Network


Latest external collaboration on country level. Dive into details by clicking on the dots.

Hotspot


Dive into the research topics where Clare P. Grey is active.

Publication


Featured researches published by Clare P. Grey.


Journal of the American Chemical Society | 2009

Real-Time NMR Investigations of Structural Changes in Silicon Electrodes for Lithium-Ion Batteries

Baris Key; Rangeet Bhattacharyya; Mathieu Morcrette; Vincent Seznec; Jean-Marie Tarascon; Clare P. Grey

Lithium-ion batteries (LIBs) containing silicon negative electrodes have been the subject of much recent investigation because of the extremely large gravimetric and volumetric capacity of silicon. The crystalline-to-amorphous phase transition that occurs on electrochemical Li insertion into crystalline Si, during the first discharge, hinders attempts to link structure in these systems with electrochemical performance. We apply a combination of static, in situ and magic angle sample spinning, ex situ (7)Li nuclear magnetic resonance (NMR) studies to investigate the changes in local structure that occur in an actual working LIB. The first discharge occurs via the formation of isolated Si atoms and smaller Si-Si clusters embedded in a Li matrix; the latter are broken apart at the end of the discharge, forming isolated Si atoms. A spontaneous reaction of the lithium silicide with the electrolyte is directly observed in the in situ NMR experiments; this mechanism results in self-discharge and potential capacity loss. The rate of this self-discharge process is much slower when CMC (carboxymethylcellulose) is used as the binder.


Nature Materials | 2013

Origin of additional capacities in metal oxide lithium-ion battery electrodes

Yan-Yan Hu; Zigeng Liu; Kyung-Wan Nam; Olaf J. Borkiewicz; Jun Cheng; Xiao Hua; Matthew T. Dunstan; Xiqian Yu; Kamila M. Wiaderek; Lin-Shu Du; Karena W. Chapman; Peter J. Chupas; Xiao-Qing Yang; Clare P. Grey

Metal fluorides/oxides (MF(x)/M(x)O(y)) are promising electrodes for lithium-ion batteries that operate through conversion reactions. These reactions are associated with much higher energy densities than intercalation reactions. The fluorides/oxides also exhibit additional reversible capacity beyond their theoretical capacity through mechanisms that are still poorly understood, in part owing to the difficulty in characterizing structure at the nanoscale, particularly at buried interfaces. This study employs high-resolution multinuclear/multidimensional solid-state NMR techniques, with in situ synchrotron-based techniques, to study the prototype conversion material RuO2. The experiments, together with theoretical calculations, show that a major contribution to the extra capacity in this system is due to the generation of LiOH and its subsequent reversible reaction with Li to form Li2O and LiH. The research demonstrates a protocol for studying the structure and spatial proximities of nanostructures formed in this system, including the amorphous solid electrolyte interphase that grows on battery electrodes.


Science | 2008

Mg/Al Ordering in Layered Double Hydroxides Revealed by Multinuclear NMR Spectroscopy

Paul J. Sideris; Ulla Gro Nielsen; Zhehong Gan; Clare P. Grey

The anion-exchange ability of layered double hydroxides (LDHs) has been exploited to create materials for use in catalysis, drug delivery, and environmental remediation. The specific cation arrangements in the hydroxide layers of hydrotalcite-like LDHs, of general formula Mg2+1–xAl3+xOH2(Anionn–x/n)·yH2O, have, however, remained elusive, and their elucidation could enhance the functional optimization of these materials. We applied rapid (60 kilohertz) magic angle spinning (MAS) to obtain high-resolution hydrogen-1 nuclear magnetic resonance (1H NMR) spectra and characterize the magnesium and aluminum distribution. These data, in combination with 1H-27Al double-resonance and 25Mg triple-quantum MAS NMR data, show that the cations are fully ordered for magnesium:aluminum ratios of 2:1 and that at lower aluminum content, a nonrandom distribution of cations persists, with no Al3+-Al3+ close contacts. The application of rapid MAS NMR methods to investigate proton distributions in a wide range of materials is readily envisaged.


Journal of Applied Crystallography | 2003

Rapid‐acquisition pair distribution function (RA‐PDF) analysis

Peter J. Chupas; Xiangyun Qiu; Jonathan C. Hanson; Peter L. Lee; Clare P. Grey; Simon J. L. Billinge

An image-plate (IP) detector coupled with high-energy synchrotron radiation was used for atomic pair distribution function (PDF) analysis, with high probed momentum transfer Qmax ≤ 28.5 A−1, from crystalline materials. Materials with different structural complexities were measured to test the validity of the quantitative data analysis. Experimental results are presented for crystalline Ni, crystalline α-AlF3, and the layered Aurivillius type oxides α-Bi4V2O11 and γ-Bi4V1.7Ti0.3O10.85. Overall, the diffraction patterns show good counting statistics, with measuring time from one to tens of seconds. The PDFs obtained are of high quality. Structures may be refined from these PDFs, and the structural models are consistent with the published literature. Data sets from similar samples are highly reproducible.


Journal of the American Chemical Society | 2009

Lithium Salt of Tetrahydroxybenzoquinone: Toward the Development of a Sustainable Li-Ion Battery

Haiyan Chen; Michel Armand; Matthieu Courty; Meng Jiang; Clare P. Grey; Franck Dolhem; Jean-Marie Tarascon; Philippe Poizot

The use of lithiated redox organic molecules containing electrochemically active C=O functionalities, such as lithiated oxocarbon salts, is proposed. These represent alternative electrode materials to those used in current Li-ion battery technology that can be synthesized from renewable starting materials. The key material is the tetralithium salt of tetrahydroxybenzoquinone (Li(4)C(6)O(6)), which can be both reduced to Li(2)C(6)O(6) and oxidized to Li(6)C(6)O(6). In addition to being directly synthesized from tetrahydroxybenzoquinone by neutralization at room temperature, we demonstrate that this salt can readily be formed by the thermal disproportionation of Li(2)C(6)O(6) (dilithium rhodizonate phase) under an inert atmosphere. The Li(4)C(6)O(6) compound shows good electrochemical performance vs Li with a sustained reversibility of approximately 200 mAh g(-1) at an average potential of 1.8 V, allowing a Li-ion battery that cycles between Li(2)C(6)O(6) and Li(6)C(6)O(6) to be constructed.


Science | 2015

Cycling Li-O2 batteries via LiOH formation and decomposition

Tao Liu; Michal Leskes; Wanjing Yu; Amy J Moore; Lina Zhou; Paul M. Bayley; Gunwoo Kim; Clare P. Grey

Solving the problems with Li-air batteries Li-air batteries come as close as possible to the theoretical limits for energy density in a battery. By weight, this is roughly 10 times higher than conventional lithium-ion batteries and would be sufficient to power cars with a range comparable to those with gasoline engines. But engineering a Li-air battery has been a challenge. Liu et al. managed to overcome the remaining challenges: They were able to avoid electrode passivation, turn limited solvent stability into an advantage, eliminate the fatal problems caused by superoxides, achieve high power with negligible degradation, and even circumvent the problems of removing atmospheric water. Science, this issue p. 530 An efficient, high-capacity, safe Li-air battery system is realized. The rechargeable aprotic lithium-air (Li-O2) battery is a promising potential technology for next-generation energy storage, but its practical realization still faces many challenges. In contrast to the standard Li-O2 cells, which cycle via the formation of Li2O2, we used a reduced graphene oxide electrode, the additive LiI, and the solvent dimethoxyethane to reversibly form and remove crystalline LiOH with particle sizes larger than 15 micrometers during discharge and charge. This leads to high specific capacities, excellent energy efficiency (93.2%) with a voltage gap of only 0.2 volt, and impressive rechargeability. The cells tolerate high concentrations of water, water being the dominant proton source for the LiOH; together with LiI, it has a decisive impact on the chemical nature of the discharge product and on battery performance.


Journal of the American Chemical Society | 2011

Conversion Reaction Mechanisms in Lithium Ion Batteries: Study of the Binary Metal Fluoride Electrodes

Feng Wang; Rosa Robert; Natasha A. Chernova; Nathalie Pereira; Fredrick Omenya; Fadwa Badway; Xiao Hua; Michael Ruotolo; Ruigang Zhang; Lijun Wu; Vyacheslav Volkov; Dong Su; Baris Key; M. Stanley Whittingham; Clare P. Grey; Glenn G. Amatucci; Yimei Zhu; Jason Graetz

Materials that undergo a conversion reaction with lithium (e.g., metal fluorides MF(2): M = Fe, Cu, ...) often accommodate more than one Li atom per transition-metal cation, and are promising candidates for high-capacity cathodes for lithium ion batteries. However, little is known about the mechanisms involved in the conversion process, the origins of the large polarization during electrochemical cycling, and why some materials are reversible (e.g., FeF(2)) while others are not (e.g., CuF(2)). In this study, we investigated the conversion reaction of binary metal fluorides, FeF(2) and CuF(2), using a series of local and bulk probes to better understand the mechanisms underlying their contrasting electrochemical behavior. X-ray pair-distribution-function and magnetization measurements were used to determine changes in short-range ordering, particle size and microstructure, while high-resolution transmission electron microscopy (TEM) and electron energy-loss spectroscopy (EELS) were used to measure the atomic-level structure of individual particles and map the phase distribution in the initial and fully lithiated electrodes. Both FeF(2) and CuF(2) react with lithium via a direct conversion process with no intercalation step, but there are differences in the conversion process and final phase distribution. During the reaction of Li(+) with FeF(2), small metallic iron nanoparticles (<5 nm in diameter) nucleate in close proximity to the converted LiF phase, as a result of the low diffusivity of iron. The iron nanoparticles are interconnected and form a bicontinuous network, which provides a pathway for local electron transport through the insulating LiF phase. In addition, the massive interface formed between nanoscale solid phases provides a pathway for ionic transport during the conversion process. These results offer the first experimental evidence explaining the origins of the high lithium reversibility in FeF(2). In contrast to FeF(2), no continuous Cu network was observed in the lithiated CuF(2); rather, the converted Cu segregates to large particles (5-12 nm in diameter) during the first discharge, which may be partially responsible for the lack of reversibility in the CuF(2) electrode.


Science | 2014

Capturing metastable structures during high-rate cycling of LiFePO4 nanoparticle electrodes

Hao Liu; Fiona C. Strobridge; Olaf J. Borkiewicz; Kamila M. Wiaderek; Karena W. Chapman; Peter J. Chupas; Clare P. Grey

Introduction The ability to achieve high cycling rates in a lithium-ion battery is limited by the Li transport within the electrolyte; the transport of Li ions and electrons within the electrodes; and, when a phase transformation is induced as a result of the Li compositional changes within an electrode, the nucleation and growth of the second phase. The absence of a phase transformation involving substantial structural rearrangements and large volume changes is generally considered to be key for achieving high rates. This assumption has been challenged by the discovery that some nanoparticulate electrode materials, most notably LiFePO4, can be cycled in a battery at very high rates, even though they cycle between two phases during battery operation. This apparent contradiction has been reconciled by the hypothesis that a nonequilibrium solid solution can be formed during reaction to bypass the nucleation step. Phase transformation from LiFePO4 (blue) to FePO4 (red). The delithiation (indicated by yellow arrows) proceeds at high rates via the formation of a nonequilibrium solid solution phase LixFePO4 (intermediate purple color), avoiding a classical nucleation process (indicated by dashed arrows). When the reaction is interrupted, the particles relax into the equilibrium configuration (shaded region), where only single-phase particles of LiFePO4 and/or FePO4 are present. Rationale To test this proposal, in situ techniques with high temporal resolution must be used to capture the fast phase transformation processes. We performed in situ synchrotron x-ray diffraction (XRD), which readily detects the structural changes and allows for fast data collection, on a LiFePO4-Li battery at high cycling rates, conditions that are able to drive the system away from equilibrium. We used an electrode comprising ~190-nm LiFePO4 particles, carbon, and binder (30:60:10 weight %), along with an electrochemical cell designed to yield reproducible results over multiple cycles, even at high rates. The high carbon content ensures that the reaction at high rates is not limited by either the electronic conductivity or ionic diffusion within the electrode composite. We compared the experimental results with simulated XRD patterns, in which the effects of strain versus compositional variation were explored. We then adapted a whole-pattern fitting method to quantify the compositional variation in the electrode during cycling. Results The XRD patterns, collected during high-rate galvanostatic cycling, show the expected disappearance of LiFePO4 Bragg reflections on charge and the simultaneous formation of FePO4 reflections. In addition, the development of positive intensities between the LiFePO4 and FePO4 reflections indicates that particles with lattice parameters that deviate from the equilibrium values of LiFePO4 and FePO4 are formed. The phenomenon is more pronounced at high currents. Detailed simulations of the XRD patterns reveal that this lattice-parameter variation cannot be explained by a LiFePO4-FePO4 interface within the particles, unless the size of the interface is similar to or greater than the size of the entire particle. Instead, the results indicate compositional variation either within or between particles. Conclusion The results demonstrate the formation of a nonequilibrium solid solution phase, LixFePO4 (0 < x < 1), during high-rate cycling, with compositions that span the entire composition between two thermodynamic phases, LiFePO4 and FePO4. This confirms the hypothesis that phase transformations in nanoparticulate LiFePO4 proceed, at least at high rates, via a continuous change in structure rather than a distinct moving phase boundary between LiFePO4 and FePO4. The ability of LiFePO4 to transform via a nonequilibrium single-phase solid solution, which avoids major structural rearrangement across a moving interface, helps to explain its high-rate performance despite a large Li miscibility gap at room temperature. The creation of a low-energy nonequilibrium path by, for example, particle size reduction or cation doping should enable the high-rate capabilities of other phase-transforming electrode materials. Watching battery materials in action When batteries get rapidly charged and discharged repeatedly, they will often stop working. This is especially true when the cycling changes the crystal structure of the battery components. Liu et al. examined the structural changes in components of a type of lithium battery (see the Perspective by Owen and Hector). Their findings explain why LiFePO4 delivers unexpectedly good electrochemical performances, particularly during rapid cycling. Science, this issue p. 10.1126/science.1252817; see also p. 1451 X-ray diffraction reveals that metastable solid solution reactions undergird the high-rate capability of LiFePO4 electrodes. [Also see Perspective by Owen and Hector] The absence of a phase transformation involving substantial structural rearrangements and large volume changes is generally considered to be a key characteristic underpinning the high-rate capability of any battery electrode material. In apparent contradiction, nanoparticulate LiFePO4, a commercially important cathode material, displays exceptionally high rates, whereas its lithium-composition phase diagram indicates that it should react via a kinetically limited, two-phase nucleation and growth process. Knowledge concerning the equilibrium phases is therefore insufficient, and direct investigation of the dynamic process is required. Using time-resolved in situ x-ray powder diffraction, we reveal the existence of a continuous metastable solid solution phase during rapid lithium extraction and insertion. This nonequilibrium facile phase transformation route provides a mechanism for realizing high-rate capability of electrode materials that operate via two-phase reactions.


Angewandte Chemie | 2012

Nanoparticulate TiO2(B): an anode for lithium-ion batteries.

Yu Ren; Zheng Liu; Frédérique Pourpoint; A. Robert Armstrong; Clare P. Grey; Peter G. Bruce

Titanates are being intensively investigated as anodes for lithium-ion batteries due to their superior safety and rate capability compared with graphite, although their higher voltage lowers the overall energy density of the lithium-ion cell. Li4Ti5O12 spinel is now used in commercial lithium-ion batteries. TiO2 possesses twice the theoretical specific capacity (335 mAh g ) compared with Li4Ti5O12 (175 mAhg ), i.e., is comparable to graphite, rendering TiO2 potentially attractive as an anode for Li-ion batteries. 3] TiO2(B) can accommodate more lithium than any other TiO2 polymorph as a bulk material (micrometer-sized particles). 4] It has been shown that nanostructured forms of TiO2(B) enhance rate capability compared to the bulk, with nanotubes of TiO2(B) or TiO2(B)/anatase microspheres exhibiting the highest rate capability to date. Their performance is exceeded by nanoparticulate TiO2(B) described here. Nanoparticulate TiO2(B) was synthesized by a procedure described in the Experimental Section. Briefly, Ti metal is dissolved in a mixture of H2O2 and NH3 in water, to which glycolic acid is added, forming a titanium glycolate complex. This is subjected to hydrothermal treatment at 160 8C for 30 min. The resulting solid was finally calcined in dry air at 300 8C for 1 h. The powder X-ray diffraction pattern (PXRD) of nanoparticulate TiO2(B) is shown in Figure 1, where it is compared with standard TiO2(B). The small particle dimensions necessarily result in peak broadening but the powder diffraction pattern follows the same intensity distribution as that of the bulk material. Confirmation that the nanoparticulate powder is TiO2(B) was obtained by high-resolution TEM (Figure 2); lattice spacings of 0.357 nm and 0.619 nm corresponding to the (110) and (001) reflections from the TiO2(B) structure (ICDD 046-1237) are observed. O NMR spectra of O-enriched TiO2 clearly show characteristic resonances from the three oxygen coordination environments OTi2, OTi3, and OTi4 in intensity ratios of approximately 1:2:1, which is characteristic of the TiO2(B) polymorph (Supporting Information, Figure S2). Rutile and anatase both contain only OTi3 environments. The TEM data demonstrate that the material is composed of nanoparticles of ca. 2.5 4.3 nm size (based on analysis of 100 nanoparticles), with a relatively narrow size distribution, and that form agglomerates of 0.3–3 mm (Figure 2a and S1). The BET surface area determined from N2 adsorption is 251 m g 1 (pore volume 0.12 cm g ) whereas, based on the primary particle size of 2.5 2.5 4.3 nm, the predicted surface area is 567 m g , i.e., more than twice the observed value and consistent with aggregation of the primary particles into porous agglomerates. One problem that often besets nanoparticles is the need to employ molecules (e.g. surfactants) in the synthesis to inhibit particle size growth. Such molecules, if they remain on the Figure 1. PXRD pattern of the TiO2(B) nanoparticles. Bulk TiO2(B) from the ICDD database is shown for comparison.


Journal of the American Chemical Society | 2011

Pair distribution function analysis and solid state NMR studies of silicon electrodes for lithium ion batteries: Understanding the (de)lithiation mechanisms

Baris Key; Mathieu Morcrette; Jean-Marie Tarascon; Clare P. Grey

Lithium ion batteries (LIBs) containing silicon negative electrodes have been the subject of much recent investigation, because of the extremely large gravimetric and volumetric capacities of silicon. The crystalline-to-amorphous phase transition that occurs on electrochemical Li insertion into crystalline Si, during the first discharge, hinders attempts to link the structure in these systems with electrochemical performance. We apply a combination of local structure probes, ex situ (7)Li nuclear magnetic resonance (NMR) studies, and pair distribution function (PDF) analysis of X-ray data to investigate the changes in short-range order that occur during the initial charge and discharge cycles. The distinct electrochemical profiles observed subsequent to the first discharge have been shown to be associated with the formation of distinct amorphous lithiated silicide structures. For example, the first process seen on the second discharge is associated with the lithiation of the amorphous Si, forming small clusters. These clusters are broken in the second process to form isolated silicon anions. The (de)lithiation model helps explain the hysteresis and the steps in the electrochemical profile observed during the lithiation and delithiation of silicon.

Collaboration


Dive into the Clare P. Grey's collaboration.

Top Co-Authors

Avatar

Peter J. Chupas

Argonne National Laboratory

View shared research outputs
Top Co-Authors

Avatar
Top Co-Authors

Avatar

Jordi Cabana

University of Illinois at Chicago

View shared research outputs
Top Co-Authors

Avatar

Karena W. Chapman

Argonne National Laboratory

View shared research outputs
Top Co-Authors

Avatar
Top Co-Authors

Avatar

Gerbrand Ceder

University of California

View shared research outputs
Top Co-Authors

Avatar
Top Co-Authors

Avatar
Top Co-Authors

Avatar
Top Co-Authors

Avatar

Olaf J. Borkiewicz

Argonne National Laboratory

View shared research outputs
Researchain Logo
Decentralizing Knowledge