)2: a possible electrode for sustainable Libased batteries showing a 4.7 V redox activity vs. Li+/Li0 Meiling Sun, Gwenaelle Rousse, Artem M. Abakumov, Matthieu Saubanère, Marie-Liesse Doublet, Juan Jose Rodriguez-Carvajal, Gustaaf Van Tendeloo, and Jean-Marie Tarascon Chem. Mater., Just Accepted Manuscript • Publication Date (Web): 25 Mar 2015 Downloaded from http://pubs.acs.org on March 26, 2015

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

Chemistry of Materials is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 23

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

Li2Cu2O(SO4)2: a possible electrode for sustainable Li-based batteries showing a 4.7 V redox activity vs. Li+/Li0 Meiling Suna, Gwenaëlle Roussea,b,c, Artem M. Abakumovd, Matthieu Saubanèrec,e, Marie-Liesse Doubletc,e, Juan Rodríguez-Carvajalf, Gustaaf Van Tendelood and Jean-Marie Tarascona,b,c * a

FRE 3677 “Chimie du Solide et Energie”, Collège de France, 11 Place Marcelin Berthelot, 75231 Paris Cedex 05, France b

Sorbonne Universités - UPMC Univ Paris 06, 4 Place Jussieu, F-75005 Paris, France.

c

Réseau sur le Stockage Electrochimique de l’Energie (RS2E), FR CNRS 3459, France

d

EMAT, University of Antwerp, Groenenborgerlaan 171, B-2020 Antwerp, Belgium

e

ICGM Institut Charles Gerhardt - CNRS and Université Montpellier, Place Eugène Bataillon, 34095 Montpellier, France f

Institut Laue Langevin, 71 avenue des Martyrs, 38042 Grenoble, France

ABSTRACT: Li-ion batteries rely on the use of insertion positive electrodes with performances scaling with the redox potential of the 3d-metals accompanying Li-uptake/removal. Although not commonly studied, the Cu2+/Cu3+ redox potential has been predicted from theoretical calculations to possibly offer a high operating voltage redox couple. We herein report the synthesis and crystal structure of a hitherto-unknown oxysulfate phase, Li2Cu2O(SO4)2, which contains infinite edge-sharing CuO4 chains and presents attractive electrochemical redox activity with respect to Li+/Li, namely amphoteric characteristics. Li2Cu2O(SO4)2 shows redox activity at 4.7 V vs. Li+/Li0 corresponding to the oxidation of Cu2+ to Cu3+ enlisting ligand holes and associated with the reversible uptake-removal of 0.3 Li. Upon reduction, this compound reversibly uptakes ~2 Li at an average potential of about 2.5 V vs. Li+/Li0, associated with the Cu2+/Cu+ redox couple. The mechanism of the reactivity upon reduction is discussed in detail, with particular attention to the occasional appearance of an oscillation wave in the discharge profile. Our work demonstrates that Cu-based compounds can indeed be fertile scientific ground in the search for new high energy density electrodes.

Introduction Li-ion batteries (LIB) have become the premier energy storage technology for portable electronics and vehicle electrification and stand as serious contenders for grid applications.1 The energy density delivered by Li-ion batteries, presently based on a carbon negative electrode, is mainly controlled by the voltage and capacity of the positive electrode materials. Today’s research on such materials is divided into two approaches: one prioritizing the design of substituted layered oxide materials having outstanding energy densities (270 mAh/g);2–4 the other one favoring the search for new polyanionic compounds which provide safety and cost advantages at the expense of energy density, which is sacrificed due to the dead weight of polyanions.5,6 Although both have led to rich discoveries, this work focuses only on the latter. Among the polyanionic compounds, LiFePO4 has become the most lauded material for the next generation of batteries to power electric vehicles.5 However, a remaining drawback for this material is its low redox voltage of 3.45 V vs. Li+/Li°, which has motivated the design of Febased polyanionic compounds having greater potentials while preserving cost, environmental and safety benefits. In the search for such materials, our group proposed,

based on the inductive effect concept,7 the replacement of the (PO4)3- polyanion in LiFePO4 by a more electronegative one (SO4)2- together with injection of the highly electronegative F- species into the structure. A new family of electrochemically active compounds termed as 3d-metal fluorosulfates (formula AxMSO4F) emerged.8–13 Among them, the triplite LiFeSO4F phase shows the highest potential thus far reported for any Fe-based inorganic compounds.6,14 Since then, many other fluorosulfates with A being either Li, Na or K ions and M being either Ni, Co, Mn or Zn were obtained and this rich crystal chemistry was extended to the preparation of numerous 3d-metal hydroxysulfates (LiFeSO4OH),15,16 and F-free sulfates (Li2Fe(SO4)2).17– 19 Further extension of the work led us to synthesize for the first time the Fe-based oxysulfate Fe2O(SO4)2, which shows a redox potential of 3 V vs. Li+/Li0 20 but whose lithiated end-member Li2Fe2O(SO4)2 is moisture sensitive, hence making its use complicated in Li-ion batteries. This was an impetus to look for related Li-based oxysulfates having greater potentials. Along that line, as deduced from DFT calculations,21,22 the 3d-metal elements from the right side of the periodic table (Co, Ni, Cu) were the most interesting. Although the sulfate crystal chemistry is rich, until now we could successfully prepare only one sulfate

ACS Paragon Plus Environment

Chemistry of Materials

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

phase combining an alkali metal (Na) and Cu, NaCuSO4F.10 Nevertheless, we decided to pursue Cu compounds because out the aforementioned metal candidates, Cu is the only one for which an oxysulfate does exist: the mineral Cu2OSO4, called dolerophanite.23–25 We report here the synthesis, structure and electrochemistry of a new Li-based copper oxysulfate phase, Li2Cu2O(SO4)2, which crystallizes in a tetragonal structure and reacts reversibly towards Li in oxidation at 4.7 V and in reduction at 2.5 V vs. Li+/Li°. Peculiar behavior is observed in reduction, specifically the appearance of oscillation waves in the discharge voltage-composition curve. Results Synthesis: Li2SO4 (Alfa Aesar, 99.7%), Li2O (Alfa Aesar, 99.5%) and CuSO4 (Alfa Aesar, Reagent grade) were used as Li and Cu-based precursors for the targeted phase with the latter being treated at 260°C under vacuum for one day to ensure its dryness. Two different routes were used. The first one consists in reacting Cu2OSO4, made by annealing CuSO4 under air at 725°C,24 with stoichiometric amounts of Li2SO4 according to the following reaction Cu2OSO4 + Li2SO4 → Li2Cu2O(SO4)2 The powders were thoroughly ball milled, pressed into a pellet and annealed under Ar for 5 days at 500°C to produce a well crystalized sample with sharp reflections on the XRD powder pattern reminiscent of a new phase as described next with, however, minute amounts of Li2SO4. To simplify the process, we tried to bypass the use of Cu2OSO4, which is time consuming to prepare. An alternative and successful single step route to Li2Cu2O(SO4)2 was identified: Li2O + 2 CuSO4 → Li2Cu2O(SO4)2 The reaction was conducted by ball milling for one hour Li2O and dried CuSO4 powders into an Ar-filled ballmill jar with a stainless-steel ball/powder weight ratio of 20. The milled powders were placed in an alumina boat and annealed at 400°C for 4 days under Ar. The recovered emerald-green powders displayed an X-ray diffraction pattern identical to samples obtained via Cu2OSO4 and with no sign of impurity phases. Structural Characterization: The X-ray diffraction (XRD) pattern of Li2Cu2O(SO4)2 was recorded in BraggBrentano geometry with a Bruker D8 Advance diffractometer equipped with a copper source (λCu-Kα1 = 1.54056 Å, λCu-Kα2 = 1.54439 Å) and a LynxEye detector. The pattern contains several peaks whose width and shape indicate a highly crystallized sample as the peak width corresponds to the instrumental one (absence of sample broadening). All diffraction peaks can be indexed using the Dicvol program26 in a tetragonal unit cell, with lattice parameters a=8.32586(2) Å and c=5.091351(17) Å. The corresponding volume (V= 352.933(2) Å3) is suitable to accommodate two formulae per unit cell. The unit cell was confirmed by transmission electron microscopy (TEM) investigation. For the TEM study, the sample was prepared in an Ar-filled glove box by crushing

Page 2 of 23

the grainy powder in a mortar in anhydrous hexane and depositing drops of suspension onto holey carbon grids. The sample was transported to the microscope column completely excluding contact with air. EDX analysis revealed the Cu:S atomic ratio of 50(1):49(1), in good agreement with the Li2Cu2O(SO4)2 formula (Supporting information, Figure S1) which was independently confirmed by atomic absorption analysis as the ratio of Li to Cu and S was found to nicely match the nominal composition. Electron diffraction (ED) patterns and energy dispersive X-ray (EDX) spectra were obtained with a Tecnai G2 electron microscope operated at 200 kV. All obtained ED patterns can be indexed on a primitive tetragonal unit cell with approximate lattice parameters a≈8.3 Å, c≈5.1 Å (Figure 1), thus confirming at first sight the unit cell found by XRD. However, inspection of the [001] ED pattern shows that the angle between the a* and b* reciprocal lattice axes slightly deviates from 90°; this may reveal a small monoclinic distortion or result from a damage caused by the electron beam (the latter is supported by synchrotron X-ray diffraction experiments). Indeed, the material demonstrates very high sensitivity to the electron beam irradiation, which makes high resolution TEM imaging virtually impossible. The only reflection condition was found to be 00l: l = 2n. In the ED patterns in Figure 1, this reflection condition is not visible due to multiple diffraction, but it was confirmed by vanishing the 00l, l ≠ 2n reflections upon tilting the crystal around the c* reciprocal lattice axis. The possible space groups are P42/m, P42 or, assuming a possible monoclinic distortion, P21/m, P21. Investigation of a possible monoclinic distortion was performed using high resolution synchrotron X-ray diffraction. The finely ground powder was loaded in a 0.5 mm diameter capillary and the sample was measured in transmission mode (λ=0.41374 Å) on the 11-BM beamline at Argonne National Lab (Figure 2a). Structural determination was undertaken from the synchrotron data using direct methods with the EXPO software.27 Absolutely no splitting of the reflections was observed, so the determination was conducted in the P42/m space group, and lead to a pertinent atomic distribution of Cu, S and O in the unit cell. As lithium atoms are weakly scattered by X-rays because of their small number of electrons, we used neutron powder diffraction to locate them. The sample (from the exact same batch) was put in a cylindrical vanadium container and measured on the HRPT neutron diffractometer at SINQ-PSI (Villigen, Switzerland) with a wavelength of 1.4928 Å.28 The partial “Cu2O(SO4)2” model deduced from XRD was implemented into the neutron diffraction pattern and difference Fourier maps were calculated to locate Li atoms using the GFourier program available in the FullProf suite of programs.29 Negative peaks (the neutron scattering length of Li is negative: bLi= –1.90 fm) can be seen in the y=1/2 section of the Fourier map (Figure 2b, right), revealing two lithium positions Li1 and Li2 corresponding to the 2d and 2f Wyckoff sites, respectively. The final structural model was obtained by a combined Rietveld refinement30 of the neutron and Synchro-

ACS Paragon Plus Environment

Page 3 of 23

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

tron XRD patterns conducted with the FullProf program,29,31 and the result is shown in Figure 2. The structural information and atomic positions within the tetragonal P42/m unit cell are compiled in Table 1. The four Cu and S atoms are placed on the Wyckoff site 4j, while oxygen is distributed among four sites. O1, O2 and O3 define the SO4 tetrahedra while O4 is on the 2e position. Figure 3 presents the structure of Li2Cu2O(SO4)2. Cu is surrounded by four oxygen atoms forming an almost perfect square planar environment (two Cu-O distances at 1.92 and two at 2.01 Å). A fifth oxygen (O2) is found at a longer distance of 2.32 Å, so as to form a square-based CuO5 pyramid. As a square planar environment is commonly expected for Jahn-Teller Cu2+ ions (electronic configuration d9), the structure is from here on described as such. The CuO4 squares are linked through edges to form Cu2O6 dimers, and are repeated with a 90° rotation under the effect of the 42 screw axis. This gives rise to infinite chains running along [001]. O4, the only oxygen not being part of SO4 tetrahedra, constitutes the link of the chain. The SO4 groups are positioned around the chains, with two ligands pointing towards Li2 giving rise to tetrahedral coordination. Li1 atoms are located in an octahedral environment, built on four oxygens linked to Cu (4 x O1) and two oxygens from a sulfate group (2 x O2). In order to confirm the refined structure, a bond valence sum analysis was performed using the Zachariasen formula Vi=Σjsij=Σexp{(d0-dij)/0.37} with the d0 parameters characterizing cation-anion pairs taken from reference 32. Results are indicated in Table 1 and the obtained valence sums are in perfect agreement with oxidation states of +1, +2, +6 and –2 for Li, Cu, S and O respectively. Meanwhile, the thermal stability of Li2Cu2O(SO4)2 was determined by thermogravimetric analysis (TGA) coupled with mass spectrometry. The TGA profiles collected by heating the compound either in air or argon indicate a decomposition temperature of 580°C (Figure 4), above which SO and SO2 were released. This decomposition temperature was confirmed by in situ high temperature XRD measurements (Figure 4, inset) which reveals CuO as the main decomposition product. A prerequisite for Li-based materials to be used as electrodes in Li-ion batteries is the existence of Li-diffusion paths within the structure. To get more information on how Li ions can travel inside the framework, we generated bond valence sum maps (BVSM) and bond valence energy landscapes (BVEL) using the soft bond valence sum parameters developed by S. Adams33. BVEL are obtained from soft-BVS parameters by transforming them valence into energy units, using a Morse-type potential for the attractive part and a screened Coulomb potential for the repulsive part.34 The probing ion, in our case Li+, is placed in all points of the unit cell and its interaction energy, using the above mentioned potential, is calculated up to a common distance of 8 Å with all ions of the structure. The previously existing Li-species are removed from the list of ions before the calculation. This approach has been successfully applied to energy materials and ionic

conductors,35,36 and was recently implemented in the program BondSTR of the FullProf Suite.29 Figure 5 shows the generated BVEL for Li+ conduction in the “LixCu2O(SO4)2” framework. A surface level corresponding to the value of 1.8 eV above the minimum was chosen to plot the BVEL. This corresponds to 0.2 eV above the minimum value leading to an infinitely connected surface for Li transport. It does not come as a surprise that the highlighted Li+ conduction paths meet the Li1 and Li2 positions we determined from neutron powder diffraction, but it also reveals that Li conduction can occur through a three dimensional pathway as shapes around the helical Cu-O chains. This was an impetus to further study the electrochemical properties of this phase as discussed next. Electrochemical Characterization: The electrochemical properties of Li2Cu2O(SO4)2 vs. Li were tested either in oxidation or reduction using Swagelok cells assembled in an argon-filled glove box and cycled in a temperature controlled environment (T=25 ± 0.5°C) using a VMP system (Biologic S.A., Claix, France) operating in galvanostatic mode. Each experiment was repeated at least twice to ensure reproducibility. The working electrode of about 6 to 8 mg of active material was made by ball milling powders of Li2Cu2O(SO4)2 with 20% in mass of carbon SP for 15 minutes in a Spex-800 mixer-miller. The positive electrode was separated from the Li metal disc negative electrode by a Whatman GF/D borosilicate glass fiber sheet saturated with 1M LiPF6 solution in a mixture of ethylene carbonate, propylene carbonate and dimethyl carbonate in 1:1:3 ratio by weight (LP100). Typical voltage– composition curves when the cell was started on oxidation or reduction are shown in Figures 6 and 9, respectively. If not otherwise specified in the figure captions, cells were cycled at C/25 (1 C means uptake or removal of 1 Li+ in one hour). a) Upon initial oxidation, the potential of the Li2Cu2O(SO4)2 cell rapidly increases to 4.5 V and then progressively reaches a plateau located at 4.75 V with the concomitant release of nearly 0.5 Li. Upon discharge, ~0.3 out of the 0.5 Li can be reinserted. This leads to a sustainable reversible capacity (Figure 6) of 0.3 Li per formula unit centered at 4.7 V vs. Li+/Li° which corresponds to ca. 20 mAh/g. Such large irreversibility between the first cycle and the subsequent ones is most likely due to electrolyte decomposition. We note that the irreversible component of the Li2Cu2O(SO4)2/C electrode capacity is dependent on grain size, with the minimum irreversibility achieved when a hand-ground rather than ball milled composite was used as positive electrode (Supporting Information, Figure S2). Attempts to remove more Li by increasing the charge cutoff voltage led to a second oxidation (∼5.1 V) pseudo-plateau. Its origin is mainly due to copious electrolyte decomposition since no extra reversible capacity was found on the subsequent discharge. Moreover, the persistence of the 4.7 V process with similar amplitude whether a 4.8 or 5.2 V charging cutoff voltage was used indicates that the material remains stable at high voltage (Supporting Information, Figure S3). This suggests that more Li might be extracted from this com-

ACS Paragon Plus Environment

Chemistry of Materials

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

pound if electrolytes highly resistant against oxidation could be found. The Li extraction-insertion mechanism in Li2Cu2O(SO4)2 was further examined by in situ X-ray diffraction (Figure 7) using a homemade electrochemical cell equipped with a beryllium window transparent to Xrays.37 During cell charging, no noticeable changes are observed in the X-ray powder pattern until the voltage reaches the plateau at 4.75 V; this supports the notion that the initial measured capacity is mainly due to electrolyte decomposition. Once the plateau is reached, a new set of XRD peaks appear (at a lower angle than those of the pristine phase) and they grow upon further delithiation at the expense of the pristine phase prior to become single phase when the voltage departs from 4.75 V. The oxidized material of composition ~ Li1.7Cu2O(SO4)2, as deduced from both coulometric titration and estimate of the irreversible capacity, is single phase. It crystallizes in a monoclinic structure with lattice parameters a=8.64 Å, b=8.59 Å, c=4.92 Å and β=93.9°. The latter corresponds to a monoclinic distortion of the pristine cell with, however, a volume 3% greater than the di-lithiated phase (364.3 Å3 vs. 352.9 Å3). Unfortunately, absorption by the Beryllium window combined with the broadening of peaks caused by the electrochemical process prevented us from more precise and definite structural determination. Note that the XRD pattern taken at the end of the subsequent discharge superimposes neatly with the pristine one. This is indicative of a fully reversible Li-driven Cu3+/Cu2+ redox process in Li2Cu2O(SO4)2. As an alternative approach to rationalize the electrochemical behavior upon oxidation, we turned to theoretical calculations. Orbital analysis and density functional calculations (DFT) were performed to access a microscopic picture of the electrochemical behavior of this phase. The phase stability diagram obtained for the LixCu2O(SO4)2 system is shown in Figure 8a. It reveals that the complete Cu2+ to Cu3+ oxidation (i.e. removal of 2 Li) would occur at 5.20 V and is strongly disfavored over partial oxidation processes associated with the formation of intermediate phases at x=1.75, 1.5 or even 1. The density of states (DOS) of the Li2Cu2O(SO4)2 phase shows that the band involved in the oxidation process is mainly of oxygen character (Supporting Information, Figure S4), suggesting an anionic rather than a cationic redox activity upon lithium removal. A deeper analysis of the electronic levels lying just below the Fermi level confirms the strong hybridization of the Cu(3d) orbitals with the O(2p) orbitals of the oxygen atoms bridging the CuO4 square-planes. These oxo-ligands (O2-) – being much less electronegative than those belonging to sulfate (SO4)2- groups – can take part to the redox process (Supporting Information, Figure S5). Upon Li removal from the Li2Cu2O(SO4)2 phase, holes are then created on these bridging oxo-ligands so that the oxidation process corresponds to the formation of “Cu2+– O–” species rather than Cu3+–O2–. This is corroborated by atomic charge analysis on Cu, S and O showing that more than 75% of the removed charge (hole) is located on the oxygen (Supporting Information, Figure S4). According to

Page 4 of 23

the phase stability diagram of Figure 8a, oxidation of the Li2Cu2O(SO4)2 electrode should proceed in two steps through the extraction of 0.5 Li each. The average voltages obtained for the x=2 to x=1.5 (4.62 V) and x=1.5 to x=1 (4.91 V) processes compare well with experiments (Supporting Information, Figure S6) although slightly underestimated. The intermediate Li1.5Cu2O(SO4)2 and Li1Cu2O(SO4)2 phases show slightly distorted CuO4 square-planes and a Cu-(O2)-Cu angle deviating from 90° which is consistent with the monoclinic distortion experimentally observed and with the loss of the 42 screw axis (Supporting information, Figure S4). Note that the full oxidation from Li1Cu2O(SO4)2 to Cu2O(SO4)2 is associated with quite a high potential (5.64 V), likely unachievable with current electrolytes. We note that the Cu2O(SO4)2 structure is however very close to that of the starting Li2Cu2O(SO4)2 phase with nearly no volume contraction and a perfectly planar CuO4 configuration. b) To further explore the electrochemical activity of this compound towards Li, a second Li2Cu2O(SO4)2/Li cell was initially started on reduction. The cell voltage rapidly drops to 2.3V and decreases continuously and smoothly to 2.1 V with the concomitant uptake of ~2 Li+ (Figure 9). Then it shows faster voltage decay down to 1.8 V corresponding to the uptake of an extra 0.4 Li+. The 2.4 inserted Li+ can be fully removed when the cell is charged back to 4.6 V (Figure 9, red curve), and presents a voltage profile quite dissimilar from the discharge curve of classical intercalation compounds: it enlists first a pseudo plateau around 3.3 V followed by a second one above 4 V. The reduction process exhibits a single redox peak on discharge as opposed to two on oxidation parted by nearly 1 V, suggesting different reacting paths. Moreover, the voltage profile for the second discharge curve does not superimpose on the first discharge curve, indicative that the compound does not return to its initial state. This difference is highlighted by the large shift of the reduction peaks between the first (solid line) and second (dashed line) derivative curves (Figure 9, bottom). Once this “electrode formation stage” (which progresses through the 3rd cycle) is over, the charge discharge curves neatly superimposed leading to a sustained capacity of 200 mAh/g upon cycling (Figure 9, bottom left inset). By limiting the charge cutoff voltage to 3.9 V we decreased the overall cell capacity to 150 mAh/g but still preserved good capacity retention (Figure 9, bottom right inset). Another interesting feature of the discharge curve is nested in the amount of Li-uptake (2.4 Li per formula unit) which is greater than expected (2 Li per formula unit) from the full reduction of Cu2+ to Cu+. This suggests that a tiny fraction of Cu+ might be reduced to metallic Cu suggesting the feasibility of extruding copper from the lattice via either displacement reactions, as already reported for Cu-based compounds,38 or conversion reactions in various oxides which usually proceed at lower voltages.39,40 While the former preserves the framework of the structure, the latter enlists a complete structural reorganization.

ACS Paragon Plus Environment

Page 5 of 23

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

To explore this last possibility, we cycled a Li2Cu2O(SO4)2/Li cell down to 0.5 V (Supporting Information, Figure S7). Decreasing the voltage from 1.8 to 0.5 V leads to an extra uptake of 2.8 Li, leading to an overall uptake of 5.2 Li per unit formula out of which most of them can be removed on oxidation till 4.6 V. The charge voltage trace is here different from that collected when the cell is only discharged to 1.8 V with the complete cycling voltage profile looking alike those collected for materials enlisting conversion reaction processes. Overall, whatever the cycling conditions, these results highlight the complexity of the reacting mechanism of Li2Cu2O(SO4)2 against Li below 3 V. To gain further insight into the mechanism of this process, in situ X-ray diffraction was performed (Figure 10) A progressive intensity loss without peak-broadening was observed so that the XRD patterns become featureless for x approaching 4, and remain as such on subsequent cycles without any evidence of Bragg peaks corresponding to Cu metal (even when the cell voltage was lowered to 0.5 V). This result does not preclude the formation of Cu, which could precipitate out of the matrix in a nanoscale form not detectable by XRD. Such an occurrence was confirmed by collecting HAADF-STEM images and EDX maps (Figure 11, a-f) on electrodes recovered from cells that were discharged to 1.8 V and 0.5 V, respectively. The data were collected with a Tecnai Osiris electron microscope equipped with a Super-X detector system and operated at 200 kV. Both samples demonstrate the presence of nanoparticles with sizes in the range of 5 – 20 nm. Such particles are more rounded and more numerous in the sample discharged to 0.5 V. EDX compositional mapping indicates that nanoparticles in the 1.8 V sample are oxidized, being most probably a copper oxide, while they are solely copper in the 0.5 V sample. Furthermore, we indirectly observed evidence for the ejection of some copper by taking an XRD pattern of a Li-disc recovered from a cell at the end of the first discharge process (Figure 10, top): the Cu Bragg peaks are clearly present. This usually implies the formation of some Cu+ ions during the electrode lithiation, which are dragged through the negative electrode by the electrical field and reduced into copper metal. To determine the outcome of CuOx and Cu nanoparticles formed during the discharge process to 1.8 V upon subsequent charging, two cells were discharged to 1.8 V and charged to 3.8 V and 4.6 V, respectively. Independent of the charging voltages, ED patterns display only diffraction rings attributed to graphitic carbon, indicative of the formation of an amorphous material (Figure 12a,b). Moreover, HAADF-STEM images do not reveal the presence of either Cu or CuOx oxide nanoparticles as supported by EDX compositional maps that show relatively homogeneous distributions of Cu, S and O (Figure 12c). However, for the 3.8 V sample, small Cu-rich regions are occasionally noted (Supporting Information Figures S8 and S9) with Cu segregation occurring in the form of alternating CuCuOx lamellas. Although such a specific texture could bear some textural/structural hints on how the electrode

is oxidized, we have not further explored these features as their appearance is sporadic. Depending upon the cycling rate, we also note within this system the striking presence of oscillating potentials; this is usually most apparent through the second discharge cycle (Figure 13a), although bursts of potential changes are also observed on the 1st, 3rd and subsequent reduction cycles and with similar amplitudes. The oscillating nature of the observed potential, which we ascertain to not be due to temperature fluctuations, is nicely confirmed by magnifying various portions of GITT measurements data (Figure 13b). From zoom 1, the onset of an increase in voltage (decrease in polarization) vs. time just 0.5 hours after applying the current pulse, together with a decrease of voltage (increase of polarization) after 1 hour, is reminiscent of the “periodic” behavior observed in the cycling curves. This tendency proceeds through the discharging process (zoom 2) and it is only at the end of the reduction (zoom 3) that the behavior of the potential under current becomes typical, with a monotonic increase of the polarization (decrease of the potential) with time. The origin of this oscillation which has also been noticed in the galvanostatic reduction of a Cu2OSO425 electrode vs. Li+/Li0, but not explained, will be addressed in the discussion section. Discussions-Conclusions In this report, we figured out the structure of a novel synthetized phase Li2Cu2O(SO4)2, described its electrochemical reactivity towards Li and demonstrated its amphoteric nature: the compound can either be oxidized or reduced enlisting respectively the Cu3+/Cu2+ and Cu2+/Cu+ redox couples. On charge, Cu2+ is oxidized to Cu3+ at 4.7 V vs. Li/Li0, one of the highest ever reported potential for such couple in any Cu-based inorganic compound, confirming, as predicted by DFT, the feasibility to achieve high redox values with Cu-based polyanionic compounds: a 5.09 V redox couple was indeed estimated for compounds containing Cu, (SO4)2- and fluorine.22 In fact, the inductive effect of the sulfate group in Li2Cu2O(SO4)2 is mitigated by the presence of the oxo-group and the absence of highly electronegative F–, resulting in the voltage to be lower than 5.09 V. Moreover, the strong hybridization of the metallic 3d-orbitals with the oxo-ligand’s 2porbitals is responsible for Cu(3d)/O(2p) band overlap, which is shown to promote anionic redox activity through the creation of holes on the bridging oxygen of the CuO4 square planes. In short Cu3+ can be viewed as Cu2+ hole, in analogy to what was found for the high Tc cuprates having Cu2+ in square planar environments. This finding suggests that the Cu-based chemistry could be fertile ground to expand the family of materials showing anionic network redox activity beyond the Li-rich lamellar oxides.4 The practical advantage offered by such a high potential is unfortunately counterbalanced by the inability to remove more than 0.3 Li per formula unit. Removing this amount of Li from the Li1 and Li2 positions destabilizes the structure, which undergoes a monoclinic distortion as

ACS Paragon Plus Environment

Chemistry of Materials

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

observed at the end of charging (4.8 V). Moreover, Li acts as a screening agent between chains: when Li is removed the electrostatic repulsion between adjacent SO4 tetrahedra is increased, which is why the volume of the unit cell increases upon (partial) delithiation. This counterintuitive effect (volume increase on charge) has been previously observed in layered materials such as LiCoO241 and also the copper oxide Li2CuO2.42 In addition to battery applications, and in light of the high Tc cuprates, Cu3+-based compounds might present superconductivity; however, as deduced from low field (20 Oe) magnetic measurements, no sign of superconductivity was found for the Li1.7Cu2O(SO4)2 phase down to 2 K (Supporting Information, Figure S10). Unlike the oxidation of Li2Cu2O(SO4)2, which basically obeys a classical deinsertion-insertion process, its reduction is more complex due to intriguing features such as i) the non-symmetric voltage trace between charge and discharge, ii) non-superimposable voltage profiles between the first and second discharges, iii) the ability to insert an amount of Li greater than allowed from solely the reduction of Cu2+ to Cu+ (2 per formula unit), with this amount increasing drastically to 5 Li as the voltage is lowered to 0.5 V, and iv) the appearance of Cu-O and Cu nanoparticles as the discharge cutoff voltage is lowered. Only a few reaction mechanisms can account for such large amounts of Li uptake. The first mechanism, mainly specific to Cu-based compounds, is a Li-driven displacement reaction leading to the growth and disappearance of metallic Cu. In the presence of Cu2+, the displacement reaction will enlist a two-electron process (Cu2+→ Cu+ → Cu0), that is the uptake of 2 Li (4 per unit formula) to fully extrude metallic copper from the matrix. The likelihood of such a process occurring here is at first sight improbable as it cannot account for the appearance of CuO in as observed by electron microscopy. An alternative and more likely explanation which integrates all the experimental facts is nested in the feasibility of our compound to entails a conversion reaction which can be described by the following reactions which occur sequentially: Li2Cu2O(SO4)2 + 2 Li+ + 2e− → 2 Li2SO4 + Cu2O Cu2O + 2 Li+ + 2 e− → Li2O + 2 Cu

(1)

(2)

We believe that Li2Cu2O(SO4)2 reacts with Li via a process involving uptake of Li by the host structure which induces the reduction of Cu2+ and Cu+. Since Cu+ usually prefers a different oxygen coordination (square planar or elongated octahedron for Cu2+ vs. dumbbell O-Cu+-O as in Cu2O43 or Cu4O344), such structural hindrance might result in a progressive decomposition of the parent phase into CuOx (reaction 1), leading to an amorphous composite. The Cu oxide then transforms, via reaction 2 which is well documented in the literature, into metallic Cu whose amount increases as the discharge potential is lowered to 0.5 V. This is consistent with the discharge voltage curve which departs from the 2.1 V plateau when 2 Li have reacted as well as with the major formation of CuOx with

Page 6 of 23

slight amount of Cu metal in the 1.8 V discharged sample and of solely Cu in the 0.5 V discharged one (Figure 11). Upon the subsequent charge, the two reactions proceed in opposite direction. This reversibility is again consistent with the fact that CuO and Cu nanoparticles are disappearing when the cell is recharged to 3.8 and 4.5 V, respectively as deduced by microscopy. Although, at the moment it is premature to speculate in greater details about the way CuOx is re-injected into the “Li-Cu-S-O” amorphous phase: we should just recall that the CuO crystal structure shows nearly equivalent edge-shared CuO4 square-planes as those occurring in Li2Cu2O(SO4)2. To confirm this mechanism, a DFT+U phase stability diagram was computed for the LixCu2O(SO4)2 system (2 ≤ x ≤ 6) assuming different electrochemical reactions, namely insertion (black circles), displacement (black triangles) and conversion reactions (blue and red squares). It is given in Figure 8b where the reference energy (green horizontal line) corresponds to the reaction enthalpy of the direct conversion of Li2Cu2O(SO4)2 into 2Cu0 + Li2O + 2Li2SO4 with the exchange of 4Li. Displacement and insertion (electrodes, namely {Li2+xCu20 x/2O(SO4)2 + x/2Cu } and Li2+xCu2O(SO4)2, respectively are found much higher in energy than the conversion electrodes whatever is the lithium content. Reactions (1) and (2) are clearly the thermodynamically favored mechanism (blue squares) which supports our mechanistic hypothesis. We note however, that a second mechanism (red squares) involving the formation of an intermediate {CuO + Cu0 + 2Li2SO4} electrode could compete with reaction (1). Although thermodynamically less favored than reactions (1)+(2) this mechanism is consistent with the appearance of CuO particles in the discharge samples and could be kinetically activated by the formation of a CuO phase which is closely related to the initial Li2Cu2O(SO4)2 phase. Lastly, on the basis of our overall proposed reaction scheme (Supporting Information, Figure S6), we believe the oscillation phenomenon seen on discharge to originate from competing kinetics between the interconnected electrochemical and chemical extrusion-disproportion processes, which yields a blocking intermediate phase that forms and decomposes repetitively, thereby impeding charge transfer and locally modifying the electrode internal resistance. Obviously, further experiments are needed to locally probe this unique feature, whose unicity might rely in compounds enlisting both an oxide injection with its subsequent reduction. In summary, we have revealed a new environmentally benign and easily synthesized compound that displays capacities of 20 mAh/g and 200 mAh/g at high (~4.7 V) and low (~2.5 V) average voltage. While the highpotential process entails a classical biphasic reversible insertion de-insertion reaction, the low-potential process proceeds via a complex displacement-conversion reaction which displays an interesting oscillation phenomenon. Obviously direct extensions of this work range from i) mastering the sustained reversibility of the 3 V process, which until now was not pursued beyond 40 cycles, and

ACS Paragon Plus Environment

Page 7 of 23

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

ii) finding ways to extract greater amounts of Li+ at high potential to fully take advantage of the theoretical energy density offered by this new compound at 4.7 V, which equals 800 Wh/kg. Finally, we hope this work will propel a more intense search for Cu-based polyanionic compounds capable of supporting the anion redox activitydriven Cu3+/Cu2+ couple, a transition metal that has been unduly disregarded. Additionally, Li2Cu2O(SO4)2 could provide a basis for better understanding S=1/2 frustrated magnetic chains, and may be of interest for the discovery of new superconductors.

SUPPORTING INFORMATION CIF file of Li2Cu2O(SO4)2. EDX analysis. HDAAF-STEM images at different states of charge/discharge. Additional electrochemical curves and DFT calculations. Magnetic susceptibility. This material is available free of charge via the Internet at http://pubs.acs.org.

AUTHOR INFORMATION Corresponding Author * Corresponding author: [email protected].

spectrometry revealing SO and SO2 gas formation above 600°C (black and red lines). The inset shows the XRD powder patterns of Li2Cu2O(SO4)2 recorded in air between 300°C and 650°C. Note the appearance of CuO from 580°C which becomes single phase above 600°C (red diamond symbols). Figure 5: Bond valence energy landscape deduced for Li2Cu2O(SO4)2. The green domains correspond to the migration paths for Li in the structure, obtained using an isosurface value of 1.8 eV over the minimum energy. Figure 6: Voltage versus composition curve of a Li2Cu2O(SO4)2/Li cell started on oxidation at C/25. The first, second and third cycles are shown in black, red and blue respectively. In each case the corresponding capacity retention (closed triangles=capacity on charge, open circles: capacity on discharge) together with the derivative curves for the second cycle are reported. Figure 7: Behavior of a Li2Cu2O(SO4)2/Li cell on charge at C/25. Left: in situ XRD patterns recorded behind a Be window while the cell is charged to 5 V (right curve). The red patterns correspond to the pristine phase, the blue one is the pattern of the fully oxidized Li1.7Cu2O(SO4)2, while the black ones correspond to biphasic domains. The pattern on top was recorded at 3.6 V at the subsequent discharge, and indicates the good reversibility of the process.

ACKNOWLEDGMENT We thank Y. Chabre and S.L. Zhang for many fruitful discussions, V. Pumjakushin for his help in neutron diffraction, D. Giaume for her help in ICP measurements, M. Courty for TGA experiments and J. Kurzman for his comments in reading the paper. Use of the 11-BM mail service of the APS at Argonne National Laboratory was supported by the U.S. Department of Energy under Contract No. DE-AC0206CH11357 and is greatly acknowledged. M.S. and M.-L.D acknowledge HPC resources from GENCI-CCRT/CINES (Grant cmm6691). Figure 1: Electron diffraction patterns of Li2Cu2O(SO4)2 along different crystallographic directions. Figure 2: Rietveld refinement of Synchrotron (a) and neutron (b) diffraction patterns of Li2Cu2O(SO4)2. The x-axis plots sin(θ)/λ which allows a direct comparison between the two patterns even if recorded at different wavelengths. The red crosses, black continuous line and bottom dark grey line represent the observed, calculated, and difference patterns, respectively. Vertical blue tick bars are the Bragg positions. The inset in a) is an enlargement at high angle. The right part of b) is a cut of the difference Fourier map at y=1/2. Blue domains correspond to negative values and reveal Li positions. Figure 3: Structure of Li2Cu2O(SO4)2. a) and b) general view of the unit cell content along [001] and [010], respectively. c) Cu coordination and Cu-O bond lengths d) the chains of edge-shared square CuO4 running along [001]. Li is green, Cu is blue, O is red, S is yellow. Figure 4: Thermal stability of Li2Cu2O(SO4)2. Thermogravimetric analysis under air (blue line) and argon (green line) between room temperature and 800°C, coupled with mass

Figure 8: Phase stability diagrams, as computed from DFT+U, for the LixCu2O(SO4)2 system for 0 ≤ x ≤ 2 (a) and 2 ≤ x ≤ 6 (b). The black circles and triangles correspond to the formation energy (∆H in eV) of LixCu2O(SO4)2 single phases 0 (insertion reaction) or {LixCu2-x/2O(SO4)2 + x/2Cu } composite electrodes (displacement reaction). For 0 ≤ x ≤ 2 and 2 ≤ x ≤ 6, the reference energies illustrated by a green horizontal line correspond to the Li2Cu2O(SO4)2 – 2Li → Cu2O(SO4)2 transformation and the direct conversion reaction 0 Li2Cu2O(SO4)2 + 4 Li → Li2O + 2Cu + 2Li2SO4, respectively. The most probable reaction mechanisms (thermodynamically favored) are illustrated by the convex hull of these phase diagrams (blue lines) while competitive reactions are indicated in dotted red lines. The average voltages vs. Li are indicated for the insertion processes only since they cannot be straightforwardly compared with experiments for conversion processes where large voltage hysteresis occurs between charge and discharge. Figure 9: Voltage versus composition curve of a Li2Cu2O(SO4)2/Li cell started on reduction, cycled between 1.8 V and 4.6 V at C/25 (red curves, left) or 2 V to 3.9 V at C/15 (blue curves, right). The bottom panel represents for each the derivative dx/dV curve for each (plain line = first cycle, dotted line = second cycle). In each case the corresponding capacity retentions are reported in inset (open circles = capacity on charge, triangles = capacity on discharge). Figure 10: Behavior of a Li2Cu2O(SO4)2/Li cell on discharge cycled at C/30. Left: in situ XRD patterns recorded behind a Be window while the cell is being discharged to 2 V (right curve). The pattern on top was recorded at the end of discharge on the Li side and shows Bragg peaks consistent with Cu metal.

ACS Paragon Plus Environment

Chemistry of Materials

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 11: HAADF-STEM images (a, d) show numerous nanoparticles, which are enriched with Cu, according to the compositional EDX maps (d, e). The intensity profiles of the Cu-K and O-K lines demonstrate that in the sample discharged to 1.8 V, the O-K signal varies concomitantly with the Cu-K signal (c) indicating that the particles are copper oxide (likely Cu2O). In the sample discharged to 0.5 V (f) the O-K signal does not depend on the presence/absence of the nanoparticles, indicating that they are metallic Cu.

(13)

Figure 12: TEM image (a), corresponding electron diffraction pattern (b), typical HAADF-STEM image and compositional EDX maps (c) of the Li2Cu2O(SO4)2 samples recharged to 3.8V. The diffraction rings are attributed to graphitic carbon. No segregation of Cu in the form of nanoparticles is observed.

(18)

Figure 13: Oscillations waves in Li2Cu2O(SO4)2. a) Voltage versus composition curve of a Li2Cu2O(SO4)2/Li cell started on reduction, cycled between 1.8 V and 4.6V at C/30, showing the presence of oscillations in the voltage profile at the second discharge; the inset is an enlargement on this region b) GITT measurements on Li2Cu2O(SO4)2. The cell was cycled at C/25, charged for 2.5 hours with relaxation times of 5 hours. The insets highlights domains for which the voltage oscillates (zooms 1 and 2) or presents the usual relaxation behavior (zoom 3).

Table 1: Structural parameters for Li2Cu2O(SO4)2, deduced from the combined Rietveld refinement of the Synchrotron XRD and neutron diffraction patterns. A bond valence sum analysis (BVS) for each atom is also reported.

(3)

(4)

(5) (6)

(7) (8)

(9) (10)

(11) (12)

(15)

(16) (17)

(19)

(20)

(21) (22) (23) (24) (25)

(26) (27)

(28)

(29) (30) (31) (32)

REFERENCES (1) (2)

(14)

Armand, M.; Tarascon, J. M. Nature 2008, 451, 652–657. Lu, Z.; Beaulieu, L. Y.; Donaberger, R. A.; Thomas, C. L.; Dahn, J. R. J. Electrochem. Soc. 2002, 149, A778–A791. Koga, H.; Croguennec, L.; Ménétrier, M.; Douhil, K.; Belin, S.; Bourgeois, L.; Suard, E.; Weill, F.; Delmas, C. J. Electrochem. Soc. 2013, 160, A786–A792. Sathiya, M.; Rousse, G.; Ramesha, K.; Laisa, C. P.; Vezin, H.; Sougrati, M. T.; Doublet, M.-L.; Foix, D.; Gonbeau, D.; Walker, W.; Prakash, A. S.; Ben Hassine, M.; Dupont, L.; Tarascon, J.M. Nat. Mater. 2013, 12, 827–835. Padhi, A. K.; Nanjundaswamy, K. S.; Goodenough, J. B. J. Electrochem. Soc. 1997, 144, 1188–1194. Barpanda, P.; Ati, M.; Melot, B. C.; Rousse, G.; Chotard, J. N.; Doublet, M. L.; Sougrati, M. T.; Corr, S. A.; Jumas, J. C.; Tarascon, J. M. Nat. Mater. 2011, 10, 772–779. Padhi, A. K.; Nanjundaswamy, K. S.; Masquelier, C.; Goodenough, J. B. J. Electrochem. Soc. 1997, 144, 2581–2586. Recham, N.; Rousse, G.; Sougrati, M. T.; Chotard, J. N.; Frayret, C.; Sathiya, M.; Melot, B. C.; Jumas, J. C.; Tarascon, J. M. Chem. Mater. 2012, 24, 4363–4370. Ati, M.; Melot, B. C.; Rousse, G.; Chotard, J.-N.; Barpanda, P.; Tarascon, J.-M. Angew. Chem. Int. Ed. 2011, 50, 10574–10577. Reynaud, M.; Barpanda, P.; Rousse, G.; Chotard, J.-N.; Melot, B. C.; Recham, N.; Tarascon, J.-M. Solid State Sci. 2012, 14, 15–20. Recham, N.; Chotard, J. N.; Dupont, L.; Delacourt, C.; Walker, W.; Armand, M.; Tarascon, J. M. Nat. Mater. 2010, 9, 68–74. Tripathi, R.; Popov, G.; Ellis, B. L.; Huq, A.; Nazar, L. F. Energy Environ. Sci. 2012, 5, 6238–6246.

(33) (34)

(35)

(36) (37)

(38) (39) (40) (41) (42)

(43) (44)

Page 8 of 23

Tripathi, R.; Ramesh, T. N.; Ellis, B. L.; Nazar, L. F. Angew. Chem. Int. Ed. 2010, 49, 8738–8742. Ati, M.; Melot, B. C.; Chotard, J.-N.; Rousse, G.; Reynaud, M.; Tarascon, J.-M. Electrochem. Commun. 2011, 13, 1280–1283. Subban, C. V.; Ati, M.; Rousse, G.; Abakumov, A. M.; Van Tendeloo, G.; Janot, R.; Tarascon, J.-M. J. Am. Chem. Soc. 2013, 135, 3653–3661. Reddy, M. A.; Pralong, V.; Caignaert, V.; Varadaraju, U. V.; Raveau, B. Electrochem. Commun. 2009, 11, 1807–1810. Reynaud, M.; Ati, M.; Melot, B. C.; Sougrati, M. T.; Rousse, G.; Chotard, J.-N.; Tarascon, J.-M. Electrochem. Commun. 2012, 21, 77–80. Reynaud, M.; Rousse, G.; Chotard, J.-N.; Rodríguez-Carvajal, J.; Tarascon, J.-M. Inorg. Chem. 2013, 52, 10456–10466. Lander, L.; Reynaud, M.; Rousse, G.; Sougrati, M. T.; LabertyRobert, C.; Messinger, R. J.; Deschamps, M.; Tarascon, J.-M. Chem. Mater. 2014, 26, 4178–4189. Sun, M.; Rousse, G.; Abakumov, A. M.; Van Tendeloo, G.; Sougrati, M.-T.; Courty, M.; Doublet, M.-L.; Tarascon, J.-M. J. Am. Chem. Soc. 2014, 136, 12658–12666. Hautier, G.; Jain, A.; Ong, S. P.; Kang, B.; Moore, C.; Doe, R.; Ceder, G. Chem. Mater. 2011, 23, 3495–3508. Mueller, T.; Hautier, G.; Jain, A.; Ceder, G. Chem. Mater. 2011, 23, 3854–3862. Effenberger, H. Monats. Chem. 1985, 116 , 927–931. Mrose, M. E. Am. Mineral. 46, 146–154. Patoux, S. Matériaux d’électrode positive à charpente polyanionique pour batteries au lithium : approches cristallochimiques et électrochimiques; PhD thesis, Amiens, 2003. Boultif, A.; Louer, D. J. Appl. Crystallogr. 1991, 24, 987–993. Altomare, A.; Burla, M. C.; Camalli, M.; Carrozzini, B.; Cascarano, G. L.; Giacovazzo, C.; Guagliardi, A.; Moliterni, A. G. G.; Polidori, G.; Rizzi, R. J. Appl. Crystallogr. 1999, 32, 339– 340. Fischer, P.; Frey, G.; Koch, M.; Könnecke, M.; Pomjakushin, V.; Schefer, J.; Thut, R.; Schlumpf, N.; Bürge, R.; Greuter, U.; Bondt, S.; Berruyer, E. Physica B: Condensed Matter 2000, 276–278, 146–147. Rodríguez-Carvajal, J. FullProf Suite. All the programs can be obtained from: http:://www.ill.eu/sites/fullprof Rietveld, H. M. J. Appl. Crystallogr. 1969, 2, 65–71. Rodríguez-Carvajal, J. Physica B 1993, 192, 55–69. Brown, I. D.; Altermatt, D. Acta Crystallogr. Sect. B-Struct. Commun. 1985, 41, 244–247. Adams, S. Solid State Ionics. 2006, 177, 1625–1630. Adams, S. In Bond Valences; Brown, I. D., Poeppelmeier, K. R., Eds.; Structure and Bonding; Springer Berlin Heidelberg, 2013; 91–128. Adams, S.; Rao, R. P. In Bond Valences; Brown, I. D., Poeppelmeier, K. R., Eds.; Structure and Bonding; Springer Berlin Heidelberg, 2014; 129–159. Avdeev, M.; Sale, M.; Adams, S.; Rao, R. P. Solid State Ionics. 2012, 225, 43–46. Leriche, J. B.; Hamelet, S.; Shu, J.; Morcrette, M.; Masquelier, C.; Ouvrard, G.; Zerrouki, M.; Soudan, P.; Belin, S.; Elkaim, E.; Baudelet, F. J. Electrochem. Soc. 2010, 157, A606–A610. Morcrette, M.; Rozier, P.; Dupont, L.; Mugnier, E.; Sannier, L.; Galy, J.; Tarascon, J.-M. Nat. Mater. 2003, 2, 755–761. Poizot, P.; Laruelle, S.; Grugeon, S.; Dupont, L.; Tarascon, J. M. Nature 2000, 407, 496–499. Cabana, J.; Monconduit, L.; Larcher, D.; Palacín, M. R. Adv. Mater. 2010, 22, E170–E192. Reimers, J. N.; Dahn, J. R. J. Electrochem. Soc. 1992, 139, 2091–2097. Prakash, A. S.; Larcher, D.; Morcrette, M.; Hegde, M. S.; Leriche, J.-B.; Masquelier, C. Chem. Mater. 2005, 17, 4406– 4415. Kirfel, A.; Eichhorn, K. Acta Crystallogr., Sect. A: Found. Crystallogr. 1990, 46, 271–284. Okeeffe, M.; Bovin, J. Am. Miner. 1978, 63, 180–185.

ACS Paragon Plus Environment

Page 9 of 23

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

Table 1. Structural parameters for Li2Cu2O(SO4)2, deduced from the combined Rietveld refinement of the Synchrotron XRD and neutron diffraction patterns. A bond valence sum analysis (BVS) for each atom is also reported.

Li2Cu2O(SO4)2, Space Group: P42/m a= 8.32586(2) Å, c=5.091351(17) Å, V= 352.933 (2) Å

3

Atom

Wyckoff site

x

y

z

B(Å2)

BVS

Li1

2d

0

0.5

0.5

2.0(3)

0.916(2)

Li2

2f

0.5

0.5

0.25

0.62(19)

1.255(4)

Cu

4j

0.16033(9)

0.06497(9)

0

0.482(11)

2.030(4)

S

4j

0.31730(18)

0.22610(17

0.5

0.60(2)

6.036(21)

O1

8k

0.3305(3)

0.1237(2)

0.7367(4)

1.32(4)

2.019(9)

O2

4j

0.1679(3)

0.3132(3)

0.5

0.85(5)

2.007(13)

O2

4j

0.4577(3)

0.3368(4)

0.5

0.75(5)

2.069(13)

O4

2e

0

0

0.25

0.46(7)

2.073(2)

2

Reliability parameters: χ =3.18; Neutrons: Bragg R-factor=4.16%, Rf-factor= 2.38%; Synchrotron: Bragg R-factor=5.39%, Rf-factor=4.57%

9 Environment ACS Paragon Plus

Chemistry of Materials

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 1

10 Environment ACS Paragon Plus

Page 10 of 23

Page 11 of 23

Intensity (arb. Units)

Figure 2

a

Synchrotron X-rays

a b Intensity (arb. Units)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

neutrons

Li1 (2d) Li2 (2f)

Li1 (2d) c y=0.5

11 Environment ACS Paragon Plus

Chemistry of Materials

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 23

Figure 3

a

b

c

Cu2+ = 3d9

Cu-O4 = 1.92 Å (x2) Cu-O1 = 2.01 Å (x2) Cu-O2 = 2.32 Å

12 Environment ACS Paragon Plus

d

Page 13 of 23

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

Figure 4

SO (48) SO2 (64)

13 Environment ACS Paragon Plus

Chemistry of Materials

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 5

14 Environment ACS Paragon Plus

Page 14 of 23

Page 15 of 23

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

Figure 6

15 Environment ACS Paragon Plus

Chemistry of Materials

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 7

16 Environment ACS Paragon Plus

Page 16 of 23

Page 17 of 23

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

Figure 8

17 Environment ACS Paragon Plus

Chemistry of Materials

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 9

18 Environment ACS Paragon Plus

Page 18 of 23

Page 19 of 23

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

Figure 10

19 Environment ACS Paragon Plus

Chemistry of Materials

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 11

20 Environment ACS Paragon Plus

Page 20 of 23

Page 21 of 23

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

Figure 12

21 Environment ACS Paragon Plus

Chemistry of Materials

4

2.55 2.5 2.45 2.4 2.7

3

3.3

x in Li x C u 2 O (SO 4 ) 2

3

4

3.5

Lix Cu 2 O(SO 4) 2

4

4.5

164

2 2.14

1

2.13 2.12

106

3

3

1.9

3

0 2.5

2.1

T/h

1

2

2.3

158 160 162

2 1.5

Volts vs. Li+/Li

(b)

0

Volts vs. Li+/Li

5

5 2.6

Volts vs. Li+/Li

(a)

Volts vs. Li+/Li

V o ltag e (V vs. L i + /Li 0 )

Figure 13

Voltage (V vs. Li +/Li)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 22 of 23

T/h

40

2.14

2

2.13 2.12

106

107

80

T/h

120

Li2Cu2O(SO4)2 ; t/h

22 Environment ACS Paragon Plus

107

160

Page 23 of 23

For Table of Contents Only

Voltage (V vs. Li)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

5

Cu3+/Cu2+ 4.7 V

4 3 2

Cu2+/Cu+ 1.5

2 2.5 3 3.5 4 x in LixCu2O(SO4)2

23 Environment ACS Paragon Plus

4.5