Processing math: 100%
Review Topical Sections

In situ Raman analyses of electrode materials for Li-ion batteries

  • Received: 10 July 2018 Accepted: 07 August 2018 Published: 14 August 2018
  • The purpose of this review is to acknowledge the current state-of-the-art and the progress of in situ Raman spectro-electrochemistry, which has been made on all the elements in lithium-ion batteries: positive (cathode) and negative (anode) electrode materials. This technique allows the studies of structural change at the short-range scale, the electrode degradation and the formation of solid electrolyte interphase (SEI) layer. We also discuss the results in the perspective of spatial and real-time investigations by in situ Raman imaging (mapping) during charge/discharge cycling.

    Citation: Christian M. Julien, Alain Mauger. In situ Raman analyses of electrode materials for Li-ion batteries[J]. AIMS Materials Science, 2018, 5(4): 650-698. doi: 10.3934/matersci.2018.4.650

    Related Papers:

    [1] Christian M Julien, Alain Mauger, Ashraf E Abdel-Ghany, Ahmed M Hashem, Karim Zaghib . Smart materials for energy storage in Li-ion batteries. AIMS Materials Science, 2016, 3(1): 137-148. doi: 10.3934/matersci.2016.1.137
    [2] Alain Mauger, Haiming Xie, Christian M. Julien . Composite anodes for lithium-ion batteries: status and trends. AIMS Materials Science, 2016, 3(3): 1054-1106. doi: 10.3934/matersci.2016.3.1054
    [3] Christian M. Julien, Alain Mauger . Functional behavior of AlF3 coatings for high-performance cathode materials for lithium-ion batteries. AIMS Materials Science, 2019, 6(3): 406-440. doi: 10.3934/matersci.2019.3.406
    [4] Andrea Somogyi, Cristian Mocuta . Possibilities and Challenges of Scanning Hard X-ray Spectro-microscopy Techniques in Material Sciences. AIMS Materials Science, 2015, 2(2): 122-162. doi: 10.3934/matersci.2015.2.122
    [5] Stephanie Flores Zopf, Matthew J. Panzer . Integration of UV-cured Ionogel Electrolyte with Carbon Paper Electrodes. AIMS Materials Science, 2014, 1(1): 59-69. doi: 10.3934/matersci.2014.1.59
    [6] Rucheng Zhu, Yota Mabuchi, Riteshkumar Vishwakarma, Balaram Paudel Jaisi, Haibin Li, Masami Naito, Masayoshi Umeno, Tetsuo Soga . Enhancing in-plane uniformity of graphene nanowalls using a rotating platform for solid-state lithium-ion battery. AIMS Materials Science, 2024, 11(4): 760-773. doi: 10.3934/matersci.2024037
    [7] Hiroki Nara, Keisuke Morita, Tokihiko Yokoshima, Daikichi Mukoyama, Toshiyuki Momma, Tetsuya Osaka . Electrochemical impedance spectroscopy analysis with a symmetric cell for LiCoO2 cathode degradation correlated with Co dissolution. AIMS Materials Science, 2016, 3(2): 448-459. doi: 10.3934/matersci.2016.2.448
    [8] Claas Hüter, Shuo Fu, Martin Finsterbusch, Egbert Figgemeier, Luke Wells, Robert Spatschek . Electrode–electrolyte interface stability in solid state electrolyte systems: influence of coating thickness under varying residual stresses. AIMS Materials Science, 2017, 4(4): 867-877. doi: 10.3934/matersci.2017.4.867
    [9] Michael A. Stamps, Jeffrey W. Eischen, Hsiao-Ying Shadow Huang . Particle- and crack-size dependency of lithium-ion battery materials LiFePO4. AIMS Materials Science, 2016, 3(1): 190-203. doi: 10.3934/matersci.2016.1.190
    [10] Dao Viet Thang, Nguyen Manh Hung, Nguyen Cao Khang, Le Thi Mai Oanh . Structural and multiferroic properties of (Sm, Mn) co-doped BiFeO3 materials. AIMS Materials Science, 2020, 7(2): 160-169. doi: 10.3934/matersci.2020.2.160
  • The purpose of this review is to acknowledge the current state-of-the-art and the progress of in situ Raman spectro-electrochemistry, which has been made on all the elements in lithium-ion batteries: positive (cathode) and negative (anode) electrode materials. This technique allows the studies of structural change at the short-range scale, the electrode degradation and the formation of solid electrolyte interphase (SEI) layer. We also discuss the results in the perspective of spatial and real-time investigations by in situ Raman imaging (mapping) during charge/discharge cycling.


    1. Introduction

    The improvement of high performance lithium-ion batteries requires an optimization of the structure of materials, the knowledge of the mechanism of the intercalation/deintercalation of Li+ into/from host materials as well as the kinetics of the electrolytic reaction [1]. Since the early report by Novák et al. [2], only few articles have addressed an overview of in situ microscopic studies on the structural and chemical behaviors of lithium-ion battery materials [3,4,5]. In this respect, in operando or in situ analytical techniques have been widely developed to investigate dynamic systems such as material growth [6], insertion/deinsertion reaction [7], phase stability [8], interface evolution [9], compatibility between materials [10], electrode degradation [11,12], etc.

    As a non-destructive and sensitive method, Raman spectroscopy (RS) is a powerful analytical technique that has the potential to study electrode materials under in situ conditions. RS involves an inelastic scattering of the electromagnetic radiation by the matter, characterized by a shift of the frequency of the scattered light with respect to the incident light. This shift is equal to the frequency of the vibrational modes of atoms or molecules that are Raman active, which depends only on their symmetry. Consequently, RS is a primary tool used to probe the structural evolution and the chemical composition of materials composing a lithium-ion battery. It gives information on the local structure (short-range order), in contrast to X-ray diffraction (XRD) that probes the long-range order of the matter. Therefore, RS is a complementary tool to diffraction techniques (X-ray, electron, neutron) to investigate the local structure of transition-metal oxides used as positive electrodes in lithium-ion batteries [13]. It proved to be very useful in characterizing the phase transformations in such materials upon cycling [14] or heat treatment [15]. Such changes can affect only the surface of the particles, in which case they are not detected by diffraction techniques. RS is also a powerful analytical technique for the knowledge of the structure of either disordered or amorphous compounds [16,17]. It should be noted that, in general, non-stoichiometry, substitutions or ion exchanges in a non-ideal material are violating the selection rules giving rise to additional bands, wavenumber shifts and/or band broadenings. For instance, Raman spectroscopy was employed to probe the nucleation and growth and structure of pulsed-laser deposited (PLD) V2O5 films [18]. The spectral features allow to establish a phase diagram mapping the effect of substrate temperature on the microstructure evolution of PLD V2O5 thin films. RS has also been widely used to study the surface phenomena of electrodes [19,20] and electrolyte including ion-solvent interactions, since the conductivity being proportional to the salt concentration, a correlation between Raman shift and the conductivity of the electrolyte can be established [21,22].

    In situ Raman spectroscopy is a direct method for real-time study of dynamic reactions during working conditions of electrochemical cells. Four spectral evolutions can be distinguished as (ⅰ) appearance of new band, (ⅱ) shift of vibrational mode, (ⅲ) change in band profile and (ⅳ) change in band intensities (Figure 1). In situ RS has been widely employed to elucidate the structural changes at the level of chemical bonds in electrode materials for lithium-ion batteries during the cycling of electrochemical cell; this analytical method is currently used to describe the variations in the valence state of transition-metal cations, lithium transport mechanism (insertion/deinsertion reaction), cycle-induced degradation, electrode–electrolyte interface evolution and state-of-charge (SOC) distribution as well [4,23]. Kinetic characterizations of Li+ insertion/extraction into/from composite cathode were also investigated by in situ Raman spectroscopy [24]. In situ RS is currently used to investigate the electrode–electrolyte interface in lithium batteries [25]. Taking the advantages in the difference in Raman spectra vs. composition or/and local symmetry, it is possible to study the spatial distribution of phases in an electrode by mapping [26]. The powerful advantage of in situ micro-Raman spectroscopy for the analysis of materials included in a lithium battery was emphasized by several authors [27,28,29].

    Figure 1. Schematic representation of spectral evolutions investigated by in situ Raman spectro-electrochemistry: (a) appearance of new band, (b) shift of vibrational mode, (c) change in band profile and (d) change in band intensities.

    The purpose of this review is to acknowledge the current state-of-the-art and the progress of in situ Raman spectroscopy that has been made on all the elements of the lithium-ion batteries: cathode and anode materials and the formation of solid electrolyte interphase (SEI) layer. We also discuss the results in the perspective of spatial and real-time investigations by in situ Raman imaging (mapping) during charge/discharge cycling.


    2. Raman single point and mapping measurements


    2.1. In situ Raman single point

    Typical in situ Raman spectroscopy is carried out using a confocal Raman spectrometer in backscattering geometry. For a description of the micro-Raman set-up, we guide the reader to the review papers Refs. [28,29]. Typical home-made electrochemical cells for in situ Raman studies were targeted to meet their experimental measurements [30,31,32,33]. Burba and Frech [30] described a spectro-electrochemical cell that was a modified industrially available coin cell with a 2-mm diameter hole drilled into the coin cell casing (Figure 2a). In situ RS cells are currently equipped with a glass, quartz or sapphire window transparent to the Raman laser light. A 100× magnification long working distance optical objective is used, providing a laser spot of approximately 3 μm2 on the sample electrode. Another in-situ cell was fabricated using the Swagelok standard [32] with a sapphire optical window (Figure 2b). A home-made in situ cell by Novák et al. [34] is presented in Figure 2c. Industry-standard 2032-coin cells (from Pred Materials International) were adapted to carried out in situ Raman mapping of electrode materials (Figures 2d, e). Fang et al. used a MgO window to cover a drilled 1/8 inch diameter hole. State-of-charge maps were collected of a 35 × 35 μm2 area considering two spectral features, i.e., Raman peak position and peak intensity [35]. Ghanty et al. [36] fabricated a in situ Raman pouch cell with an optical sodium borosilicate window covering an aperture in the aluminum-coated polyamide outer shell (Figure 2f).

    Figure 2. Cross section of home-made in situ Raman spectro-electrochemical cells (not drawn to scale): (a) Modified coin cells with a 2-mm diameter hole for in situ Raman spectro-electrochemical measurements of Li(x)V2O5 for lithium rechargeable batteries. From Ref. [30]. (b) Swagelok-type cell with sapphire optical window from Ref. [31]. (c) Schematic in situ Raman cell for the study of Li+ ion intercalation in graphite negative electrode. From Ref. [33]. (d) home-made cell design with glass window. From Ref. [34]. (e) in-situ setup with CR2016 coin cell adapted with quartz window. From Ref. [37]. (f) pouch cell with borosilicate window. From Ref. [36].

    2.2. Optical-skin depth

    A Raman spectrum represents the response of nanoparticles contained in the volume defined by the focal spot times the penetration depth δp that depends on the probed sample. The axial resolution in opaque materials is determined by the optical-skin depth (δp) of the laser beam penetrating into a material. Because of the optical absorption of the light in electrode materials, the observed Raman spectra are induced in a thin surface layer; thus, due to small δp value, Raman spectroscopy can be considered as a surface analytical method in most cases. The Raman scattering intensity is then weak because of the small material volume probed (V < 1 μm3). The skin depth is given by the relation:

    δp=(2λ/μσ)1/2 (1)

    where λ is the laser wavelength, σ the electronic conductivity and μ the magnetic permeability.

    Consequently, using laser excitation with higher wavelength increases the penetration depth and an increase in the electronic conductivity, e.g., from semiconductor to metal, results in decrease of δp. According to the Beer–Lambert law, the intensity I(ξ) of the laser light inside the materials is an exponential function of the absorption coefficient α of the material as I(ξ) = I0exp[−αξ]. The optical penetration depth defined as the distance at which the intensity decreases to 1/e (37%), where e is the base of the natural system of logarithms, we have δp = α−1. For instance, δp ≈ 1 μm in crystalline silicon, δp ≈ 100 nm in amorphous silicon and δp ≈ 100 nm in LixSi [38,39,40]. Currently, the experiments are performed at very low laser power Wlaser ≈ 0.02 mW using a 0.1% filter that corresponds to a specific power of 600 W·cm−2. An optical-skin depth is less than 50 nm in highly oriented pyrolytic graphite examined with a 514.5 nm green laser beam [41], while δp falls to few nanometers in delithiated Li1−xNi1/3Mn1/3Co1/3O2. Another experimental case is the low optical-skin depth (δp ≈ 30 nm) of carbon coating onto LiFePO4 particles [42,43]. Thus, to probe the electrode materials of LIBs, the choice of laser excitation line at high wavelength λ > 700 nm is preferred, rather than the solid-state 532 nm excitation.


    2.3. In situ Raman imaging

    In situ Raman mapping consists in the spatial characterization of an electrode at a multitude of locations that allows to define the chemical distribution and inhomogeneity of particles during the cell charge–discharge process and electrode degradation as well. A Raman image is formed by hyperspectral data set, in which a pixel is a complete Raman spectrum. Raman micro-spectroscopy as a spatially resolved analytical method is currently used to generate local surface images, to construct state-of-charge maps of electrodes, or to investigate the kinetic aspect of Li+ intercalation–deintercalation process in a single particle. The feasibility of Raman mapping was first explored by Panitz et al. [44] who investigated the lithium intercalation in graphite electrode under potentiostatic and galvanostatic conditions. Results indicated that, at a potential of 0.2 V vs. Li+/Li, the lithium is not homogeneously inserted over the graphite electrode and a new Raman band evolved at ca. 1850 cm−1 at potential 0.18 V vs. Li+/Li. The same group developed an in-situ cell to monitor the Li insertion in carbon electrode by changing the optical geometry that improved the signal-to noise ratio by a factor of 20 [45]. The mapping of the state-of-charge distribution of LiCoO2 (LCO) was investigated with a special resolution of few micrometers, which enabled to test the intensity, width and position of the Raman peaks by constructing the Raman images of the A1g mode (595 cm−1) [46]. The remarkable potential of Raman imaging was applied to study the local degradation behavior in composite cathode materials such as LiNi0.8Co0.15Al0.05O2 and LiNi0.33Mn0.33Co0.33O2 [47]. The authors investigated the nanoscale modifications of the surface composition, structure, SOC, and determined the electronic conductivity at the electrode surface as well. Raman imaging was performed on an area of 52 × 75 μm2 at 0.7 μm resolution. An image is formed of color-coded pixels, which represent the relative intensities of specific Raman peaks and the chemical surface composition, i.e., active material and carbonaceous additive (acetylene black, graphite, etc.) characterized by the Raman G- and D-bands. Currently, the Raman spectrum of an electrode particle varies significantly with the position on the surface. For example, the spectrum of LiNi0.33Mn0.33Co0.33O2 displays bands at 475 and 554 cm−1 associated with the vibrations of the Ni–O bonds having intensities strongly coupled with electronic states that provoke Raman resonance behavior. Consequently, the relative intensities I475/I554 vary significantly from one particle to another. Fang et al. [35] adapted industry-standard coin cells for high spatial resolving in situ Raman mapping. By extraction of the local frequency of the A1g mode of electrodes with rock-salt-type structure, the authors demonstrated an easy-to-implement electrode design using LiNi0.5Mn0.3Co0.2O2 (NMC) as active material.

    Quantification of the Li+ ion transport in liquid electrolyte, i.e., LiClO4 dissolved in dimethyl carbonate (DMC), was performed by combining high-spatial-resolution confocal Raman microscopy and microfluidic technique. [48]. The Raman peak at ca. 523 cm−1 corresponding to the O–C–O deformation mode of DMC develops a sideband with the increase of LiClO4 concentration. Its analysis using a Voigt profile was used to quantify the ionic diffuse transport.

    Using high-speed and high-definition in situ Raman imaging, several groups have investigated the local SOC and state-of-health (SOH) distribution of electrodes. Nishi et al. [49] visualized the SOC distribution in LiCoO2 at various rates and showed the limitation of ionic transport. Nanda et al. [50] investigated the SOC inhomogeneity in fresh and aged layered cathodes, while Slautin et al. [51] studied the degradation paths in LiMn2O4 spinel structures. As an example of experimental procedure, Raman maps of the LiNi0.8Co0.15Al0.05O2 (NCA) positive electrode were performed by collecting spectra in a range of Raman shift from 250 to 1800 cm−1 across a 10 μm square region of the electrode surface [50]. Results obtained from the composite electrode NCA + carbon black + PVdF binder are shown in Figure 3. Typical map is obtained with a minimum of 1000 data points. Each spectrum was treated considering three Raman peaks of NCA centered at ca. 475,550 and 620 cm−1 attributed to the vibrations of Ni–O bonds, while bands of the carbon are observed at ~1350 cm−1 (D band), 1500 cm−1 (amorphous carbon) and 1590 cm−1 (G band).

    Figure 3. Raman mapping and SOC surface analysis of NCA composite cathode. (a) Map showing the spectral intensity as the total area under the NCA Raman peaks. The solid (orange) arrow points to NCA particle interior and the dotted arrow points to interface between particle and carbon-filled binder. (b) SOC map derived from analysis of NCA Raman peaks. (c) Representative Raman spectra of the NCA active particles indicated by the red arrows for two values of the SOC parameter. (d) Histogram showing the frequency distribution of the SOC parameter. The quantity on the abscissa (A475/A550) × (I475/I550) is a measure of the SOC, as described in the text. Electrode was charged galvanostatically to 4.2 V at 3C rate without any potentiostatic step. Reprinted by permission from Ref. [50].

    3. Positive electrode (cathode) materials


    3.1. Rock-salt-type compounds

    Structural properties of layered-type compounds such as LiCoO2, LiNiO2, LiNixCo1−xO2, LiNi0.8Co0.1.5Al0.05O2 as well as LiNixMnyCozO2 with various Co contents were widely investigated by in situ X-ray diffraction [52]. These compounds crystallize with a lamellar structure of rock-salt-type of R-3m space group. For instance, the poor electrochemical cycling behavior of Ni-rich NMC (LiNi0.5Mn0.2Co0.3O2) cathode is due to the structural phase transition between hexagonal phases H2 to another hexagonal phase H3 (at 4.7 V) upon Li+ de-intercalation, which can be suppressed in long-term cycleability.


    3.1.1. LiCoO2 (LCO)

    LiCoO2 prepared by solid-state reaction or by wet chemistry followed by a post-annealing treatment at 800–900 ℃ (denoted HT-LCO) crystallizes in the rhombohedral system R-3m with the α-NaFeO2 structure built by alternating ordered layers of CoO6 and LiO6 octahedra along the c-axis. At a lower annealing temperature, Ta ≈ 400 ℃, LCO crystallizes in the cubic NaCl-type structure (Fd3m space group) (LT-LCO) with cations distributed among octahedral sites of the cubic oxygen close-packed array. While the HT- and LT-LCO structures cannot be efficiently discriminated by X-ray diffraction, Raman spectroscopy has been successful to determine the phase of LiCoO2 synthesized in the temperature range 400–900 ℃ [16].


    3.1.1.1. Phase identification of Li1−xCoO2

    The phase identification of Li1−xCoO2 as a function of the Li extraction is currently established via the A1g and Eg phonon Raman active (gerade) modes (Table 1), which are observed at 596 and 487 cm−1, respectively. The A1g mode is the symmetrical stretching of Co–O (atomic displacement parallel to c-axis) whereas the Eg mode is the symmetrical deformation (atomic displacement perpendicular to c-axis). Gross and Hess [32] demonstrated the presence of the resonance Raman effect for LiCoO2 cathode under electrochemical conditions upon switching the laser excitation by evaluating the A1g/Eg integrated intensity ratio. A1g/Eg increases from 0.9 (for λ = 632.8 nm) to 2.3 (for l = 532 nm) due to resonance process involving the dd electronic transition from Co-t2g to Co-eg bands (optical absorption at 2.1 eV (591 nm)) and under resonance conditions an overtone of the A1g band occurs [12]. The resonance enhancement for LCO materials was also pointed out when using a green laser excitation for in situ studies [32].

    Table 1. Atomic position, site symmetries and Raman active modes for LT- and HT-LiCoO2 compounds.
    Space group Atom Wyckoff positions Raman modes
    R-3m Li (0, 0, ½) 3b A1g + Eg
    Co (0, 0, 0) 3a
    O (0, 0, ¼) 6c
    Fd3m Li (0, 0, 0) 16c A1g + Eg + 2F2g
    Co (½, ½, ½) 16d
    O (¼, ¼, ¼) 32e
     | Show Table
    DownLoad: CSV

    Several articles have reported in situ RS experiments of LCO electrode materials [2,14,20,31,32,53,54,55,56,57,58,59] but only few of them mentioned the phase transformation upon lithium extraction/insertion. The early work by Inaba et al. [20] showed the change in the Raman spectra of Li1−xCoO2 prepared by electrochemical extraction of Li ions. Results indicated a line broadening of a second hexagonal phase and the monoclinic phase. Novák et al. [2] addressed the rapid disappearance of the A1g band with increasing potential (V > 3.95 V). This Raman band emerges again on discharge (at ca. 3.7 V) with a lower intensity, which was attributed to a lithium-deficient LCO electrode.

    As shown in Figure 4, the phase evolution during the charge of Li1−xCoO2 (lithium extraction) is as follows: (ⅰ) for 0 ≤ x ≤ 0.1 the two phonon (normal) modes at 596 and 487 cm−1 are invariant and correspond to the hexagonal H1 phase, (ⅱ) two new Raman bands located at 473 and 572 cm−1 appear at x > 0.1 due to the formation of the hexagonal H2 phase that coexists with H1 up to x ≈ 0.3, i.e., in a two-phase system, (ⅲ) for x > 0.3 the LCO-type normal modes disappeared while the modes of the H2 phase shift toward lower wavenumbers to 462 and 567 cm−1, respectively, due to the expansion of the Co–O bonds along the c-axis (Figure 5). It is worthy to note that the intensities of the Raman peaks decrease significantly due to the transition from semiconductor to metal for x > 0.15. Novák et al. [2] suggested an increase of electrical conductivity by about two orders of magnitude in Li1−xCoO2 for x > 0.04 correlated with the change of colour of the particles from dark to light. Similar description was previously reported by Inaba et al. [20], while Itoh et al. [14] conceived the formation of a Raman-inactive phase. Song et al. [15] investigated the phase change of LCO films deposited on to Co substrate using a spot of 1 μm of 514.5 nm laser irradiation by in situ RS. A remarkable phase change of the hexagonal film deposited at 200 ℃ to cubic spinel is due to reaction of LCO films with the Co substrate forming a Co-rich phase or Li deficient phase Li1−xCoO2. The laser power (≤50 mW·cm−2) provokes a shift accompanied by broadening and depressed intensity of the hexagonal A1g (589 cm−1) and Eg (478 cm−1) bands, and the occurrence of two new bands at ca. 523 and 681 cm−1 that correspond to the cubic spinel Fd3m phase. In the harmonic approximation, the phonon frequencies are given by:

    Figure 4. Raman spectra of the Li1−xCoO2 electrode recorded during the charge process.
    Figure 5. Phase diagram of the Li1−xCoO2 electrode.
    ν(k/μ)1/2 (2)

    where k is the force constant and μ the reduced mass (μ−1 = mLi−1 + (2mO)−1). As Li and Co co-exist in the Li sites, μ becomes larger, which results in band softening and broadening. Such a structural transformation has been confirmed by FTIR and RS spectroscopy of sol-gel LCO powders [60] showing the growth of the rock-salt structure (R-3m symmetry) upon annealing heat-treatment in the range 400–900 ℃. Upon prolonged cycling, degradation of LCO cathodes is evidenced by disappearance of the peak at 579 cm−1 and the emergence of new bands at 515 and 674 cm−1 assigned to the mode vibration of Li2O and Co3O4, respectively, both being electrochemically inactive components [61].


    3.1.1.2. Surface modification upon cycling

    Spatially-resolved in situ Raman diagnostics were performed to analyze the variation of chemical composition of composite electrodes based on the mixture of LCO active particles, carbon additive and polyvinylidene fluoride (PVdF) binder, and the degradation mechanism of materials [12,56,58]. Taking into account the resonance enhancement for LCO materials, i.e., maximum for green laser excitation, the wavelength-dependent spatially resolved analysis demonstrated significant variation of chemical composition across the electrode surface [32]. Using the same technique, Fukumitsu et al. [57] investigated the SOC distribution in cross-section of LiCoO2 cathode and observed the inhomogeneity of the active particles where Li+ ions did not completely return after discharging. Such an inhomogeneous distribution and the mapping of the LCO electrode degradation were also correlated with battery performance by several groups [31,55,56,57]. Hausbrand et al. [12] addressed the degradation aspects upon overcharge of LiCoO2/PVdF/carbon black composite (85:10:5) electrodes by in situ RS imaging the A1g phonon band of LCO. After cycling at C/12 rate, the composite electrode showed a redistribution of the chemical composition that consists in the formation of surface layers; thus, the loss of electrical contact between LCO particles were evidenced by mapping, while Gross and Hess claimed no significant structural changes observed from in situ RS experiments [32]. Nishi et al. [49] investigated the local SOC distribution of LCO cathode. The results showed that individual LCO particles were discharged at various rates with inhomogeneous current distribution and part of them was irreversibly charged upon subsequent cycles due to the increasing electrical resistance between LCO grains. As an example, Figure 6 displays the Raman images recorded during the first and second cycle of LCO (charge at 0.4C and discharge at 1C) in the potential range 2.5–4.4 V vs. Li+/Li. This electrode shows an inhomogeneous distribution of the SOC of the particles upon cycling.

    Figure 6. Raman images recorded during the first and second cycle of LCO (charge at 0.4C and discharge at 1C) in the potential range 2.5–4.4 V vs. Li+/Li. The SOC distribution is indicated by arrows: particles charged and discharged at slower rate during both cycles (red), particles charged at slower rate in the first cycle (yellow) and particles charged and discharged at normal rate during the first cycle but at a slower rate in the second cycle (blue). Reprinted by permission from Ref. [49].

    Recently, Otoyama et al. [59] reported the Raman mapping of the composite electrode, i.e., mixture of active LiCoO2 and Li2S–P2S5 solid electrolyte before and after the initial charging reaction. While the Raman images show the expected distribution of reactions after full charge test, several LCO particles do not display structural changes. Kostecki et al. [47] reported similar studies performed on LiNi0.8Co0.15Al0.05O2 (NCA) composite cathode material. Observation of the Raman microscopy mapping of the electrode surface showing nonuniform electrode state-of-charge was correlated with the significant impedance rise and capacity fade of the lithium-ion cell graphite//NCA.


    3.1.2. Mixed layered oxides LiNiaMnbCo1−a−bO2 (NMC)

    Mixed layered oxides, namely lithium nickel-manganese-cobalt oxides LiNiaMnbCocO2 with c = 1 − ab (named NMC) are solid solutions of the binary LiMO2 (M = Ni, Mn, Co) compounds. They are identified by the cation stoichiometry, for instance NMC532 for LiNi0.5Mn0.3Co0.2O2. NMC materials crystallize with the α-NaFeO2 structure belonging to the R-3m (D3d5) space group. The local structure of LiNi1/3Mn1/3Co1/3O2 (NMC333) synthesized by wet chemistry with various chelate to metal-ion ratio was studied Raman spectroscopy [62]. Analysis of the intensity A1g and Eg modes by integral of the Lorentzian individual band allows to estimate the cationic mixing, i.e., partial occupancy of Ni2+ in Li sites. The intensity of the Ni–O and O–Ni–O vibrations (ν1 and ν4, respectively) are smaller in the sample with higher cationic mixing. The larger width of these ν1 and ν4 vibrations means that the lifetime of phonons is shorter, another evidence of the larger Ni cationic disorder, in agreement with the structural analysis. An important issue for NMC materials is the cation mixing, i.e., filling of Li sites by Ni2+ ions. Employing a combination of XRD and spectroscopic methods (Raman and NMR spectroscopy), Ben-Kamel et al. [63] reported the effect of varying Co content on the local structure of as-synthesized LiNixCoyMnzO2 oxides and showed that the cation mixing effect during cycling decreases with increasing Co content, illustrating the usefulness of Raman spectroscopy in this issue. The effect of high voltage charging on NMC532 cathodes has been investigated by in situ Raman mapping by analyzing the high-frequency phonon mode. Above the potential 4.4 V vs. Li+/Li, the Raman component at 595 cm−1 shifts above 600 cm−1, which corresponds to a capacity loss similar to the one observed in spinel LiMn2O4 [64]. The local charging profiles of NMC532 particles during cycling was studied by construction of high-spatial resolved SOC images of 35 μm square portion of electrode. The delithiation process provokes a downshift in the frequency of A1g and Eg peaks, accompanied with a decrease in the A1g peak in intensity and an increase in the Eg peak intensity. The local topography was obtained from data by numerically extracting the frequency that highlights the inhomogeneity in the SOC distribution when the cathode is charged further to 4.21 V [35]. Raman mapping was conducted for NMC333 composite cathode in solid-state batteries. Using 75Li2S–25P2S5 solid electrolyte, mapping images displayed uniformly delithiated and lithiated NMC particles [65]. The ratio I595/I474 was used to determine the electrode state-of-charge with I595/I474 = 4.0 for low SOC and I595/I474 = 0.85 for high SOC.

    In situ Raman spectroscopy of Ni-rich Li1+x(NiyCozMnz)wO2 (0.005 < x < 0.03; y:z = 8:1, w is nearly 1) electrodes demonstrated the structural transformation using an original pouch cell with sodium borosilicate glass window [36]. Figure 7a presents the deconvoluted Raman spectrum of a NMC811 electrode in the spectral range of 400–650 cm−1, which shows six active Raman modes according to the approach proposed by Julien and co-workers [63]. In situ Raman spectra collected during Li+ extraction (charge process) in the electrode potential from 3.8 to 4.3 V vs. Li+/Li are shown in Figure 7b. Spectral changes are correlated with the structural transitions from initial hexagonal phase H1 to phase H2 and to coexisting phases H1 + H2. The bandwidth (FWHM) of A1g(Ni) and Eg(Ni) modes remained unchanged upon the potential charge 3.7–3.8 V corresponding to the H1 phase, while increased up to a maximum at 3.9–4.0 V (phases H1 and H1 + H2) and further decreased in the potential range 4.0–4.3 V implying a structural transition from disorder to order that occurs for x > 0.56 in LixNi0.8Mn0.1Co0.1O2 electrode. During the discharge, the Raman response shows quite reversible behavior with low deviations from the original ν and FWHM values.

    Figure 7. (a) Raman spectrum of a NMC811 electrode with the deconvolution of six vibrational components. (b) In situ Raman spectra collected during Li+ ion extraction (charge process) in the electrode potential from 3.8 to 4.3 V vs. Li+/Li. Reprinted by permission from Ref. [36].

    3.1.3. Li-rich layered–layered oxides (LLNMC)

    The exact nature of Li-rich layered-layered cathode materials (LLNMC) of formula Li1+δ(NiaMnbCo1−a−b)1−δO2 or xLi2MnO3·(1 − x)LiMO2 (with M = Ni, Mn, Co) is still the subject of debate; two models are proposed, one considering the structure of a lamellar compound with Li2MnO3 domains, the other proposing a solid solution as an integrated structure of Li2MnO3 (C2/m) and LiMO2 (R-3m) layered components. For δ > 0, the overlithiated oxide are also known as high-energy NMC (HE-NMC). Li2MnO3 is electrochemically activated during the first charge at a potential plateau of 4.5 V. Depending on the composition and charging conditions, LLNMC is enabling to deliver more than 250 mAh·g−1 at potential of 5 V vs. Li+/Li. This class of materials is often noted xLi2MnO3·(1 − x)LiMO2 based on the X-ray diffraction patterns. Among this series, 0.5Li2MnO3·0.5LiNi0.5Mn0.5O2 is equivalent to Li1.2Ni0.2Mn0.6O2 and 0.5Li2MnO3·0.5LiNi0.33Mn0.33Co0.33O2 to Li1.2Mn0.54Ni0.13Co0.13O2. Both of them are promising cathode materials. The layered structure of Li2MnO3 or Li(Li1/3Mn2/3)O2 is described in the monoclinic C2/m (C2h3) space group; there is an extended cation ordering in the slabs with the Li/Mn ratio of 1/2 with Li ions fully occupying one 2b Wyckoff site of the slabs along the c-axis, while Mn ions occupy the 4g site. Factor group analysis predicts 15 Raman active modes as 7Ag + 8Bg.

    Lanz et al. [66] stated that, prior to charging, i.e., in the open-circuit voltage (OCV) state, the LLNMC framework is a mixture of the Li2MnO3 and layered LiMO2 phases observed by in situ XRD and in situ Raman spectroscopy. The local structure of Li-rich NMC with the composition Li1.20Mn0.54Co0.13Ni0.13O2 studied by Raman spectroscopy was compared with that of layered compounds LiCoO2, LiNi1/3Mn1/3Co1/3O2 and Li(Li1/3Mn2/3)O2 (or Li2MnO3). This comparison supports that Li1.20Mn0.54Co0.13Ni0.13O2 is a solid solution of Li2MnO3 and LiNi1/3Mn1/3Co1/3O2 instead of separated domains of these components [67]. However, an additional Raman peak located at ~660 cm−1 was attributed to the vibrational mode of Mn–O bond also observed in other layered phases such as HT-Li0.52MnO2.1(with I41/amd symmetry) [68] simply shifted by the fact that the Mn–O bond is shorter in LLNMC.

    In situ Raman spectroscopy of Li-rich layered oxides xLi2MnO3·(1 − x)LiMO2 (M = Ni, Co, Mn) was carried out for several compositions x = 0.3, 0.5, and 0.7 [37,66,69,70,71,72,73]. The general trends consisted in the determination of the Li2MnO3 activation during the first cycle characterized by a modification of the ionic local coordination and an increase of the ionic disorder of the Li-rich layered structure. Singh et al. [69] reported in situ Raman spectra of Li1.2Ni0.175Co0.1Mn0.52O2 during galvanostatic charge and discharge process between OCV (~3.0 V) and 4.8 V. During the charge, a new peak emerges at ~544 cm−1 due to the oxidation of Ni2+ ions, and a change in the Raman peak at 445 cm−1 (up to 4.1 V) reflects the extraction of Li+ ions from the transition-metal layer, contrary to the previous claim of a movement of Li+ ions to vacant site at lower potential 2.92 V [74]. It was concluded that, in the voltage range 4.1–4.4 V, lithium ions are removed from both the transition-metal layers and lithium layers. At upper state of charge at 4.55–4.6 V, the oxygen removal becomes severe, and associated with the diffusion of the transition metal ions into the vacant sites. In situ Raman spectroscopy was applied to characterize the electrochemical activation of Li2MnO3 in LLNMC cathodes. The fingerprint of this activation lies in the main peak at 615 cm−1 (Ag band of Li2MnO3 phase), which shifts to higher wavenumbers ~630 cm−1 upon of charging above 4.4 V vs. Li+/Li [70]. The same group of researchers reported the emergence of the new Raman band at ~545 cm−1 for stoichiometric Li1.1(Ni1/3Co1/3Mn1/3)O2.1 during the first charging process at low voltage in the range 4.1–4.3 V vs. Li+/Li. The activation of Li2MnO3 possibly results in the formation of Li2O and MnO2, the former being responsible for the band at ~545 cm−1 [66]. This assumption was previously proposed by Hy et al. [72] by studying the oxygen-related surface reactions occurring on Li1.2Ni02Mn0.6O2 using in-situ surface enhanced Raman spectroscopy (SERS). Li2O was directly detected during the extended voltage plateau at 4.5 V that leads to the hydrolysis of the electrolyte and the formation of Li2CO3. Rao et al. [71] reported the spectral modifications of the Li1.2Mn0.54Ni0.13Co0.13O2 electrode during charging from OCV to 4.7 V (Figure 8). While the Raman spectrum is invariant until the charge at 4.2 V, a new broad peak due to the Ni3+–O bonds appears at ca. 542 cm−1 at 4.25 V, which becomes sharp in the potential range 4.35–4.47 V. Other notable changes associated with the removal of Li+ ions consist in the vanished 445 cm−1 peak and a shift of the 490 cm−1 peak due to the increase of the interlayer distance. Finally, in the potential region 4.25–4.5 V, a strong increase of the Raman signal background is attributed to the Li2MnO3 activation indicating a local environment modification with simultaneous lithium extraction and oxygen release. During discharge, the Raman spectrum returns to the initial situation with suppression of the background intensity of the peak.

    Figure 8. In situ Raman spectra of Li1.2Mn0.54Ni0.13Co0.13O2 cathode cycled at C/20 rate. Reprinted by permission from Ref. [71].

    Particular attention has been devoted to the effects of cycling on the structure of lithium-rich NMC cathode compounds, and of Li2MnO3–LiMO2 (M = Ni, Mn, Co) composites (LR-NMC), the later being an integrated structure of the C2/m (Li2MnO3) and R-3m (LiMO2) layered components. Several investigators have shown the gradual transformation of Li-rich NMC materials cycled from a discharge state at 2.0 V to the charging voltage as high as 4.9 V [75,76]. Using a combination of XRD, TEM and ex situ Raman spectroscopy, Amalraj et al. [75] demonstrated a partial layered-spinel structural transition in the initial charge to 4.7 V vs. Li+/Li. Similar results on carbon coated Li1.2Mn0.56Ni0.16Co0.08O2 were reported by Kumar-Nayak et al. [76]. The Li-rich NMC Raman features were interpreted in terms of a gradual layered to spinel transformation provoked by electrochemical cycling.

    Long-term cycling of high-energy sol-gel synthesized cathodes, i.e., Li-rich layered-layered oxides xLi2MnO3·(1 − x)LiNi2/3Co1/6Mn1/6O2 (x = 0.3, 0.5 and 0.7) was investigated by in situ Raman spectroscopy for 35 cycles. During the rate capability measurements at different current densities, the A1g (602 cm−1) and Eg (490 cm−1) phonon modes were unchanged without the occurrence of 630–670 cm−1 peaks corresponding to the phonons mode of the cubic spinel-like phase. The increase of the bandwidth of the phonon modes after electrochemical tests was attributed to the formation of the SEI layer [73]. Huang et al. [37] reported similar behavior and showed that, using in-situ electrochemical Raman spectroscopy correlated with the differential capacity (dQ/dV) curves, the oxidation reactions Ni2+ → Ni3+ → Ni4+ appeared in the potential range 3.70–4.45 V with a good reversibility in Li-rich 0.5LiNi0.5Mn0.5O2·0.5Li2MnO3 material. Recently, the study of the fate of Li2MnO3 (C2/m) from a commercially produced LLNMC cathode material having the composition 0.49Li2MnO3·0.51LiNi0.37Co0.24Mn0.39O2 (Toda Kogyo Corp. HE5050) adopted the domain model. The best spectral response in terms of signal-to-noise, was obtained using the 785 nm laser excitation rather than lower wavelengths [77]. On charge, the C2/m phase was delithiated at the same rate as the R-3m phase, while on discharge the monoclinic component was reformed latter.


    3.1.4. LiNi0.8Co0.15Al0.05O2 (NCA)

    Aluminum-doped Ni-rich layered structure, i.e., LiNi0.8Co0.15Al0.05O2 (named NCA) is considered as a promising cathode material for high-power batteries with a high specific discharge capacity of ~180 mAh·g−1 at 1C-rate and a superior thermal stability [78]. Increase of the Li+ ion diffusivity in the layered R-3m lattice is due to the aluminum substitution which enlarges the c-axis parameter [79]. In situ Raman microscopy was investigated to evaluate the kinetic features of Li extraction/insertion of individual NCA particles [24], the SOC distribution for deep discharge regime [47], and the NCA electrode degradation upon long term cycling [11]. Using a sealed spectro-electrochemical cell, in situ monitoring of the SOC of NCA composite electrodes (84 wt% NCA, 4 wt% carbon black, 4 wt% SFG-6 graphite and 8 wt% PVdF) showed that the rate of charge and discharge of a selected particle is dependent of the time of the process and location in the electrode. In situ Raman spectra collected using the 632 nm laser line are dominated by a group of bands in the spectral range 470–580 cm−1 characteristic of the Ni–O bonds in NCA oxides. It is assumed that the integrated peak ratio I475/I553 is proportional to the amount of lithium in one particle, which allows the measure of SOC in individual particle. However, the plot of I475/I553 versus electrode potential does not show symmetric behavior during charge-discharge cycle, due to the non-uniform particle distribution [24].

    Performing Raman microscopy mapping of NCA cathode material in a pouch-type lithium-ion cell, Kostecki et al. [47] showed that the SOC distribution of particles was non-homogeneous despite the cycling at deep discharge regime. The NCA electrode degradation upon long term cycling was also studied by the same group [11]. Raman image was collected on an area 52 μm × 75 μm using the red excitation source (632.8 nm) with a spot of ~1.2 μm of power adjusted at 0.1 mW exhibiting the Raman features of NCA (peaks at 475 and 554 cm−1) and carbon (D- and G-band at ~1350 and ~1600 cm−1, respectively (Figure 9a). As shown in Figure 9b, the Raman image displays different colored pixels, red, blue, and green corresponding to the integrated band intensities of NCA, and D, G carbon bands of single spectrum, respectively. Figure 9c shows that the NCA spectral response, the intensity band ratio I475/I554 and the peak FWHM are position dependent. These results evidence the spectral distortion due to the variation of the excited electronic states. Similar results obtained from mapping of Li1−x(NiyCozAl1−y−z)O2 maintained at different SOCs were explained in terms of a link between the variation of composition and the structural integrity at the single particle level [50].

    Figure 9. (a) Average Raman spectrum of the fresh composite LiNi0.8Co0.15Al0.05O2 electrode, (b) Raman image of the composite cathode where red, blue, and green pixels correspond to the integrated band intensities of Eg and A1g modes of NCA, and D-, G-band of carbon of each spectrum, respectively; (c) Raman microscope spectra of three individual NCA particles. Reprinted by permission from Ref. [11].

    3.2. Spinel-type compounds


    3.2.1. LiMn2O4 (LMO)

    Among the three-dimensional frameworks, spinel LiMn2O4, is an interesting cathode material as it is a low-cost material with low environmental impact. It generally exhibits reduced volume expansion upon lithium extraction/insertion reaction. LiMn2O4 grows in the cubic structure (Fd3m space group), in which cations occupy octahedral (Mn: 16d) and tetrahedral (Li: 8a) sites in the cubic close-packed array of oxygens occupying 32e Wyckoff positions. Extraction of Li ions from Li1−xMn2O4 (0 < x < 1) occurs with two voltage plateaus at ca. 3.95 and 4.15 V vs. Li+/Li in between the LiMn2O4/Li0.5Mn2O4/λ-MnO2 end phases that correspond to the oxidation of Mn3+ to Mn4+ [80]. Li1−xMn2O4 is a stable phase: the removal of Li from Li1−xMn2O4 (oxidation process) occurs with a gradual decrease of the unit cell volume through a two-step redox reaction at the two voltage plateaus. However, the identification of the phases during Li extraction from LiMn2O4 to λ-MnO2 has been debated in several publications involving two- or three-phase structural models [81,82,83,84,85,86].

    A complete factor group analysis of the vibrational modes of the ideal AB2O4 spinel structure was given by White and DeAngelis [87] and revisited by Ammundsen et al. [88]. In the Oh7 symmetry there are five Raman active modes represented by:

    ΓLiMn2O4=A1g+Eg+3T2g (3)

    which decomposes in A1g + Eg + 2T2g as internal modes and 1T2g as an external (translational) mode. The experimental Raman spectrum displays 6 peaks distributed as follows: (ⅰ) the low-frequency peak at ca. 162 cm−1 is the T2g(1) translational mode that derives from the Li sublattice vibration, (ⅱ) the Eg and T2g(2) modes are observed at 434 et 455 cm−1, respectively, (ⅲ) due to the presence of mixed valence state of Mn (Mn3.5+) in LiMn2O4, the A1g mode related to the Mn–O symmetric stretching mode is split into two components at 577 and 625 cm−1 attributed to the vibrations ν(Mn+Ⅳ–O) and ν(Mn+Ⅲ–O), respectively, (ⅳ) the high-frequency band at 654 cm−1 is the T2g(3) mode predominantly characterized by large oxygen displacements.

    Investigations of the phase evolution of LiMn2O4 during charge–discharge process have been reported by several groups using in situ Raman spectroscopy [89,90,91,92,93,94]. The main spectral fingerprints of the spectra of LiMn2O4 and λ-MnO2 consist in intense peaks at 593 and 628 cm-1. Kanoh et al. [89] studied the electro-insertion of Li+ in λ-MnO2 using in situ Raman spectroscopy, and demonstrated that the logarithm of the intensity ratio log(I593/I628) is an accurate measure of the variation of the lithium content, x, in Li1−xMn2O4 spinel lattice. In their early work, Huang and Frech [90] investigated the phase diagram of stoichiometric Li1−xMn2O4 and non-stoichiometry Li1.1−xMn2O4 cathode materials. Raman spectra of the former electrode (0.1 < 1 − x < 1.0) showed a single-phase followed by a two-phase process, and finally by another single-phase reaction, whereas the latter electrode displayed the presence of λ-MnO2. Modifications of the vibrational properties of a single microparticle of Li1−xMn2O4 were monitored in situ via simultaneous Raman spectroscopy and cyclic voltammetry. In the potential range 4.021–4.118 V, a markedly asymmetric band centered at ca. 600 cm−1 are attributed to the T2 and A1 modes of the Li0.5Mn2O4 (half-charged) phase, whereas two prominent bands located at 495 and 588 cm−1 are due to the T2g(2) and A1g modes of λ-MnO2 (full charged), respectively, for electrode voltage higher than 4.131 V [92]. Shi et al. [93] reported similar Raman experiments on LiMn2O4 single crystal microelectrode and showed that the SOC derived from optical data displays two well-defined steps up to fully oxidized material. Anzue et al. [93] studied the effect of the excited laser lines of 647.1 nm (1.97 eV), 514.5 nm (2.48 eV) and 457.9 nm (2.71 eV) on in situ Raman spectra of Li1−xMn2O4. They found a resonance enhancement effect for x = 1 (λ-MnO2). This energy dependence attributed to the variation of the bandgap as a function of the Li content. Using a confocal Raman microscopy method, Slautin et al. [51] investigated the degradation paths in LiMn2O4 electrodes. Upon cycling, it is shown that the Mn3O4 phase is formed and dissolved that induces a local disturbance of the lithiation state.

    Figure 10 shows the in-situ Raman spectra recorded using the 532 nm laser excitation line of Li1−xMn2O4 electrode during the charge at C/20 rate as a function of the electrode potential in the range 3.2–4.45 V vs. Li+/Li [91]. The evolution of the frequency shift of Raman peaks that describes the Li1−xMn2O4 phase diagram in the composition range 0 ≤ x ≤ 0.88 is presented in Figure 11. Three distinct single phases are identified from the spectral features during the charge process: (ⅰ) the phase of the initial compound LiMn2O4 (phase 1) is maintained up to x = 0.10, (ⅱ) this phase coexists with an intermediate phase (phase 2) up to x = 0.6, (ⅱ) above which a single intermediate phase (phase 2) is identified in a short range of composition 0.6 < x < 0.66. (ⅲ) a two-phase system between phase 2 and phase 3 is then observed in the range 0.7 ≤ x ≤ 0.8, (ⅳ) the single Li-poor phase (phase 3) is observed for x > 0.80. Note that the intermediate phase 2 has been previously reported from X-ray diffraction [86] and by Raman spectroscopy [88,91] analysis. Ammundsen et al. [88] identified this phase as an ordered structure Li1−xMn2O4 (0.5 ≤ x ≤ 0.7) with lower symmetry (space group F-43m, point group Td2), due to a distortion of the initial cubic lattice and vacancies in Li tetrahedral sites. According to White and DeAngelis, the ordered cubic spinel (Ⅲ) displays 12 Raman active modes represented by ΓLi0.5Mn2O4 = 4A1 + 3E + 6T2. It is worthy to make the following remarks: (ⅰ) using neutron and ex situ X-ray diffraction studies, the intermediate phase was only observed by Liu et al. [82] due to the difficulty to distinguish bi-phased domains at low C-rate (<C/10). On another hand, Raman spectroscopy is predominantly a weak penetrating probe, which investigates the delithiated phases formed at the surface of particles. (ⅱ) It seems that the phase diagram in the 4-volt region is sensitive to the Li stoichiometry; as shown by in situ X-ray diffraction [95] lithium-deficient Li0.99Mn2O4 and lithium-rich Li1.04Mn1.93O4 show three cubic phases with the existence of two-phase domains due to the overlap of the adjacent phase.

    Figure 10. In situ Raman spectra of Li1−xMn2O4 electrode during the charge at C/20 rate as a function of the potential electrode in the range 3.2–4.45 V vs. Li+/Li. Spectra were recorded using the 532 nm laser excitation line.
    Figure 11. Evolution of the frequency shift of Raman peak describing the Li1−xMn2O4 phase diagram in the composition range 0 ≤ x ≤ 0.88.

    Micro-Raman mapping was used to investigate the local structure of Ce-doped LiMn2O4 as shown in Figure 12a with the Raman spectra of Li–Mn–O components (Figure 12b). By probing the A1g mode shift, spectral components observed in Figure 12a are as follows: (ⅰ) the largest area (blue regions) represents the pristine spinel (628 cm−1), (ⅱ) some spots originate from the monoclinic Li2MnO3 phase (616 cm−1, white spots), (ⅲ) a small fraction of distorted lattice spinel is shown by green spots (646 cm−1) and (ⅳ) a small fraction of signal comes from CeO2 particles [96].

    Figure 12. (a) in situ Raman mapping of Ce-doped LiMn2O4, (b) Raman spectra of Li-Mn–O components. Reprinted by permission from Ref. [96].

    3.2.2. LiNi0.5Mn1.5O4 (LNM)

    LiNi0.5Mn1.5O4 spinel (denoted LMN hereafter) is considered to be a promising high-voltage (4.7 V vs. Li+/Li) cathode materials for powering hybrid electrical vehicles (HEVs) and plug-in hybrid electric vehicles (PHEVs). However, this material suffers from the presence of Jahn-Teller Mn3+ ions, cation ordering, and oxygen vacancies, which affect its high-rate performance. Recent investigations showed that the partial substitution of Ni and Mn by cations, such as Cr, Co, Fe, Nb, Mg, Zn, etc. [97,98,99,100,101,102,103] stabilizes the disordered LNM phase (Fd-3m space group) and eliminates the impurities of LiNiOx, leading to electrochemical improvements. Typical Raman spectra of pristine and doped LNM electrodes are presented in Figure 13 with attribution of the spectral features [103].

    Figure 13. Raman spectra of as-made LNM electrodes. Reprinted by permission from Ref. [103].

    The in-situ XRD studies on the phase evolution of LNM cathodes during cycling reveal that the doped materials have identical phase diagrams showing three cubic phases and the patterns reveal a decrease of the amount of Mn3+ Jahn-Teller ions leading to an increase of the rate capability [101]. In-situ Raman spectroscopy of doped LiNi0.5Mn1.5O4 spinel cathodes was employed to investigate the valence state variations of nickel and manganese, as well as the local structure change during galvanostatic charge–discharge [102,103]. The early report of the in-situ Raman spectra of LiNixMn2−xO4 (0 < x < 0.5) thin films collected in the potential range 3.4–5.0 V vs. Li+/Li showed a Raman band located at 540 cm−1 appearing at ~4.7 V, assigned to the Ni4+–O bond. It was suggested that, upon cycling, the redox reactions Ni2+/3+ and Ni3+/4+ occurred in the potential range 4.4–4.7 V and 4.7–5.0 V, respectively [102].

    Figure 14 shows the Raman spectra of Cr- and Co-doped samples collected between 3.5 and 4.9 V [103]. In the wavenumber range of 100–800 cm−1 the Raman spectra contain the bands of the pristine LNM with eleven features: (ⅰ) the bands at 162 and 220 cm−1 are attributed to the translation mode of molecular vibration, (ⅱ) the 486,593 and 639 cm−1 bands are assigned to the stretching mode of Mn–O bond, (ⅲ) the vibration modes at 408,498,528 and 613 cm−1 originated from the Ni–O bond. In addition, a new band is introduced by the doping at 672 cm−1 and is attributed to A1g mode of Cr3+–O/Co3+–O. Near the end of charge process at ca. 4.80 V, several new bands are detected among which (ⅰ) the T2g(T) band at 170 cm−1 is attributed to the translation mode, and (ⅱ) the T2g band at 538 cm−1 is due to the presence of Ni4+–O bond in the delithiated lattice. These Raman bands are clearly observed at V ≥ 4.78 (x ~ 0.32) and V ≥ 4.82 (x ~ 0.28) for Cr- and Co-doped LMN, respectively. The high-wavenumber band at 672 cm−1 assigned to Cr3+–O/Co3+–O vibration remains fixed during cycling, which proves that the valence state of Cr and Co ions is invariant [103].

    Figure 14. Evolution of in situ Raman spectra in one charge–discharge cycle at C/12 rate for (a) LiMn1.45Ni0.45Cr0.1O4 and (b) LiMn1.45Ni0.45Co0.1O4. Reprinted by permission from Ref. [103].

    3.3. Vanadium pentoxide (V2O5)

    V2O5 is an attractive electrode material which can accommodate 3 Li per mole of oxide to deliver a specific capacity of 450 mAh·g−1 within the potential range 1.5–4.5 V vs. Li+/Li [104]. Vanadium pentoxide, V2O5, with high-valent V5+ cations, are layered structures which consist of edge and corner-sharing VO5 square pyramids attached by weak chemical bonds favourable to Li insertion process. The phase diagram of LixV2O5 is rather complex with reversible and irreversible phases across the composition 0 < x < 3. Over the composition 0 ≤ x ≤ 3, LixV2O5 undergoes the phase transitions with the corresponding potential plateaus α → ε, ε → δ and δ → γ at 3.4, 3.2 and 2.3, respectively. The γ-phase is irreversibly transformed into the ω-phase at the potential 2.05 V (x > 2). From XRD patterns, it was shown that, upon lithiation from α to to ε-phase, the lattice parameter, a, decreases continuously, while the lattice parameter b remains almost unchanged. In situ Raman spectra of Li intercalated V2O5 have been measured on crystal [19,30,105] and thin films [106,107] to characterize structural changes in LixV2O5 according to different C-rates, but also to study the abnormal stress modification upon phase transitions [107,108]. The first in situ Raman spectrometry studies of electrochemically lithiated and delithiated V2O5 materials were reported by Frech et al. [10,30,105]. Findings of Raman patterns of LixV2O5 electrodes are as follows: (ⅰ) a shift of the low-frequency B1g/B3g Raman peak (translational mode) from 144 to 154 cm−1, (ⅱ) a progressive intensity decay of the spectral response in the range 195–715 cm−1 attributed to the reduction of V5+ to V4+ generating negative charge carriers, (ⅲ) the subsequent reduction of the optical-skin depth due to the metallic nature of the Li-intercalated V2O5 electrode, (ⅳ) the disappearance of the vanadyl stretching mode at 996 cm−1 at electrode potential of ~3.27 V vs. Li+/Li, replaced by a new peak at 983 cm−1 observed at ~3.40 V related to the coexistence of the α- and ε-LixV2O5 phases, (ⅴ) the new peak at 983 cm−1 shifts down to 972 cm−1 for further degree of Li insertion corresponding to the ε-LixV2O5/δ-LixV2O5 two-phase system, (ⅵ) finally, the γ-LixV2O5 phase is characterized by a peak at 975 cm−1 and the disappearance of the peaks at 196,302,484 and 700 cm−1 due to a loss of crystallinity in the γ-phase. Jung et al. [108] reported that the Raman patterns are not detected in the γ → ω phase transformation suggesting that the ω-phase has a quasi-amorphous structure. Studies of Li insertion in LixV2O5 film, suggested that the shift of the low-frequency peak from 144 to 154 cm−1 is due to the increase in the restoring force consequently to the tensile stress in the LixV2O5 film [109]. This assumption was confirmed by a combined approach (micro-Raman and optical stress sensor) correlating the Raman shift of B1g + B3g mixed modes with the more tensile stress as a function of the Li content in the LixV2O5 electrode [108]. The initial V2O5 orthorhombic symmetry (Pmmn space group) is kept upon Li intercalation in the LixV2O5 film (0 ≤ x ≤ 0.94) due to a solid-solution behavior leading to the typical Raman signature of the ε-Li0.94V2O5 phase. During electrode charging, the local structure is completely recovered, the reason for the excellent electrochemical reversibility of V2O5 thinfilms prepared by atomic layer deposition (ALD).

    The changes of the Raman spectra during the discharge process (lithiation) depend on the V2O5 synthesis technique [108]. In the works by the Pereira–Ramos's group [106,110,111,112] the Raman spectra differs for ALD samples, for rf-magnetron sputtering films and for sol-gel powders. Baddour-Hadjean et al. assumed that the decrease in the intensity of the low-frequency B1g/B3g Raman active mode located at 144 cm−1 reflects the structural disorder occurring upon V2O5 lithiation [113]. Therefore, the peak height I144 variation was correlated with the Li content in the LixV2O5 cathode and the dependence of the induced disorder during cycling with C-rates. Jung et al. [107] analyzed the charge rate dependent stress and structure changes in V2O5 electrode synthesized by ALD using an in-situ Raman micro-spectrometer equipped with a 532 green laser combined with a stress monitor. As shown in Figure 15, an exponential dependence of I144 on the degree of Li intercalation is observed in the composition range 0 < x < 0.2. When the cathode is discharged at high C-rate (>C/4), the slope of ln(I144) increases, which reveals an increase of the structural disorder in LixV2O5.

    Figure 15. Plot of the logarithm of the peak height of the 144 cm−1 Raman band as a function of Li content in LixV2O5 (adapted from Ref. [107]).

    3.4. Manganese dioxide (MnO2)

    Manganese dioxide, MnO2 has been extensively studied as an active material for Leclanché dry cells, lithium-ion batteries and pseudocapacitors. The most popular phases used in energy-storage systems are γ-MnO2 (pyrolusite/Ramsdellite intergrowth), α-MnO2 (hollandite or cryptomelane phase) and δ-MnO2 (layered birnessite phase). The crystallization in one of these different phases is determined by the synthesis conditions [114].

    The charge storage mechanism (Li+ ions incorporation) of a pseudocapacitive α-MnO2 electrode was probed using in operando Raman spectroscopy. Raman spectra were obtained using a 514 nm excitation source and a confocal slit of 5 μm. the MnO2 electrode was tested with different cations in the electrolytes (Li+, Na+ and K+) in order to validate the best electrolyte of the pseudocapacitor [115]. Four major active Raman bands labelled ν1 to ν4 located at ca. 640,577,515 and 390 cm−1, respectively, were considered for this purpose, since they describe the intrinsic vibration modes of pristine MnO2. According to Julien et al. [116,117], the ν1 band (620–650 cm−1) is assigned to the symmetric stretching vibration of Mn–O bond in the MnO6 octahedral environment; the ν2 band (570–590 cm−1) is assigned to the Mn–O vibration along the chains of the MnO2 network, whereas the band at 390 cm−1 originates from the Mn–O bending vibrations. Based on the monoclinic structure (C2/m symmetry) of Li-intercalated MnO2, the irreducible representations of the Raman active modes are classified as:

    ΓLiMnO2=10Ag+8Bg (4)

    Figure 16 shows the in-situ Raman spectra evolution of MnO2 film cycled within pseudocapacitive conditions in 2 mol·L−1 LiNO3 aqueous electrolyte between 0.7, 0 and 0.7 V vs. Ag/AgCl. Changes in the Raman spectrum are as follows. (ⅰ) At the beginning of the cycle (0.7 V), water molecules fill the interlayer spacing with Mn in the highest oxidation state (~Mn4+). As the cathodic reaction proceeds (0.7 V to 0.0 V), the position of the ν1 peak shows a red shift while the ν2 band position underwent a blue shift. As pointed out by Julien et al. [118] the band shifts imply a reduction of the interlayer spacing upon the cathodic process due to the replacement of Li+ ions for water molecules. (ⅱ) At 0 V, the spectrum is close to that of Li1.0MnO2 [119]. While the band frequencies ν1 (616 cm−1), ν2 (586 cm−1) and ν3 (480 cm−1) match well with spectral features of Li1.0MnO2, the weak ν4 is observed at 402 cm−1 instead of 420 cm−1, which could be due the lower amount of incorporated Li+ ions. (ⅲ) The ν1 and ν2 bands become broadened due to the structural distortion induced by the increase of Mn3+ Jahn-Teller ions, (ⅳ) the evolution of the peak intensity ratio I2)/I1) (polarizability change) is attributed to the exchange of neutral H2O by charged Li+ cations and (ⅴ) during the cathodic process, the weak band ν3 displays a doublet, while the ν4 band shows reverse behavior [120].

    Figure 16. In situ Raman spectra evolution of MnO2 film cycled in 2 mol·L−1 LiNO3 aqueous electrolyte between 0.7, 0 and 0.7 V vs. Ag/AgCl. Reprinted by permission from Ref. [120].

    Note that the reversible expansion and shrinkage in lattice spacing during charge transfer at manganese sites upon reduction/oxidation of MnO2 demonstrated by the Raman spectroscopy is a direct proof that the pseudocapacitance involves intercalation or insertion of cations into the bulk of the oxide structure and is not limited to only the surface in contact with the electrolyte, as it was thought in the early stage of the study of the charge storage mechanism in this material.

    The origin of high stability of the birnessite δ-MnO2 pseudo-capacitive electrode synthesized by electroplating was elucidated using operando Raman spectroscopy at different states of charge/discharge. In situ spectra were collected using a green laser (λ = 514.5 nm) and a long working distance of 50× objective lens. The electrochemical tests of the electrode were carried out galvanostatically between 0 and 1 V vs. Ag/AgCl at a 60 A·g−1 current density in solution containing 1 mol·L−1 Na2SO4 as electrolyte [121]. During charge/discharge cycling, three major Raman bands ν1 (677 cm−1), ν2 (589 cm−1) and ν3 (507 cm−1) show frequency shifts due to the structural evolution. For increasing applied potential 0.0 → 1.0 V (charge process) the ν2 band at 589 cm−1 shifts down to 572 cm−1, which corresponds to the expansion of the interplanar spacing due to the modification of the oxidation state of Mn4+ occurring for extraction of Na+ ions and incorporation of water from the solution. During charge, the new peak ν1 (677 cm−1) is assigned to an intermediate Mn–O oxide state because of the coexistence of Mn4+ and Mn3+ cations. The reverse mechanism occurs during discharge (Na+ ions insertion). Moreover, theoretical calculations also demonstrated that the incorporation of Na+ into δ-MnO2 is more energetically favorable than H+ at all available sites of the interlayer space.


    3.5. LiFePO4 olivine

    In the last decade, LiFePO4 (LFP), member of the family of ABIIXO4 olivine orthophosphates, has become an attractive positive electrode material for Li-ion batteries because their low cost and excellent thermal stability [122]. According to the group factor analysis of the LFP structure with Pnma space group (D2h16 point group), the 36 Raman active modes at the centre of Brillouin zone are classified as:

    ΓLiFePO4=11Ag+7B1g+11B2g+7B3g (5)

    Raman spectrum of LFP includes: (ⅰ) the high-frequency bands in the region 908–1126 cm−1 that are stretching modes of P–O bonds, (ⅱ) vibrations in the medium-frequency range between 487 and 691 cm−1 that are O–P–O bending internal modes of the PO43− anion, (ⅲ) external modes in the low-frequency region between 175 and 335 cm−1 and (ⅳ) the four peaks between 402 and 445 cm−1 correspond to the Li cage modes in octahedral coordination with O2− anions [123]. Wu et al. [124] studied the evolution with x of in situ Raman spectra of LixFePO4 samples (non-carbon coated) with different morphologies, i.e., nanoparticles and bulk particles, during galvanostatic charge/discharge process. The phase change from LiFePO4 to FePO4 is characterized by the shifts in the internal modes (stretching region), the appearance of peaks at 175 and 244 cm−1, the intensity growth of external modes and the peak broadening of near 960 cm−1. The symmetric stretch (ν1 at 1102 cm−1) is displaced to lower frequencies and splits in two bands at 1093 and 1091 cm−1. Among the antisymmetric stretchs (ν3), the peak at 1023 cm−1 disappears in the spectrum of FePO4, while a red shift is observed for bands at 1011 and 978 cm−1; the symmetric stretch (ν1) blue-shifts from 953 to 975 cm−1. Finally, a new peak at 908 cm−1 comes from the red shift of an asymmetry stretch. The largest LFP particles show incomplete delithiation, which could be due to the effect of anti-site defects on the surface preventing Li+ ion motion in the 1D channels as evidenced by TEM experiments [125]. Siddique et al. [126] have combined two in operando techniques, micro-Raman spectroscopy and X‐ray diffraction to study the Li1−xFePO4 phase change in samples with distinct length scales, i.e., particle level scale (≈1 μm) and macroscopic scale (≈several cm) using LiPF6 in EC + DMC solution. In addition to the symmetric stretching mode (ν1) of PO43− anions, the Raman spectrum displays the characteristic bands from the electrolyte: (ⅰ) the peaks at 915 and 895 cm−1 are assigned to C–O stretchs of carbonate groups in non-solvated DMC and EC, respectively, which shift to 932 and 904 cm−1 in solvated solution, (ⅱ) the peak at 714 cm−1 is attributed to the symmetric C = O ring deformation in EC, which moves to 720 cm−1 in solvated EC. In contrast with ex situ measurements [127], the in-situ Raman of the delithiated LFP electrode was able to display conclusive interpretation of the peak profile at 950 cm−1 peak nor new bands in the range 910–1050 cm−1. In particular, from the spectroscopic study of ion-solvent interactions, it was shown that the intensity of the C–O stretch is a linear function of the salt concentration that leads a blue shift in solvated products. The relative intensity of solvation, Ir, is expressed as [128]:

    Ir=Is/(Is+In) (6)

    where Is is the intensity of solvated molecule and In the intensity of bulk non-solvated molecule. Therefore, the variation of the Li+ concentration at the LFP particle surface, CLi, depends on the electrode potential and C-rate. CLi decreases at the beginning of the charge and reached a minimum for the potential of 3.45 V (potential plateau), then CLi increases till the charge at 4.0 V.


    3.6. Li3V2(PO4)3 (LVP)

    Polyphosphate material such as α-Li3V2(PO4)3 (LVP) has a complex three-dimensional network with a monoclinic structure that provides mobility for three Li+ ions, providing a high specific capacity of 197 mAh·g−1. According to the group factor analysis of the LVP structure with P21/n space group (C2h point group), there are 120 Raman active modes at the centre of Brillouin zone, which are classified as:

    ΓαLi3V2(PO4)3=60Ag+60Bg (7)

    Only 19 vibrational modes are observed in the unpolarized Raman spectrum located at ca. 1078, 1059, 1030, 1009,975,652,604,560,505,454,430,375,349,290,255,224,169,134 and 118 cm−1. Burba and Frech [129] studied the vibration spectra of electrochemically delithiated Li3V2(PO4)3 powders (0 ≤ x ≤ 3) and reported the high sensitivity of the (PO4)3− modes with extraction of Li+ ions and change of the vanadium oxidation state. After the first discharge process, Yin et al. [130] found the (PO4)3− band broadening that is assigned to some local disorder in the relithiated lattice. The thermal phase stability of α-Li3V2(PO4)3 was investigated by in situ Raman spectroscopy by varying the irradiation power intensity from 5 to ~600 kW·cm−2 (using a 1.3 μm spot diameter in air). The increase of temperature resulted in phase transitions in the sequence [8]:

    αLi3V2(PO4)3βγαLiVOPO4 (8)

    where the α-LiVOPO4 material is the oxidation phase under air atmosphere.


    3.7. β-AgVO3 (SVO)

    Silver vanadium oxide (SVO) has an important technological application in medical power sources as cathode material of lithium batteries for implantable cardio-pacemakers and defibrillators [131]. The structural changes and of Li inserted β-AgVO3 nanoparticle were studied by in situ Raman spectroscopy using the blue excitation line (λ = 488 nm) combined with atomistic simulation to determine the lithium migration pathways [132]. Upon Li insertion, the reduction sites in the β-AgVO3 lattice can be divided into three steps that correspond to the potential plateau in the discharge profile: (ⅰ) the most favorable reduction on sites V1/V4 occurs at higher potential 3.55–3.05 V, (ⅱ) Ag2/Ag3 are reduced and substituted at medium potential 3.05–2.55 V and (ⅲ) finally, Li+ ion reduces V2/V3/Ag4 and further reduces V1/V4 in the potential range 2.55–1.75 V vs. Li+/Li. In the first step, in which vanadium is reduced from V5+ to V4+, the intensities of Raman peaks at 266,530,772,808,847 and 886 cm−1 are enhanced at 3.35 V and are further weakened at 3.05 V; the stretching vibration at 513 cm−1 of V–O bonds in pyramids exhibits a large blue shift to 530 cm−1, while the asymmetric bending mode of (VO4)3− pyramids at 278 cm−1 has a red-shift to 266 cm−1 related to the distortion of V1/V4 pyramids upon Li introduction in the host lattice. In the second step, the Raman spectrum becomes strongly attenuated due to the loss of resonance and to the increase of electrical conductivity that reduces the optical-skin depth. However, a blue shift of the Raman band at 807 cm−1 (to 811 cm−1) due to the shortened V–O bond is observable. In the third step, no significant modification of the Raman features is detected in the potential range 2.55–2.05 V, while the band at 771 cm−1 shifts to 775 cm−1 between 1.85 and 1.75 V and the peak at 756 cm−1 splits into two components at 750 and 761 cm−1. Moreover, the reduction of vanadium at the V1/V4 site occurring at 1.95 V is characterized by the spit of the band at 730 cm−1 in two well-resolved peaks at 723 and 734 cm−1. These changes of the vibrational modes indicate a significant lattice distortion of β-AgVO3 with Li insertion.


    4. Negative electrodes


    4.1. Carbonaceous materials

    Numerous works prior 1998 have been devoted to the studies of local structure of carbonaceous materials used as negative (anodes) materials, for Li-ion batteries. The term carbonaceous include various forms of carbon materials from highly disordered networks (turbostratic graphite, glassy carbons), to defect-free well crystallized crystals, i.e., highly oriented pyrolytic graphite (HOPG) characterized by its Lc and La parameters indicating dimensions perpendicular and parallel to the basal plane. A relation that is a measure of the structural order in the graphene sheets of carbonaceous materials has been established between the correlation length La (expressed in nm) and the derived spectral parameter, i.e., intensity ratio of Raman carbon-like bands I(D)/I(G) [133]:

    La=4.382I(D)/I(G) (9)

    In this section, we will not expose in detail the results of in situ Raman but will only give some hints about the latest research.


    4.1.1. Graphite intercalation compounds (GICs)

    The layered structure of graphite consists of regular stacked graphene sheets arranged in hexagonal AB stacking sequence of P63/mmc space group (D2h4 point group), of which the factor group analysis gives rise to two Raman active modes for q = 0:

    Γgraphite=2E2g (10)

    The spectrum of graphite exhibits a low-frequency prominent peak E2g1 mode at 42 cm−1 (rigid-layer shear mode) that originates from the in-plane sliding of graphene sheets, and the predominant peak at 1580 cm−1 attributed to the E2g2 mode (in-plane C–C stretching mode) that is usually termed as G band. An additional Raman band at ~1350 cm−1 (D band) appears for polycrystalline graphites and non-graphitic carbons, which is the A1g breathing mode activated by defects and grain boundaries. Overtone bands are also observed at ~2670 cm−1 (2D) and ~3250 cm−1 (2G) [134].

    Since the early work of the Japanese's group [135], the changes in the local structure of graphite intercalation compounds (GIC) have been widely studied by in situ Raman spectroscopy [25,34,44,136,137,138,139,140,141,142,143,144,145,146,147]. At the open circuit potential (ca. 3.0 V vs. Li+/Li), the Raman spectrum exhibits three bands in the region between 1000–3000 cm−1: a weak D-band, a predominant G-band and a weak 2D-band. The staging process of Li intercalation into HOPG and natural graphite was first reported by Inaba et al. [135]. Using in situ Raman spectroscopy, they observed that Li intercalation into graphite proceeds via a series of staged GICs phases, noted by a stage index, n, which represents the number of graphene sheets separating the intercalated layers. It is claimed that Raman spectral changes are correlated with the potential plateaus recorded on the charging reaction as follows: a phase transition from dilute stage-1 to stage-4 followed by the stage-2' (LiC18) → stage-2 (LiC12) and stage-2 (LiC12) → stage-1 (golden LiC6) transitions. This model will be modified in further investigations of Lonza KS‐44 and KS‐6 graphite materials [136]. Panitz et al. [44] presented in situ Raman patterns and imaging of electrode made of TIMREX SFG 44 synthetic graphite with PVdF binder, which reveal a new band at ca. 1850 cm−1 at potential negative to 0.18 V vs. Li+/Li assigned to the complex between Li ions and decomposition of the aprotic solution. This electrolyte/graphite interaction was later reconsidered by Novák et al. [137] in terms of SEI. The so-called dilute stage-1 phase of GICs was deeply studied on KS-44 carbon particles (8–50 μm in diameter) embedded in heat-treated Ni foils [138]. Kostecki and McLarnon [139] investigated the structural degradation of graphite electrodes by portraying the graphite D/G band ratio of the anode via a 50 μm × 75 μm cross-section Raman image.

    During lithiation (discharge process from OCV (ca. 3.0 V) to 0.05 V), modifications of the first- and second-order Raman bands may be classified into four steps: (ⅰ) in the potential range 3.0–0.6 V, the D-band intensity vanishes, (ⅱ) a blue-shift of the G-band from 1580 cm−1 to 1590 cm−1 occurs at ~0.19 V along with a weakening of the 2D(1) band intensity due to the formation of dilute stage-1, (ⅲ) in the potential window 0.19–0.10 V, the G-band splits into the E2g2(ⅰ) (1575 cm−1) and E2g2(b) (1601 cm−1) bands, accompanied by a major red-shift of the 2D(2) band due to the formation of stage-2 and stage-1 and (ⅳ) finally, a gradual loss of all distinct Raman peaks occurs below 0.10 V vs. Li+/Li with the appearance of a weak band located at ca. 1370 cm−1. The formation of highly conductive GIC at low-stage number is at the origin of the loss of all Raman features [148]. Note that the formation of the G-band doublet E2g2(ⅰ) (1578 cm−1) and E2g2(b) (1600 cm−1) is also observed in the case of the electrochemical intercalation of tetraethylammonium (Et4N+) and tetrafluoroborate (BF4) into microcrystalline graphite [149]. The use of propylene carbonate (PC) provokes graphite exfoliation characterized by a new Raman band (E-band) at high wavenumbers of 1609 cm−1 between 0.9 and 0.005 V [144].

    In the reverse reaction (lithium de-intercalation) four stages can be also considered: (ⅰ) a broad G-band at 1592 cm−1 and the weak band at 1370 cm−1 recurs at ca. 0.14 V, (ⅱ) the E2g2(b) band of the stage-2 GIC is formed at ca. 0.15 V from the blue-shift of the G-band from 1592 to 1598 cm−1, (ⅱ) both the G band doublet, E2g2(ⅰ) (1573 cm−1) and E2g2(b) (1601 cm−1) and the 2D(2) band reappear in the potential range 0.17–0.22 V and (ⅳ) the sharp singlet G-band at 1586 cm−1 and the 2D(1) band are reformed coming to the delithiated structure between ca. 0.3–1.5 V. Note that the doublet E2g2(ⅰ), E2g2(b) is associated respectively with carbon vibrations in interior graphite layers and in-bounding graphite layers adjacent to the intercalate planes. This splitting is attributed to changes in symmetry at the boundary layer and to the change in electronic states upon Li+ ion intercalation at stage n > 2. When intercalated graphite appears as stage-1 and -2, the E2g2(ⅰ) band vanishes because of the absence of graphite interior layer. By taking into consideration the intensity ratio of the E2g doublet (R), Solin has proposed a quantitative measure of the intercalation stage index, n, by the following equation [150]:

    R=IiIb=σiσbn22(n>2) (11)

    where Ii and Ib are the intensities of the interior E2g2(i) and bounding E2g2(b) layer modes, respectively, and σib is a stage-independent ratio of the Raman scattering cross-section from the interior and bounding layers, equal to 1 for lithium intercalation.

    Typical in situ Raman spectro-electrochemical measurements of graphite electrode is presented in Figure 17, where spectral features were taken on different points that show the inhomogeneity of the particle distribution in the electrode [143]. Detailed real-time Raman spectra from a single KS-44 graphite microflake electrode determined that the transition between dilute stage-1 and stage-4 of lithiated GIC occurs in the potential range 0.174–0.215 V vs Li+/Li [141]. Graphite surface disorder was detected by the variation in electrode potential of the appearance of the G-band doublet associated to the formation of stage-4 GIC [143]. During lithium intercalation in hydrogen-rich amorphous carbon, lithium-carbon interaction leads to an increase of IG/ID and formation of Li–C bands at ~480 and ~700 cm−1 [151]. Raman imaging of synthetic graphite LF-18D particles with an average particle diameter of 22 μm (Chuetsu Graphite Works Co.) showed that the Li does not intercalate in a homogeneous manner; it varies with the article morphology and local conditions [142]. In situ Raman experiments evidenced the intercalation effect of both large cation (EMI+) and anion (TFSI) in KS44 microcrystalline graphite and https://www.aimspress.com/aimspress-data/aimsmates/2018/4/PICACTIF activated carbon. Intercalation is detrimental for graphite with an increased ID/IG ratio, while possible intercalation of activated carbon upon cycling is evidenced by the blue-shift of the D-band [146]. The surface region of graphitic negative electrode (microcrystalline graphite) was examined by in situ real-time Raman measurements under potential control during Li+-ion insertion and extraction [33]. A special attention was paid to the double resonance 2D-band that shifts from 2681 to 2611 cm−1. The change of the electronic structure of the intercalated compound and the C–C bonding of stage 3 and 4 was evidenced by the band shape transformation into a single Lorentzian in the potential range 0.24–0.15 V vs. Li+/Li. It was concluded that the Daumas–Hérold model can be applied, where the graphene sheets are flexible and deform around Li intercalated domains.

    Figure 17. Typical in situ Raman spectro-electrochemical measurements of graphite electrode. (a) Potential profile during the lithiation of graphite showing the staged phase. (b and c) Spectral features as a function of the electrode potential taken on different points that show the inhomogeneity of the particle distribution in the electrode. Reprinted by permission from Ref. [143].

    In situ Raman spectro-electrochemical measurements were also carried out to study the degradation of edge-plane graphite negative-electrodes [152], the SEI formation [153], the influence of the salt in the electrolyte solution on the structural degradation of graphite [154], the surface evolution of a single graphite particle with ethylene carbonate/dimethyl carbonate [155], the behavior of graphite electrodes in electrolyte solution containing fluorinated phosphoric esters [25], the decrease in the surface crystallinity of graphite negative-electrodes at high potentials in LiPF6-based electrolyte solution [156], the correlations of concentration changes of the electrolyte salt with the surface state of a HOPG graphite electrode, the Li+ ion intercalation and deintercalation behaviors of graphitized carbon nanospheres [157] and the structure of heat-treated graphene nanoflake-based anodes [158].


    4.1.2. Li insertion in other carbonaceous materials

    By contrast to graphite, carbonaceous materials with non-graphitic structure exhibit poor galvanostatic responses with slopping potential profiles that originate from disorder. Coke, carbon black and glassy carbon (GC) deliver low-specific capacity, i.e., 223 mAh·g−1 for GC against 372 mAh·g−1 for graphite. The spectral modifications of the Raman E2g2 band of mesocarbon microbeads (MCMBs) annealed at 2800 ℃ presents a Li insertion mechanism similar to that of graphite with formation of staged phases [159]. Totir and Scherson [160] investigated the electrochemical lithium intercalation of embedded KS‐44 graphite and MCF28 carbon microfibers and reported spectral patterns in good agreement with similar materials. The graphene oxide (GO) and reduced graphene oxide (rGO) obtained by scanning the cell potential from 0 to −1 V vs. SCE in an aqueous electrolyte were characterized by in situ Raman spectro-electrochemical study, which showed a clear shift of the G-band from 1610 to 1585 cm−1 for GO and rGO, respectively [161]. In situ Raman and nuclear magnetic resonance measurements were performed to study the trapped Li ions in the SEI of rGO [162]. It is shown that the graphitic G-band weakens and vanishes fast at potential > 0.3 V without staged GIC and suggested that the monolithic defective sites are randomly accommodated by Li+ ions, which cover the rGO surface and cause the G-band disappearance.


    4.2. Lithium titanate Li4Ti5O12 (LTO)

    Li4Ti5O12 (LTO) is an anode material being a zero-strain network with high thermal stability. It crystallizes in the cubic structure with Fd3m space group (Oh7 point group) and can be expressed as Li[Li1/3Ti5/3]O4 in the spinel notation; thus, the Raman-allowed phonon peaks are represented by species in Eq 3, while factor group analysis predicts 4 Raman-active modes for the lithiated phase Li2[Li1/3Ti5/3]O4 at q = 0:

    Γ=A1g+Eg+2Tg (12)

    The Raman spectrum of the pristine material exhibits three intense Raman bands centered at 671,430 and 232 cm−1 along with two weak Raman peaks at 347 and 271 cm−1. The vibration at 671 cm−1 (A1g mode) is ascribed to the symmetric Ti–O stretch of TiO6 octahedra, while the peak at 430 cm−1 (Eg mode) originate from the asymmetric Li–O stretch of LiO4 tetrahedra and the 232 cm−1 band (Tg mode) from δ(Ti–O). The Raman peak at 232 cm−1 and other weak bands are assigned as T2g mode [163]. During discharge, the lithiated Li1+x[Li1/3Ti5/3]O4 electrode can accommodate x ≈ 0.94 Li with a CV curve showing a very wide potential plateau around 1.55 V vs. Li+/Li. Insertion of Li+ ions in octahedral 16c empty sites of the Li1+x[Li1/3Ti5/3]O4 framework and the reduction of Ti ions in TiO6 octahedra provoke a gradual blue shift of Raman bands at 430 and 232 cm−1, while the high-wavenumber peak at 671 cm−1 is maintained [164]. In situ Raman studies show that, at the end of the discharge, the spinel transforms in rock-salt lattice and, as the electronic conductivity increases considerably due to the presence of Ti4+ and Ti3+ ions, the Raman line intensities gradually weaken with x and finally disappear at x ≈ 0.94 due to the decrease of the optical-skin depth [165].


    4.3. Titanium dioxide TiO2

    TiO2 has been exhaustively investigated due to its attractive properties; it is low cost, thermally stable, environmentally safe, etc., and it is considered as a promising candidate for use in Li-ion batteries. TiO2 exists in different polymorphic forms: anatase, rutile, brookite, TiO2-B (bronze), TiO2-R (ramsdellite), TiO2-H (hollandite), etc. [166]. It appears that the potential profile of the Li//TiO2 couple is strongly influenced by the structure; as an example, the working voltage of Li//anatase TiO2 cell is higher than that of Li//rutile TiO2 cell. Anatase TiO2 (tetragonal structure, space group I41/amd, D4h19 point group) can accommodate reversibly x ≈ 0.5 Li that converts Li0.5TiO2 to orthorhombic phase (space group Imma, D2h28 point group) [167]. Only nanocrystalline TiO2 particles (<7 nm in size) are able to host x ≈ 1 Li [168]. The lattice dynamics of pristine anatase TiO2 has been subject to numerous works (see [29] and Refs. herein). Its first-order Raman analysis gives rise to 6 active vibrations at q = 0:

    ΓTiO2=A1g+3Eg+2B1g (13)

    which have been experimentally observed at 635,515,395,195 and 142 cm−1 [169]. The symmetry change from D4h14 to D2h28 induced by elongated equatorial Ti–O bonds in Li0.5TiO2 produces 9 Raman active modes represented by:

    ΓLi0.5TiO2=3Ag+3B2g+3B3g (14)

    where B2g and B3g modes come from the splitting of the Eg mode. Non-stoichiometry and phonon confinement in nanoparticles (>10 nm in size) produce a blue shift and a broadening of the lowest-frequency Eg vibration. Several studies reported in situ Raman spectro-electrochemistry of Li-inserted anatase TiO2 [170,171,172,173,174,175,176,177]. Dinh et al. [170] combined in situ Raman and in situ transmittance spectra of electrochromic TiO2 anatase thin films and demonstrated that the colored film exhibit five characteristic bands at 629,531,316,224 and 176 cm−1. Hardwick et al. [173] studied three nanosized TiO2 anatase powders (80, 15 and 8 nm in size) and, using in situ Raman microscopy, followed the increase in electrical conductivity as lithiation proceeds, with a marked band intensity decrease for x > 0.3 in LixTiO2. In situ Raman spectro-electrochemistry was a powerful tool to detect the contamination of anatase TiO2 in the TiO2(B) and TiO2(rutile) polymorphs during Li+ insertion/deinsertion [176]. In situ Raman spectro-electrochemistry showed that structural modifications during Li insertion in anatase TiO2 nanoparticles can be understood by a mechanism of Li-poor tetragonal → orthorhombic → Li-rich tetragonal double phase transformation. For Li0.09TiO2 (at 1.75 V electrode potential), a blue-shift of the band at 142 cm−1 is observed (Δν = 5 cm−1) followed by the two-phase process characterized by the occurrence of two bands at 165 and 177 cm−1 that corresponds to the potential plateau at 1.4 V. The additional Raman peaks associated with the end-composition Li0.5TiO2 are recorded at 625,555,528,355,339, and 315 cm−1 [175].


    4.4. Silicon

    Silicon is an attractive anode material because of its low cost, its large specific capacity, i.e., 4200 mAh·g−1 when lithiated to Li4.4Si (or Li22Si5), and its relatively low redox potential, i.e., 0.3–0.4 V above Li+/Li [178]. However, the electrochemical performance of Li–Si alloys degrades due to the large volume expansion, i.e., 420% from Si to Li4.4Si. During discharge the Li–Si alloy is formed and subsequently a transition from crystalline (c-Si) to amorphous (a-Si) silicon occurs. In this context, few researchers have studied the phase transformation using in situ Raman microscopy [179,180,181,182,183]. One of the earlier in situ Raman investigation of silicon anode was conducted by Holzapfel et al. [184]. Unlithiated Si material exhibits the typical first-order (q = 0) Raman peak at 520 cm−1 (optical phonon) of the crystalline phase that significantly decreases at potential below 0.88 V upon Li alloying. For further lithiation, the Si peak vanishes totally and a broad band grows at ca. 480 cm−1 when the discharge reaches 0.09 V, which could indicate the growth of an amorphous LixSi phase [179]. Nanda et al. [180] characterized the surface of the Si/C composite anode at a microscopic level and monitored the alloying/dealloying behavior of Si/C with Li, using in situ Raman microscopy study. As shown in the Raman spectra of silicon as a function of the first discharge between 2.0 and 0.45 V (Figure 18), the Raman peak of Si peak does not change (discharged at a current density of 90 μA·cm−2), while at 0.2 V the signal substantially decreases and disappears below 0.1 V due to the conversion of the crystalline Si into an alloyed Li–Si phase.

    Figure 18. In situ Raman spectra of a Li//Si electrode discharged at a current density of 90 μA·cm−2. Reprinted by permission from Ref. [180].

    Long et al. [185] examined the effects of p-type (boron) and n-type (phosphorus) dopants on the lithiation of crystalline Si. The transition to amorphous Si associated with the Li insertion was monitored by in situ Raman spectroscopy. A plot of the height of phonon peak shows that the n-type Si exhibits a phonon decay at potential of 0.09 V, against 0.68 V for the p-type Si. This effect is explained in terms of energy states available to the electron associated with the Li+ ion. The lithiation of silicon decorated by plasmonic metals, i.e., Ag and Au, was characterized by in situ Raman microscopy and Raman mapping on a section of 20 × 30 μm2 using a laser wavelength of 532 nm [186]. The structural evolution, the internal strain and stress in the lithiated and delithiated Si nanoparticles were monitored by both in situ X-ray diffraction and in operando Raman spectroscopy [182,187]. Upon the formation of Li–Si alloys, the drop of the phonon intensity results first from the amorphization of the c-Si framework and secondly from the decrease of the optical-skin depth. There are no changes of peak position during the first discharge, while a blue shift occurs during delithiation. In contrast, the upshift that originates from compressive stress is followed by the downshift due to tensile stress during the second cycle and so forth. From the experimental Raman shifts, one can estimate the magnitude of the stress using the relation [182]:

    σ=230(ωsω0) (15)

    where ω0 and ωs are the wavenumbers for the relaxed and of the stressed particles, respectively. Yang et al. [188] experienced the lithiation behavior of Cu-nanowire/Si nano-particle composite anodes using in situ micro-Raman mapping, showing that the Li–Si reaction does not occur in a homogeneous manner due to disconnected active particles. Recently, Sakaguchi et al. strenuously determine silicon anode properties [189,190,191]. The deterioration mechanism of Si anode was analyzed using Raman mapping of Li–Si alloying and dealloying reactions [189]. Evolution of the Raman shift of the 520 cm−1 line of phosphorous-doped Si revealed that doping monitored the Li–Si alloying reaction and the formation of the c-Li15Si4 phase [190].


    4.5. Other anode materials


    4.5.1. Antimony

    The active researches of high-performance anode materials have motivated the consideration of metal-containing compounds in the form of intermetallics such as Sb, InSb, Cu2Sb, AlSb, etc. that make alloys upon lithiation [192]. Antimony is capable of 3Li uptake forming an alloy with a theoretical specific capacity of 660 mAh·g−1 based on the Li3Sb reaction [193]. The lithium alloying of carbon-coated antimony microparticles synthesized by a one-pot sol-gel auto-combustion route was analyzed by in situ Raman spectroscopy using a red laser-line (λ = 632.8 nm) as excitation source [194].

    Using in situ Raman spectro-electrochemistry in the potential between 1.5 and 0.02 V, a comparison with commercially available Sb nanoparticles of same size (ca. 5–50 μm) shows a superiority of sol-gel prepared Sb (Figure 19). The Raman spectrum exhibits the signal due to D- and G-band of the disordered carbon coating and the phonons of antimony at 150 cm−1 (A1g mode) and 112 cm−1 (Eg mode), which is sensitive to the confinement and strain of particles. During lithiation, the Raman spectrum of Sb is maintained until ~0.9 V vs. Li+/Li and then the intensity of phonon vibrations decreases until smearing out at 0.02 V. Some degree of long range ordering is recovered upon delithiation to 1.5 V with the reemergence of the A1g peak at 152 cm−1.

    Figure 19. (a) Discharge–charge profile of the LixSb anode. (b) In situ Raman spectra collected in the potential between OCV and 0.02 V vs. Li+/Li. Reprinted by permission from Ref. [194].

    4.5.2. ZnM2O4 (M = Mn, Fe)

    Among the ternary Zn-based oxides, the spinel ZnMn2O4 material reacts with lithium with both conversion and alloying process, which results in a specific capacity 1024 mAh·g−1. In situ Raman spectra of mesoporous ZnMn2O4 microspheres were collected in the potential range 2.0–0.01 V vs. Li+/Li (Figure 20) [195]. The structural changes were monitored by following the evolution of the three Raman bands located at 681 cm−1 (A1g symmetry involving motion of oxygen in AO4 tetrahedra), 386 and 324 cm−1 (involving the vibration of BO6 octahedra) according to the usual notations for compounds of generic form AB2O4. When the Mn3+ ions are reduced to Mn2+ during the lithiation process at 1.21 V, the A1g band begins to wane and disappears at 0.16 V, while the strength of Raman bands at 386 and 324 cm−1 decreases. At this potential Zn2+ and Mn2+ are reduced to Zn0 and Mn0. During the charge process (delithiation) at 1.25 V, a band grows at 694 cm−1 which is characteristic of the asymmetric Mn–O stretch for R-Mn2O3 associated with the oxidation of Mn0 to Mn3+. Above 1.52 V, a band located at 407 cm−1 assigned to the E2 mode of ZnO is the fingerprint of oxidized Zn0. Therefore, the two weak Raman bands at 324 and 386 cm−1 for ZnMn2O4 reappear at the end of the charge process.

    Figure 20. In situ Raman spectra showing the lithiation and delithiation of ZnMn2O4 at representative potentials. Reprinted by permission from Ref. [195].

    Cabo-Fernandez et al. [196] investigated the delithiation process of carbon-coated ZnFe2O4 (ZFO) nanoparticles. Like ZnMn2O4, ZFO is a low-cost anode material with a theoretical specific capacity ~1000 mAh·g−1 that reacts with lithium via a conversion/alloying process. The first-order Rama spectrum displays 5 bands located at 647 cm−1 (A1g), 467 cm−1 (T2g), 340 cm−1 (T2g), 246 cm−1 (Eg) and 221 cm−1 (T2g). The conversion reaction that occurs at ~0.8 V leads the formation of Zn0 particles, which enhance the Raman signal of the SEI components. In the electrode potential between 0.80 and 0.69 V, the Raman spectrum is dominated by the reduction products of electrolyte carbonate solvents which form the SEI.


    5. Solid electrolyte interface (SEI)

    The solid electrolyte interface is a chemical layer formed on the anode side, i.e., lithium metal or graphite, that is indispensable for the electrochemical stability of Li-ion batteries using organic (aprotic) electrolyte. A SEI is the result of solvation of Li ions, i.e., electrolyte reduction because of the sufficiently low potential of graphite, resulting in the formation of a film deposited on the electrode surface in the first cycle of charge/discharge. After the second cycle, the SEI suppresses the electrolyte decomposition acting as an energetic barrier between the Fermi level of the graphite and the high occupied molecular orbital (HOMO) of the electrolyte. Analysis of the SEI using in situ Raman spectroscopy has been reported by few research teams [197,198,199,200,201] only, because the SEI layer is usually very thin so that the Raman signal difficult to be detected. The composition and formation of the SEI were analyzed using copper electrodes after lithium plating. From Raman spectra Schmitz et al. [197] have detected semicarbonates such as Li2CO3 and Li2C2 as SEI components. In situ Raman micro-spectroscopy and Raman mapping reveal that: (ⅰ) Li2C2 is not formed at a potential > 0 V vs. Li+/Li, (ⅱ) Li2C2 is located on the plated lithium and (ⅲ) Li2CO3 is homogeneously distributed over the copper sheet. The presence of Li2C2 on the metallic Li surface has been considered as a product of laser degradation rather than an actual SEI species [202,203]. The SEI formation onto SiO2-coated Au nanoparticles was studied using in situ surfaced-enhanced Raman spectroscopy (SERS) in lithium-ion battery [201]. The effect of the additive vinylene carbonate (VC) in the ethylene carbonate (EC) electrolyte was investigated by monitoring the band intensity of silicon, which exhibits different amorphization rates between bulk and surface. Park et al. [204] investigated the mechanism of the electrolyte–electrode interface reaction using in situ Raman spectroscopy during the second charging process. Due to the low permeability of the SEI, there is a variation of ion concentration during charging and discharging a Li-ion battery [205]. This change in concentration nearby the separator has been investigated by in situ Raman spectroscopy using ultrafine microfiber probes [206,207]. Recently, Yamanaka et al. [22,207] performed such measurements with a graphite//LiFePO4 cell using an electrolyte solution of 1 mol·L−1 LiClO4 dissolved in ethylene carbonate (EC) and diethyl carbonate (DEC) solution. Intensities of three Raman peaks were analyzed: at 718 and 730 cm−1 (symmetric ring deformation mode of EC and solvated EC with Li+ and 935 cm−1 (symmetric mode of the anion ClO4). The properties of the SEI were modified by adding film-forming additives to the EC:DEC solution. Based on in situ Raman spectro-electrochemistry, the concerted effect of the SEI and the adjacent separator film generates an effective barrier for Li ions [208]. The same research's group [209] performed a chronoamperometric pretreatment of HOPG anode at 0.9 V vs. lithium applied for 20 h that decrease the resistance charge transfer and capacitance at the surface of graphite. The main result of this pretreatment consists in the suppression of the concentration changes.


    6. Conclusions

    In this review, we have shown that, among the various in situ methods, Raman spectro-electrochemistry is a highly effective technique in intensive field of researches covering all aspects of battery operation. Regarding the structural response of electrode materials, the contribution of both in-situ Raman spectro-electrochemistry and Raman mapping were determining for the issue of real time charge–discharge cycling because they are powerful tool for studying the short-range order (local environment) in lattices. In situ analyses are carried out to track several effects occurring during charging and discharging process in lithium batteries such as phase stability, structural modifications of the electrode materials during insertion/deinsertion reaction, interface evolution, compatibility between materials, kinetics of Li+ ions diffusing in the electrodes et electrolytes, electrode degradation and formation of the SEI layer.

    As other in operando analytical methods, in situ Raman spectroscopy is used not only to improve the cycle stability of entire batteries, but to provide experimental evidence of new insights of Li+-ion insertion/deinsertion mechanism. However, due to the short optical penetration depth, Raman spectroscopy is essentially a surface analysis technique and the spectral response cannot be extended to the bulk of a sample that makes some difference with results obtained from in situ X-ray diffraction when the rate of analysis is not enough slow.


    Conflict of interest

    The authors declare that there is no conflict of interest regarding the publication of this manuscript.


    [1] Julien CM, Mauger A, Vijh A, et al. (2016) Principles of Intercalation, In: Julien CM, Mauger A, Vijh A, et al., Lithium Batteries: Science and Technology, Cham, Switzerland: Springer, 69–91.
    [2] Novák P, Panitz JC, Joho F, et al. (2000) Advanced in situ methods for the characterization of practical electrodes in lithium-ion batteries. J Power Sources 90: 52–58. doi: 10.1016/S0378-7753(00)00447-X
    [3] Shao M (2014) In situ microscopic studies on the structural and chemical behaviors of lithium-ion battery materials. J Power Sources 270: 475–486. doi: 10.1016/j.jpowsour.2014.07.123
    [4] Harks PPRML, Mulder FM, Notten PHL (2015) In situ methods for Li-ion battery research: A review of recent developments. J Power Sources 288: 92–105. doi: 10.1016/j.jpowsour.2015.04.084
    [5] Weatherup RS, Bayer BC, Blume R, et al. (2011) In situ characterization of alloy catalysts for low-temperature graphene growth. Nano Lett 11: 4154–4160. doi: 10.1021/nl202036y
    [6] Yang Y, Liu X, Dai Z, et al. (2017) In situ electrochemistry of rechargeable battery materials: Status report and perspectives. Adv Mater 29: 1606922. doi: 10.1002/adma.201606922
    [7] Shu J, Ma R, Shao L, et al. (2014) In-situ X-ray diffraction study on the structural evolutions of LiNi0.5Co0.3Mn0.2O2 in different working potential windows. J Power Sources 245: 7–18.
    [8] Membreño N, Xiao P, Park KS, et al. (2013) In situ Raman study of phase stability of α-Li3V2(PO4)3 upon thermal and laser heating. J Phys Chem C 117: 11994–12002. doi: 10.1021/jp403282a
    [9] Kong F, Kostecki R, Nadeau G, et al. (2001) In situ studies of SEI formation. J Power Sources 97–98: 58–66. doi: 10.1016/S0378-7753(01)00588-2
    [10] Mehdi BL, Qian J, Nasybulin E, et al. (2015) Observation and quantification of nanoscale processes in lithium batteries by operando electrochemical (S)TEM. Nano Lett 15: 2168–2173. doi: 10.1021/acs.nanolett.5b00175
    [11] Kerlau M, Marcinek M, Srinivasan V, et al. (2007) Studies of local degradation phenomena in composite cathodes for lithium-ion batteries. Electrochim Acta 52: 5422–5429. doi: 10.1016/j.electacta.2007.02.085
    [12] Hausbrand R, Cherkashinin G, Ehrenberg H, et al. (2015) Fundamental degradation mechanisms of layered oxide Li-ion battery cathode materials: Methodology, insights and novel approaches. Mater Sci Eng B-Adv 192: 3–25. doi: 10.1016/j.mseb.2014.11.014
    [13] Julien C (2000) Local environment in 4-volt cathode materials for Li-ion batteries, In: Julien C, Stoynov Z, Materials for Lithium-ion Batteries, NATO Science Series (Series 3. High Technology), Dordrecht: Springer, 309–326.
    [14] Itoh T, Sato H, Nishina T, et al. (1997) In situ Raman spectroscopic study of LixCoO2 electrodes in propylene carbonate solvent systems. J Power Sources 68: 333–337. doi: 10.1016/S0378-7753(97)02539-1
    [15] Song SW, Han KS, Fujita H, et al. (2001) In situ visible Raman spectroscopic study of phase change in LiCoO2 film by laser irradiation. Chem Phys Lett 344: 299–304. doi: 10.1016/S0009-2614(01)00677-7
    [16] Julien C (2000) Local cationic environment in lithium nickel–cobalt oxides used as cathode materials for lithium batteries. Solid State Ionics 136–137: 887–896. doi: 10.1016/S0167-2738(00)00503-8
    [17] Julien CM, Gendron F, Amdouni N, et al. (2006) Lattice vibrations of materials for lithium rechargeable batteries. VI: Ordered spinels. Mater Sci Eng B-Adv 130: 41–48. doi: 10.1016/j.mseb.2006.02.003
    [18] Ramana CV, Smith RJ, Hussain OM, et al. (2004) Growth and surface characterization of V2O5 thin films made by pulsed-laser deposition. J Vac Sci Technol A 22: 2453–2458. doi: 10.1116/1.1809123
    [19] Zhang X, Frech R (1998) In situ Raman spectroscopy of LixV2O5 in a lithium rechargeable battery. J Electrochem Soc 145: 847–851. doi: 10.1149/1.1838356
    [20] Inaba M, Iriyama Y, Ogumi Z, et al. (1997) Raman study of layered rock‐salt LiCoO2 and its electrochemical lithium deintercalation. J Raman Spectrosc 28: 613–617. doi: 10.1002/(SICI)1097-4555(199708)28:8<613::AID-JRS138>3.0.CO;2-T
    [21] Battisti D, Nazri GA, Klassen B, et al. (1993) Vibrational studies of lithium perchlorate in propylene carbonate solutions. J Phys Chem 97: 5826–5830. doi: 10.1021/j100124a007
    [22] Yamanaka T, Nakagawa H, Tsubouchi S, et al. (2017) In situ Raman spectroscopic studies on concentration of electrolyte salt in lithium-ion batteries by using ultrafine multifiber probes. ChemSusChem 10: 855–861. doi: 10.1002/cssc.201601473
    [23] Yang Y, Liu X, Dai Z, et al. (2017) In situ electrochemistry of rechargeable battery materials: Status report and perspectives. Adv Mater 29: 1606922. doi: 10.1002/adma.201606922
    [24] Lei J, McLarnon F, Kostecki R (2005) In situ Raman microscopy of individual LiNi0.8Co0.15Al0.05O2 particles in a Li-ion battery composite cathode. J Phys Chem B 109: 952–957.
    [25] Nakagawa H, Dom Yi, Doi T, et al. (2014) In situ Raman study of graphite negative-electrodes in electrolyte solution containing fluorinated phosphoric esters. J Electrochem Soc 161: A480–A485. doi: 10.1149/2.007404jes
    [26] Zhang X, Cheng F, Zhang K, et al. (2012) Facile polymer-assisted synthesis of LiNi0.5Mn1.5O4 with a hierarchical micro–nano structure and high rate capability. RSC Adv 2: 5669–5675. doi: 10.1039/C2RA20669B
    [27] Amaraj SF, Aurbach D (2011) The use of in situ techniques in R&D of Li and Mg rechargeable batteries. J Solid State Electr 15: 877–890. doi: 10.1007/s10008-011-1324-9
    [28] Stancovski V, Badilescu S (2014) In situ Raman spectroscopic–electrochemical studies of lithium-ion battery materials: a historical overview. J Appl Electrochem 44: 23–43. doi: 10.1007/s10800-013-0628-0
    [29] Baddour-Hadjean R, Pereira-Ramos JP (2010) Raman microspectrometry applied to the study of electrode materials for lithium batteries. Chem Rev 110: 1278–1319. doi: 10.1021/cr800344k
    [30] Burba CM, Frech R (2006) Modified coin cells for in situ Raman spectro-electrochemical measurements of LixV2O5 for lithium rechargeable batteries. Appl Spectrosc 60: 490–493. doi: 10.1366/000370206777412167
    [31] Gross T, Giebeler L, Hess C (2013) Novel in situ cell for Raman diagnostics of lithium-ion batteries. Rev Sci Instrum 84: 073109. doi: 10.1063/1.4813263
    [32] Gross T, Hess C (2014) Raman diagnostics of LiCoO2 electrodes for lithium-ion batteries. J Power Sources 256: 220–225. doi: 10.1016/j.jpowsour.2014.01.084
    [33] Sole C, Drewett NE, Hardwick LJ (2014) In situ Raman study of lithium-ion intercalation into microcrystalline graphite. Faraday Discuss 172: 223–237. doi: 10.1039/C4FD00079J
    [34] Novák P, Goers D, Hardwick L, et al. (2005) Advanced in situ characterization methods applied to carbonaceous materials. J Power Sources 146: 15–20. doi: 10.1016/j.jpowsour.2005.03.129
    [35] Fang S, Yan M, Hamers RJ (2017) Cell design and image analysis for in situ Raman mapping of inhomogeneous state-of-charge profiles in lithium-ion batteries. J Power Sources 352: 18–25. doi: 10.1016/j.jpowsour.2017.03.055
    [36] Ghanty C, Markovsky B, Erickson EM, et al. (2015) Li+-ion extraction/insertion of Ni-rich Li1+x(NiyCozMnz)wO2 (0.005 < x < 0.03; y:z = 8:1, w ≈ 1) electrodes: in situ XRD and Raman spectroscopy study. ChemElectroChem 2: 1479–1486. doi: 10.1002/celc.201500160
    [37] Huang JX, Li B, Liu B, et al. (2016) Structural evolution of NM (Ni and Mn) lithium-rich layered material revealed by in-situ electrochemical Raman spectroscopic study. J Power Sources 310: 85–90. doi: 10.1016/j.jpowsour.2016.01.065
    [38] Dash WC, Newman R (1955) Intrinsic optical absorption in single-crystal germanium and silicon at 77 K and 300 K. Phys Rev 99: 1151–1155. doi: 10.1103/PhysRev.99.1151
    [39] Brodsky MH, Title RS, Weiser K, et al. (1970) Structural, optical, and electrical properties of amorphous silicon films. Phys Rev B 1: 2632–2641. doi: 10.1103/PhysRevB.1.2632
    [40] Pollak E, Salitra G, Baranchugov V, et al. (2007) In situ conductivity, impedance spectroscopy, and ex situ Raman spectra of amorphous silicon during the insertion/extraction of lithium. J Phys Chem C 111: 11437–11444. doi: 10.1021/jp0729563
    [41] Tuinstra F, Koenig JL (1970) Raman spectrum of graphite. J Chem Phys 53: 1126–1130. doi: 10.1063/1.1674108
    [42] Kostecki R, Schnyder B, Alliata D, et al. (2001) Surface studies of carbon films from pyrolyzed photoresist. Thin Solid Films 396: 36–43. doi: 10.1016/S0040-6090(01)01185-3
    [43] Julien CM, Zaghib K, Mauger A, et al. (2006) Characterization of the carbon coating onto LiFePO4 particles used in lithium batteries. J Appl Phys 100: 063511. doi: 10.1063/1.2337556
    [44] Panitz JC, Joho F, Novák P (1999) In situ characterization of a graphite electrode in a secondary lithium-ion battery using Raman microscopy. Appl Spectrosc 53: 1188–1199. doi: 10.1366/0003702991945650
    [45] Panitz JC, Novák P, Haas O (2001) Raman microscopy applied to rechargeable lithium-ion cells -Steps towards in situ Raman imaging with increased optical efficiency. Appl Spectrosc 55: 1131–1137. doi: 10.1366/0003702011953379
    [46] Panitz JC, Novák P (2001) Raman spectroscopy as a quality control tool for electrodes of lithium-ion batteries. J Power Sources 97–98: 174–180. doi: 10.1016/S0378-7753(01)00679-6
    [47] Kostecki R, Lei J, McLarnon F, et al. (2006) Diagnostic evaluation of detrimental phenomena in high-power lithium-ion batteries. J Electrochem Soc 153: A669–A672. doi: 10.1149/1.2170551
    [48] Forster JD, Harris SJ, Urban JJ (2014) Mapping Li+ concentration and transport via in situ confocal Raman microscopy. J Phys Chem Lett 5: 2007–2011. doi: 10.1021/jz500608e
    [49] Nishi T, Nakai H, Kita A (2013) Visualization of the state-of-charge distribution in a LiCoO2 cathode by in situ Raman imaging. J Electrochem Soc 160: A1785–A1788. doi: 10.1149/2.061310jes
    [50] Nanda J, Remillard J, O'Neill A, et al. (2011) Local state-of-charge mapping of lithium-ion battery electrodes. Adv Funct Mater 21: 3282–3290. doi: 10.1002/adfm.201100157
    [51] Slautin B, Alikin D, Rosato D, et al. (2018) Local study of lithiation and degradation paths in LiMn2O4 battery cathodes: confocal Raman microscopy approach. Batteries 4: 21. doi: 10.3390/batteries4020021
    [52] Li J, Shunmugasundaram R, Doig R, et al. (2016) In situ X-ray diffraction study of layered Li–Ni–Mn–Co oxides: Effect of particle size and structural stability of core–shell materials. Chem Mater 28: 162–171. doi: 10.1021/acs.chemmater.5b03500
    [53] Graetz J, Gabrisch H, Julien CM, et al. (2003) Raman evidence of spinel formation in delithiated cobalt oxide. Electrochemical society Proceedings Volume 2003-28; Proceedings-electrochemical Society PV, Symposium, Lithium and lithium-ion battery, 28: 95–100.
    [54] Yu L, Liu H, Wang Y, et al. (2013) Preferential adsorption of solvents on the cathode surface of lithium ion batteries. Angew Chem Int Edit 52: 5753–5756. doi: 10.1002/anie.201209976
    [55] Kuwata N, Ise K, Matsuda Y, et al. (2012) Detection of degradation in LiCoO2 thin films by in situ micro Raman spectroscopy, In: Chowdari BVR, Kawamura J, Mizusaki J, et al. (Eds.), Solid State Ionics: Ionics for Sustainable World, Proceedings of the 13th Asian Conference, Singapore: World Scientific, 138–143.
    [56] Gross T, Hess C (2014) Spatially-resolved in situ Raman analysis of LiCoO2 electrodes. ECS Trans 61: 1–9. doi: 10.1149/06112.0001ecst
    [57] Fukumitsu H, Omori M, Terada K, et al. (2015) Development of in situ cross-sectional Raman imaging of LiCoO2 cathode for Li-ion battery. Electrochemistry 83: 993–996. doi: 10.5796/electrochemistry.83.993
    [58] Heber M, Schilling C, Gross T, et al. (2015) In situ Raman and UV-Vis spectroscopic analysis of lithium-ion batteries. MRS Proceedings 1773: 33–40. doi: 10.1557/opl.2015.634
    [59] Otoyama M, Ito Y, Hayashi A, et al. (2016) Raman imaging for LiCoO2 composite positive electrodes in all-solid-state lithium batteries using Li2S–P2S5 solid electrolytes. J Power Sources 302: 419–425. doi: 10.1016/j.jpowsour.2015.10.040
    [60] Porthault H, Baddour-Hadjean R, Le Cras F, et al. (2012) Raman study of the spinel-to-layered phase transformation in sol–gel LiCoO2 cathode powders as a function of the post-annealing temperature. Vib Spectrosc 62: 152–158. doi: 10.1016/j.vibspec.2012.05.004
    [61] Park Y, Kim NH, Kim JY, et al. (2010) Surface characterization of the high voltage LiCoO2/Li cell by X-ray photoelectron spectroscopy and 2D correlation analysis. Vib Spectrosc 53: 60–63. doi: 10.1016/j.vibspec.2010.01.004
    [62] Zhang X, Mauger A, Lu Q, et al. (2010) Synthesis and characterization of LiNi1/3Mn1/3Co1/3O2 by wet-chemical method. Electrochim Acta 55: 6440–6449. doi: 10.1016/j.electacta.2010.06.040
    [63] Ben-Kamel K, Amdouni N, Mauger A, et al. (2012) Study of the local structure of LiNi0.33+dMn0.33+dCo0.33−2dO2 (0.025 ≤ d ≤ 0.075) oxides. J Alloy Compd 528: 91–98. doi: 10.1016/j.jallcom.2012.03.018
    [64] Hayden C, Talin AA (2014) Raman spectroscopic analysis of LiNi0.5Co0.2Mn0.3O2 cathodes stressed to high voltage.
    [65] Otoyama M, Ito Y, Hayashi A, et al. (2016) Raman spectroscopy for LiNi1/3Mn1/3Co1/3O2 composite positive electrodes in all-solid-state lithium batteries. Electrochemistry 84: 812–816. doi: 10.5796/electrochemistry.84.812
    [66] Lanz P, Villevieille C, Novák P (2014) Ex situ and in situ Raman microscopic investigation of the differences between stoichiometric LiMO2 and high-energy xLi2MnO3·(1 − x)LiMO2 (M = Ni, Co, Mn)· Electrochim Acta 130: 206–212. doi: 10.1016/j.electacta.2014.03.004
    [67] Koga H, Croguennec L, Mannessiez P, et al. (2012) Li1.20Mn0.54Co0.13Ni0.13O2 with different particle sizes as attractive positive electrode materials for lithium-ion batteries: Insights into their structure. J Phys Chem C 116: 13497–13506. doi: 10.1021/jp301879x
    [68] Julien CM, Massot M (2003) Lattice vibrations of materials for lithium rechargeable batteries III. Lithium manganese oxides. Mater Sci Eng B-Adv 100: 69–78. doi: 10.1016/S0921-5107(03)00077-1
    [69] Singh G, West WC, Soler J, et al. (2012) In situ Raman spectroscopy of layered solid solution Li2MnO3–LiMO2 (M = Ni, Mn, Co). J Power Sources 218: 34–38. doi: 10.1016/j.jpowsour.2012.06.083
    [70] Lanz P, Villevieille C, Novák P (2013) Electrochemical activation of Li2MnO3 at elevated temperature investigated by in situ Raman microscopy. Electrochim Acta 109: 426–432. doi: 10.1016/j.electacta.2013.07.130
    [71] Rao CV, Soler J, Katiyar R, et al. (2014) Investigations on electrochemical behavior and structural stability of Li1.2Mn0.54Ni0.13Co0.13O2 lithium-ion cathodes via in-situ and ex-situ Raman spectroscopy. J Phys Chem C 118: 14133–14141. doi: 10.1016/j.jpowsour.2014.10.032
    [72] Hy S, Felix F, Rick J, et al. (2014) Direct in situ observation of Li2O evolution on Li-rich high-capacity cathode material, Li[NixLi(1−2x)/3Mn(2−x)/3]O2 (0 ≤ x ≤ 0.5). J Am Chem Soc 136: 999–1007. doi: 10.1021/ja410137s
    [73] Shojan J, Rao-Chitturi V, Soler J, et al. (2015) High energy xLi2MnO3·(1 − x)LiNi2/3Co1/6Mn1/6O2 composite cathode for advanced Li-ion batteries.J Power Sources 274: 440–450. doi: 10.1016/j.jpowsour.2014.10.032
    [74] Grey CP, Yoon W, Reed J, et al. (2004) Electrochemical activity of Li in the transition-metal sites of O3 Li[Li (1−2x)/3Mn(2−x)/3Nix]O2. Electrochem Solid St 7: A290–A 293. doi: 10.1149/1.1783113
    [75] Amalraj F, Talianker M, Markovsky B, et al. (2013) Study of the lithium-rich integrated compound xLi2MnO3·(1 − x)LiMO2 (x around 0.5; M = Mn, Ni, Co; 2:2:1) and its electrochemical activity as positive electrode in lithium cells. J Electrochem Soc 160: A324–A337.
    [76] Kumar-Nayak P, Grinblat J, Levi M, et al. (2014) Electrochemical and structural characterization of carbon coated Li1.2Mn0.56Ni0.16Co0.08O2 and Li1.2Mn0.6Ni0.2O2 as cathode materials for Li-ion batteries. Electrochim Acta 137: 546–556. doi: 10.1016/j.electacta.2014.06.055
    [77] Wu Q, Maroni VA, Gosztola DJ, et al. (2015) A Raman-based investigation of the fate of Li2MnO3 in lithium- and manganese-rich cathode materials for lithium ion batteries. J Electrochem Soc 162: A1255–A1264. doi: 10.1149/2.0631507jes
    [78] Bak SM, Nam KW, Chang W, et al. (2013) Correlating structure changes and gas evolution during the thermal decomposition of charged LixNi0.8Co0.15Al0.05O2 cathode material. Chem Mater 25: 337–351. doi: 10.1021/cm303096e
    [79] Julien C, Camacho-Lopez MA, Lemal M, et al. (2002) LiCo1−yMyO2 positive electrodes for rechargeable lithium batteries. I. Aluminum doped materials. Mater Sci Eng B-Adv 95: 6–13. doi: 10.1016/S0921-5107(01)01010-8
    [80] Julien CM, Mauger A, Vijh A, et al. (2016) Cathode Materials with Monoatomic Ions in a Three-Dimensional Framework, In: Julien CM, Mauger A, Vijh A, et al., Lithium Batteries: Science and Technology, Cham, Switzerland: Springer, 163–199.
    [81] Ohzuku T, Kitagawa M, Hirai T (1990) Electrochemistry of manganese dioxide in lithium nonaqueous cell. III. X-ray diffractional study on the reduction of spinel-related manganese dioxide. J Electrochem Soc 137: 769–775.
    [82] Liu W, Kowal K, Farrington GC (1998) Mechanism of the electrochemical insertion of lithium into LiMn2O4 spinels. J Electrochem Soc 145: 459–465. doi: 10.1149/1.1838285
    [83] Xia Y, Yoshio M (1996) An investigation of lithium ion insertion into spinel structure Li–Mn–O compounds. J Electrochem Soc 143: 825–833. doi: 10.1149/1.1836544
    [84] Richard MN, Keotschau I, Dahn JR (1997) A cell for in situ x‐ray diffraction based on coin cell hardware and Bellcore plastic electrode technology. J Electrochem Soc 144: 554–557. doi: 10.1149/1.1837447
    [85] Yang XQ, Sun X, Lee SJ, et al. (1999) In situ synchrotron X‐ray diffraction studies of the phase transitions in LixMn2O4 cathode Materials. Electrochem Solid St 2: 157–160.
    [86] Kanamura K, Naito H, Yao T, et al. (1996) Structural change of the LiMn2O4 spinel structure induced by extraction of lithium. J Mater Chem 6: 33–36. doi: 10.1039/jm9960600033
    [87] White WB, DeAngelis A (1967) Interpretation of the vibrational spectra of spinels. Spectrochim Acta A 23: 985–995. doi: 10.1016/0584-8539(67)80023-0
    [88] Ammundsen B, Burns GR, Islam MS, et al. (1999) Lattice dynamics and vibrational spectra of lithium manganese oxides: A computer simulation and spectroscopic study. J Phys Chem B 103: 5175–5180. doi: 10.1021/jp984398l
    [89] Kanoh H, Tang W, Ooi K (1998) In situ Raman spectroscopic study on electroinsertion of Li+ into a Pt/l-MnO2 electrode in aqueous solution. Electrochem Solid St 1: 17–19.
    [90] Huang W, Frech R (1999) In situ Raman spectroscopic studies of electrochemical intercalation in LixMn2O4-based cathodes. J Power Sources 81–82: 616–620. doi: 10.1016/S0378-7753(99)00231-1
    [91] Julien C, Massot M (2003) Lattice vibrations of materials for lithium rechargeable batteries I. Lithium manganese oxide spinel. Mater Sci Eng B-Adv 97: 217–230. doi: 10.1016/j.mseb.2003.09.028
    [92] Dokko K, Shi Q, Stefan IC, et al. (2003) In situ Raman spectroscopy of single microparticle Li+-intercalation electrodes. J Phys Chem B 107: 12549–12554. doi: 10.1021/jp034977c
    [93] Anzue N, Itoh T, Mohamedi M, et al. (2003) In situ Raman spectroscopic study of thin-film Li1−xMn2O4 electrodes. Solid State Ionics 156: 301–307. doi: 10.1016/S0167-2738(02)00693-8
    [94] Shi Q, Takahashi Y, Akimoto J, et al. (2005) In situ Raman scattering measurements of a LiMn2O4 single crystal microelectrode. Electrochem Solid St 8: A521–A524. doi: 10.1149/1.2030507
    [95] Sun X, Yang XQ, Balasubramanian M, et al. (2002) In situ investigation of phase transitions of Li1+yMn2O4 spinel during Li-ion extraction and insertion. J Electrochem Soc 149: A842–A848.
    [96] Michalska M, Ziołkowska DA, Jasinski JB, et al. (2018) Improved electrochemical performance of LiMn2O4 cathode material by Ce doping. Electrochim Acta 276: 37–46. doi: 10.1016/j.electacta.2018.04.165
    [97] Liu J, Manthiram A (2009) Understanding the improved electrochemical performances of Fe-substituted 5 V spinel cathode LiMn1.5Ni0.5O4. J Phys Chem C 113: 15073–15079.
    [98] Alcantara R, Jaraba M, Lavela P, et al. (2003) Structural and electrochemical study of new LiNi0.5TixMn1.5−xO4 spinel oxides for 5-V cathode materials. Chem Mater 15: 2376–2382. doi: 10.1021/cm034080s
    [99] Zhong GB, Wang YY, Yu YQ, et al. (2012) Electrochemical investigations of the LiNi0.45M0.10Mn1.45O4 (M = Fe, Co, Cr) 5 V cathode materials for lithium ion batteries. J Power Sources 205: 385–393.
    [100] Liu D, Hamel-Paquet J, Trottier J, et al. (2012) Synthesis of pure phase disordered LiMn1.45Cr0.1Ni0.45O4 by a post-annealing method. J Power Sources 217: 400–406. doi: 10.1016/j.jpowsour.2012.06.063
    [101] Zhu W, Liu D, Trottier J, et al. (2014) Comparative studies of the phase evolution in M-doped LixMn1.5Ni0.5O4 (M = Co, Al, Cu and Mg) by in-situ X-ray diffraction. J Power Sources 264: 290–298. doi: 10.1016/j.jpowsour.2014.03.122
    [102] Dokko K, Mohamedi M, Anzue N, et al. (2002) In situ Raman spectroscopic studies of LiNixMn2−xO4 thin film cathode materials for lithium ion secondary batteries. J Mater Chem 12: 3688–3693. doi: 10.1039/B206764A
    [103] Zhu W, Liu D, Trottier J, et al. (2014) In situ Raman spectroscopic investigation of LiMn1.45Ni0.45M0.1O4 (M = Cr, Co) 5 V cathode materials. J Power Sources 298: 341–348.
    [104] Cocciantelli JM, Doumerc JP, Pouchard M, et al. (1991) Crystal chemistry of electrochemically inserted LixV2O5. J Power Sources 34: 103–111. doi: 10.1016/0378-7753(91)85029-V
    [105] Zhang X, Frech R (1996) In situ Raman spectroscopy of LixV2O5 in a lithium rechargeable battery. ECS Proceedings 17: 198–203.
    [106] Baddour-Hadjean R, Navone C, Pereira-Ramos JP (2009) In situ Raman microspectrometry investigation of electrochemical lithium intercalation into sputtered crystalline V2O5 thin films. Electrochim Acta 54: 6674–6679. doi: 10.1016/j.electacta.2009.06.052
    [107] Jung H, Gerasopoulos K, Talin AA, et al. (2017) In situ characterization of charge rate dependent stress and structure changes in V2O5 cathode prepared by atomic layer deposition. J Power Sources 340: 89–97. doi: 10.1016/j.jpowsour.2016.11.035
    [108] Jung H, Gerasopoulos K, Talin AA, et al. (2017) A platform for in situ Raman and stress characterizations of V2O5 cathode using MEMS device. Electrochim Acta 242: 227–239. doi: 10.1016/j.electacta.2017.04.160
    [109] Julien C, Ivanov I, Gorenstein A (1995) Vibrational modifications on lithium intercalation in V2O5 films. Mater Sci Eng B-Adv 33: 168–172. doi: 10.1016/0921-5107(95)80032-8
    [110] Baddour-Hadjean R, Pereira-Ramos JP, Navone C, et al. (2008) Raman microspectrometry study of electrochemical lithium intercalation into sputtered crystalline V2O5 thin films. Chem Mater 20: 1916–1923. doi: 10.1021/cm702979k
    [111] Baddour-Hadjean R, Raekelboom E, Pereira-Ramos JP (2006) New structural characterization of the LixV2O5 system provided by Raman spectroscopy. Chem Mater 18: 3548–3556. doi: 10.1021/cm060540g
    [112] Baddour-Hadjean R, Marzouk A, Pereira-Ramos JP (2012) Structural modifications of LixV2O5 in a composite cathode (0 < x < 2) investigated by Raman microspectrometry. J Raman Spectrosc 43: 153–160. doi: 10.1002/jrs.2984
    [113] Baddour-Hadjean R, Pereira-Ramos JP (2007) New structural approach of lithium intercalation using Raman spectroscopy. J Power Sources 174: 1188–1192. doi: 10.1016/j.jpowsour.2007.06.177
    [114] Julien C, Massot M (2002) Spectroscopic studies of the local structure in positive electrodes for lithium batteries. Phys Chem Chem Phys 4: 4226–4235. doi: 10.1039/b203361e
    [115] Chen D, Ding D, Li X, et al. (2015) Probing the charge storage mechanism of a pseudocapacitive MnO2 electrode using in operando Raman spectroscopy. Chem Mater 27: 6608–6619. doi: 10.1021/acs.chemmater.5b03118
    [116] Julien CM, Massot M, Poinsignon C (2004) Lattice vibrations of manganese oxides: Part I. Periodic structures. Spectrochim Acta A 60: 689–700. doi: 10.1016/S1386-1425(03)00279-8
    [117] Hashem AM, Abuzeid H, Kaus M, et al. (2018) Green synthesis of nanosized manganese dioxide as positive electrode for lithium-ion batteries using lemon juice and citrus peel. Electrochim Acta 262: 74–81. doi: 10.1016/j.electacta.2018.01.024
    [118] Julien C, Massot M, Baddour-Hadjean R, et al. (2003) Raman spectra of birnessite manganese dioxides. Solid State Ionics 159: 345–356. doi: 10.1016/S0167-2738(03)00035-3
    [119] Hwang SJ, Park HS, Choy JH, et al. (2001) Micro-Raman spectroscopic study on layered lithium manganese oxide and its delithiated/relithiated derivatives. Electrochem Solid St 4: A213–A216. doi: 10.1149/1.1413704
    [120] Chen D, Ding D, Li X, et al. (2015) Probing the charge storage mechanism of a pseudocapacitive MnO2 electrode using in operando Raman spectroscopy. Chem Mater 27: 6608–6619. doi: 10.1021/acs.chemmater.5b03118
    [121] Yang L, Cheng S, Wang J, et al. (2016) Investigation into the origin of high stability of δ-MnO2 pseudo-capacitive electrode using operando Raman spectroscopy. Nano Energy 30: 293–302. doi: 10.1016/j.nanoen.2016.10.018
    [122] Zaghib K, Dontigny M, Perret P, et al. (2014) Electrochemical and thermal characterization of lithium titanate spinel anode in C-LiFePO4//C-Li4Ti5O12 cells at sub-zero temperatures. J Power Sources 248: 1050–1057. doi: 10.1016/j.jpowsour.2013.09.083
    [123] Paques-Ledent MT, Tarte P (1974) Vibrational studies of olivine-type compounds 2. Orthophosphates, orthoarsenates and orthovanadates AIBIIXVO4. Spectrochim Acta A 30: 673–689.
    [124] Wu J, Dathar GK, Sun C, et al. (2013) In situ Raman spectroscopy of LiFePO4: size and morphology dependence during charge and self-discharge. Nanotechnology 24: 424009.
    [125] Ramana CV, Mauger A, Gendron F, et al. (2009) Study of the Li-insertion/extraction process in LiFePO4/FePO4. J Power Sources 187: 555–564. doi: 10.1016/j.jpowsour.2008.11.042
    [126] Siddique NA, Salehi A, Wei Z, et al. (2015) Length‐scale‐dependent phase transformation of LiFePO4: An in situ and operando study using micro‐Raman spectroscopy and XRD. ChemPhysChem 16: 2383–2388. doi: 10.1002/cphc.201500299
    [127] Burba CM, Frech R (2004) Raman and FTIR spectroscopic study of LixFePO4 (0 ≤ x ≤ 1). J Electrochem Soc 151: A1032–A1038. doi: 10.1149/1.1756885
    [128] Morita M, Asai Y, Yoshimoto N, et al. (1998) A Raman spectroscopic study of organic electrolyte solutions based on binary solvent systems of ethylene carbonate with low viscosity solvents which dissolve different lithium salts. J Chem Soc Faraday Trans 94: 3451–3456. doi: 10.1039/a806278a
    [129] Burba CM, Frech R (2007) Vibrational spectroscopic studies of monoclinic and rhombohedral Li3V2(PO4)3. Solid State Ionics 177: 3445–3454. doi: 10.1016/j.ssi.2006.09.017
    [130] Yin SC, Grondey H, Strobel P, et al. (2003) Electrochemical property: Structure relationships in monoclinic Li3V2(PO4)3. J Am Chem Soc 125: 10402–10411. doi: 10.1021/ja034565h
    [131] Takeuchi KJ, Marschilok AC, Davis SM, et al. (2001) Silver vanadium oxide and related battery applications. Coordin Chem Rev 219–221: 283–310.
    [132] Bao Q, Bao S, Li CM, et al. (2007) Lithium insertion in channel-structured b-AgVO3: in situ Raman study and computer simulation. Chem Mater 19: 5965–5972. doi: 10.1021/cm071728i
    [133] Tuinstra F, Koenig JL (1970) Raman spectrum of graphite. J Chem Phys 53: 1126–1130. doi: 10.1063/1.1674108
    [134] Reich S, Thomsen C (2004) Raman spectroscopy of graphite. Philos T R Soc A 362: 2271–2288. doi: 10.1098/rsta.2004.1454
    [135] Inaba M, Yoshida H, Yida H, et al. (1995) In situ Raman study on electrochemical Li intercalation into graphite. J Electrochem Soc 142: 20–26. doi: 10.1149/1.2043869
    [136] Huang W, Frech R (1998) In situ Raman studies of graphite surface structures during lithium electrochemical intercalation. J Electrochem Soc 145: 765–770.
    [137] Novák P, Joho F, Imhof R, et al. (1999) In situ investigation of the interaction between graphite and electrolyte solutions. J Power Sources 81–82: 212–216.
    [138] Luo Y, Cai WB, Scherson DA (2002) In situ, real-time Raman microscopy of embedded single particle graphite electrodes. J Electrochem Soc 149: A1100–A1105. doi: 10.1149/1.1492286
    [139] Kostecki R, McLarnon F (2003) Microprobe study of the effect of Li intercalation on the structure of graphite. J Power Sources 119–121: 550–554.
    [140] Luo Y, Cai WB, Xing XK, et al. (2004) In situ, time-resolved Raman spectromicrotopography of an operating lithium-ion battery. Electrochem Solid St 7: E1–E5. doi: 10.1149/1.1627972
    [141] Shi Q, Dokko K, Scherson DA (2004) In situ Raman microscopy of a single graphite microflake electrode in a Li+-containing electrolyte. J Phys Chem B 108: 4798–4793.
    [142] Migge S, Sandmann G, Rahner D, et al. (2005) Studying lithium intercalation into graphite particles via in situ Raman spectroscopy and confocal microscopy. J Solid State Electr 9: 132–137. doi: 10.1007/s10008-004-0563-4
    [143] Hardwick LJ, Hahn M, Ruch P, et al. (2006) Graphite surface disorder detection using in situ Raman microscopy. Solid State Ionics 177: 2801–2806. doi: 10.1016/j.ssi.2006.03.032
    [144] Hardwick LJ, Buqa H, Holzapfel M, et al. (2007) Behaviour of highly crystalline graphitic materials in lithium-ion cells with propylene carbonate containing electrolytes: An in situ Raman and SEM study. Electrochim Acta 52: 4884–4891. doi: 10.1016/j.electacta.2006.12.081
    [145] Hardwick LJ, Marcinek M, Beer L, et al. (2008) An investigation of the effect of graphite degradation on irreversible capacity in lithium-ion cells. J Electrochem Soc 155: A442–A447.
    [146] Hardwick LJ, Ruch PW, Hahn M, et al. (2008) In situ Raman spectroscopy of insertion electrodes for lithium-ion batteries and supercapacitors: First cycle effects. J Phys Chem Solids 69: 1232–1237. doi: 10.1016/j.jpcs.2007.10.017
    [147] Nakagawa H, Ochida M, Domi Y, et al. (2012) Electrochemical Raman study of edge plane graphite negative-electrodes in electrolytes containing trialkyl phosphoric ester. J Power Sources 212: 148–153. doi: 10.1016/j.jpowsour.2012.04.013
    [148] Underhill C, Leung SY, Dresselhaus G, et al. (1979) Infrared and Raman spectroscopy of graphite-ferric chloride. Solid State Commun 29: 769–774.
    [149] Hardwick LJ, Hahn M, Ruch P, et al. (2006) An in situ Raman study of the intercalation of supercapacitor-type electrolyte into microcrystalline graphite. Electrochim Acta 52: 675–680. doi: 10.1016/j.electacta.2006.05.053
    [150] Solin SA (1990) Graphite Intercalation Compounds, Berlin: Springer-Verlag.
    [151] Brancolini G, Negri F (2004) Quantum chemical modeling of infrared and Raman activities in lithium-doped amorphous carbon nanostructures: hexa-peri-hexabenzocoronene as a model for hydrogen-rich carbon materials. Carbon 42: 1001–1005. doi: 10.1016/j.carbon.2003.12.016
    [152] Nakagawa H, Domi Y, Doi T, et al. (2012) In situ Raman study on degradation of edge plane graphite negative-electrodes and effects of film-forming additives. J Power Sources 206: 320–324. doi: 10.1016/j.jpowsour.2012.01.141
    [153] Perez-Villar S, Lanz P, Schneider H, et al. (2013) Characterization of a model solid electrolyte interphase/carbon interface by combined in situ Raman/Fourier transform infrared microscopy. Electrochim Acta 106: 506–515. doi: 10.1016/j.electacta.2013.05.124
    [154] Nakagawa H, Domi Y, Doi T, et al. (2013) In situ Raman study on the structural degradation of a graphite composite negative-electrode and the influence of the salt in the electrolyte solution. J Power Sources 236: 138–144. doi: 10.1016/j.jpowsour.2013.02.060
    [155] Lanz P, Novák P (2014) Combined in situ Raman and IR microscopy at the interface of a single graphite particle with ethylene carbonate/dimethyl carbonate. J Electrochem Soc 161: A1555–A1563. doi: 10.1149/2.0021410jes
    [156] Domi Y, Doi T, Nakagawa H, et al. (2016) In situ Raman study on reversible structural changes of graphite negative-electrodes at high potentials in LiPF6-Based electrolyte solution. J Electrochem Soc 163: A2435–A2440. doi: 10.1149/2.1301610jes
    [157] Yamanaka T, Nakagawa H, Tsubouchi S, et al. (2017) Correlations of concentration changes of electrolyte salt with resistance and capacitance at the surface of a graphite electrode in a lithium ion battery studied by in situ microprobe Raman spectroscopy. Electrochim Acta 251: 301–306. doi: 10.1016/j.electacta.2017.08.119
    [158] Bhattacharya S, Alpas AT (2018) A novel elevated temperature pre-treatment for electrochemical capacity enhancement of graphene nanoflake-based anodes. Mater Renew Sustain Energy 7: 3.
    [159] Inaba M, Yoshida H, Ogumi Z (1996) In situ Raman study of electrochemical lithium insertion into mesocarbon microbeads heat‐treated at various temperatures. J Electrochem Soc 143: 2572–2578. doi: 10.1149/1.1837049
    [160] Totir DA, Scherson DA (2000) Electrochemical and in situ Raman studies of embedded carbon particle electrodes in nonaqueous liquid electrolytes. Electrochem Solid St 3: 263–265.
    [161] Ramesha GK, Sampath S (2009) Electrochemical reduction of oriented graphene oxide films: an in situ Raman spectroelectrochemical study. J Phys Chem C 113: 7985–7989. doi: 10.1021/jp811377n
    [162] Tang W, Goh BM, Hu MY, et al. (2016) In situ Raman and nuclear magnetic resonance study of trapped lithium in the solid electrolyte interface of reduced graphene oxide. J Phys Chem C 120: 2600–2608. doi: 10.1021/acs.jpcc.5b12551
    [163] Julien CM, Massot M, Zaghib K (2007) Structural studies of Li4/3Me5/3O4 (Me = Ti, Mn) electrode materials: local structure and electrochemical aspects. J Power Sources 136: 72–79.
    [164] Mukai K, Kato Y, Nakano H (2014) Understanding the zero-strain lithium insertion scheme of Li[Li1/3Ti5/3]O4: structural changes at atomic scale clarified by Raman spectroscopy. J Phys Chem C 118: 2992–2999. doi: 10.1021/jp412196v
    [165] Shu J, Shui M, Xu D, et al. (2011) Design and comparison of ex situ and in situ devices for Raman characterization of lithium titanate anode material. Ionics 17: 503–509. doi: 10.1007/s11581-011-0544-4
    [166] Julien CM, Mauger A, Vijh A, et al. (2016) Anodes for Li-Ion Batteries, In: Julien CM, Mauger A, Vijh A, et al., Lithium Batteries: Science and Technology, Cham, Switzerland: Springer, 323–429.
    [167] Ohzuku T, Takehara Z, Yoshizawa S (1979) Non-aqueous lithium/titanium dioxide cell. Electrochim Acta 24: 219–222. doi: 10.1016/0013-4686(79)80028-6
    [168] Wagemaker M, Borghols WJH, Mulder FM (2007) Large impact of particle size on insertion reactions. A case for anatase LixTiO2. J Am Chem Soc 129: 4323–4327.
    [169] Mukai K, Kato Y, Nakano H (2014) Understanding the zero-strain lithium insertion scheme of Li[Li1/3Ti5/3]O4: structural changes at atomic scale clarified by Raman spectroscopy. J Phys Chem C 118: 2992–2999. doi: 10.1021/jp412196v
    [170] Dinh NN, Oanh NTT, Long PD, et al. (2003) Electrochromic properties of TiO2 anatase thin films prepared by a dipping sol–gel method. Thin Solid Films 423: 70–76. doi: 10.1016/S0040-6090(02)00948-3
    [171] Smirnov M, Baddour-Hadjean R (2004) Li intercalation in TiO2 anatase: Raman spectroscopy and lattice dynamic studies. J Chem Phys 121: 2348–2355. doi: 10.1063/1.1767993
    [172] Baddour-Hadjean R, Bach S, Smirnov M, et al. (2004) Raman investigation of the structural changes in anatase LixTiO2 upon electrochemical lithium insertion. J Raman Spectrosc 35: 577–585. doi: 10.1002/jrs.1200
    [173] Hardwick LJ, Holzapfel M, Novák P, et al. (2007) Electrochemical lithium insertion into anatase-type TiO2: An in situ Raman microscopy investigation. Electrochim Acta 52: 5357–5367. doi: 10.1016/j.electacta.2007.02.050
    [174] Ren Y, Hardwick LJ, Bruce PG (2010) Lithium intercalation into mesoporous anatase with an ordered 3D pore structure. Angew Chem Int Edit 49: 2570–2574. doi: 10.1002/anie.200907099
    [175] Gentili V, Brutti S, Hardwick LJ, et al. (2012) Lithium insertion into anatase nanotubes. Chem Mater 24: 4468–4476. doi: 10.1021/cm302912f
    [176] Laskova BP, Frank O, Zukalova M, et al. (2013) Lithium insertion into titanium dioxide (anatase): A Raman study with 16/18O and 6/7Li isotope labeling. Chem Mater 25: 3710–3717. doi: 10.1021/cm402056j
    [177] Laskova BP, Kavan L, Zukalova M, et al. (2016) In situ Raman spectroelectrochemistry as a useful tool for detection of TiO2(anatase) impurities in TiO2(B) and TiO2(rutile). Monatsh Chem 147: 951–959. doi: 10.1007/s00706-016-1678-x
    [178] Mauger A, Xie H, Julien CM, (2016) Composite anodes for lithium-ion batteries, status and trends. AIMS Mater Sci 3: 1054–1106. doi: 10.3934/matersci.2016.3.1054
    [179] Li H, Huang X, Chen L, et al. (2000) The crystal structural evolution of nano-Si anode caused by lithium insertion and extraction at room temperature. Solid State Ionics 135: 181–191. doi: 10.1016/S0167-2738(00)00362-3
    [180] Nanda J, Datta MK, Remillard JT, et al. (2009) In situ Raman microscopy during discharge of a high capacity silicon–carbon composite Li-ion battery negative electrode. Electrochem Commun 11: 235–237. doi: 10.1016/j.elecom.2008.11.006
    [181] Zeng Z, Liu N, Zeng Q, et al. (2016) In situ measurement of lithiation-induced stress in silicon nanoparticles using micro-Raman spectroscopy. Nano Energy 22: 105–110.
    [182] Tardif S, Pavlenko E, Quazuguel L, et al. (2017) Operando Raman spectroscopy and synchrotron X-ray diffraction of lithiation/delithiation in silicon nanoparticle anodes. ACS Nano 11: 11306–11316. doi: 10.1021/acsnano.7b05796
    [183] Miroshnikov Y, Zitoun D (2017) Operando plasmon-enhanced Raman spectroscopy in silicon anodes for Li-ion battery. J Nanopart Res 19: 372. doi: 10.1007/s11051-017-4063-8
    [184] Holzapfel M, Buqa H, Hardwick LJ, et al. (2006) Nano silicon for lithium-ion batteries. Electrochim Acta 52: 973–978. doi: 10.1016/j.electacta.2006.06.034
    [185] Long BR, Chan MKY, Greeley JP, et al. (2011) Dopant modulated Li insertion in Si for battery anodes: theory and experiment. J Phys Chem C 115: 18916–18921. doi: 10.1021/jp2060602
    [186] Miroshnikov Y, Yang J, Shokhen V, et al. (2018) Operando micro-Raman study revealing enhanced connectivity of plasmonic metals decorated silicon anodes for lithium-ion batteries. ACS Appl Energy Mater 1: 1096–1105. doi: 10.1021/acsaem.7b00220
    [187] Zeng Z, Liu N, Zeng Q, et al. (2016) In situ measurement of lithiation-induced stress in silicon nanoparticles using micro-Raman spectroscopy. Nano Energy 22: 105–110.
    [188] Yang J, Kraytsberg A, Ein-Eli Y (2015) In situ Raman spectroscopy mapping of Si based anode material lithiation. J Power Sources 282: 294–298. doi: 10.1016/j.jpowsour.2015.02.044
    [189] Shimizu M, Usui H, Suzumura T, et al. (2015) Analysis of the deterioration mechanism of Si electrode as a Li-ion battery anode using Raman microspectroscopy. J Phys Chem C 119: 2975–2982. doi: 10.1021/jp5121965
    [190] Domi Y, Usui H, Shimizu M, et al. (2016) Effect of phosphorus-doping on electrochemical performance of silicon negative electrodes in lithium-ion batteries. ACS Appl Mater Inter 8: 7125–7132. doi: 10.1021/acsami.6b00386
    [191] Yamaguchi K, Domi Y, Usui H, et al. (2017) Elucidation of the reaction behavior of silicon negative electrodes in a bis(fluorosulfonyl)amide based ionic liquid electrolyte. ChemElectroChem 4: 3257–3263. doi: 10.1002/celc.201700724
    [192] Zhang WJ (2011) A review of the electrochemical performance of alloy anodes for lithium-ion batteries. J Power Sources 196: 13–24. doi: 10.1016/j.jpowsour.2010.07.020
    [193] Weppner W, Huggins RA (1977) Determination of the kinetic parameters of mixed-conducting electrodes and application to the system Li3Sb. J Electrochem Soc 124: 1569–1578. doi: 10.1149/1.2133112
    [194] Drewett NE, Aldous AL, Zou J, et al. (2017) In situ Raman spectroscopic analysis of the lithiation and sodiation of antimony microparticles. Electrochim Acta 247: 296–305. doi: 10.1016/j.electacta.2017.07.030
    [195] Zhong X, Wang X, Wang H, et al. (2018) Ultrahigh-performance mesoporous ZnMn2O4 microspheres as anode materials for lithium-ion batteries and their in situ Raman investigation. Nano Res 11: 3814. doi: 10.1007/s12274-017-1955-y
    [196] Cabo-Fernandez L, Mueller F, Passerini S, et al. (2016) In situ Raman spectroscopy of carbon-coated ZnFe2O4 anode material in Li-ion batteries-Investigation of SEI growth. Chem Commun 52: 3970–3973. doi: 10.1039/C5CC09350C
    [197] Schmitz R, Muller RA, Schmitz RW, et al. (2013) SEI investigations on copper electrodes after lithium plating with Raman spectroscopy and mass spectrometry. J Power Sources 233: 110–114. doi: 10.1016/j.jpowsour.2013.01.105
    [198] Hy S, Felix F, Chen YH, et al. (2014) In situ surface enhanced Raman spectroscopic studies of solid electrolyte interphase formation in lithium ion battery electrodes. J Power Sources 256: 324–328. doi: 10.1016/j.jpowsour.2014.01.092
    [199] Yamanaka T, Nakagawa H, Tsubouchi S, et al. (2017) In situ diagnosis of the electrolyte solution in a laminate lithium ion battery by using ultrafine multi-probe Raman spectroscopy. J Power Sources 359: 435–440.
    [200] Yamanaka T, Nakagawa H, Tsubouchi S, et al. (2017) In situ Raman spectroscopic studies on concentration change of electrolyte salt in a lithium ion model battery with closely faced graphite composite and LiCoO2 composite electrodes by using an ultrafine microprobe. Electrochim Acta 234: 93–98. doi: 10.1016/j.electacta.2017.03.060
    [201] Yamanaka T , Nakagawa H, Tsubouchi S, et al. (2017) Effects of pored separator films at the anode and cathode sides on concentration changes of electrolyte salt in lithium ion batteries. Jpn J Appl Phys 56: 128002. doi: 10.7567/JJAP.56.128002
    [202] Naudin C, Bruneel JL, Chami M, et al. (2003) Characterization of the lithium surface by infrared and Raman spectroscopies. J Power Sources 124: 518–525. doi: 10.1016/S0378-7753(03)00798-5
    [203] Galloway TA, Cabo-Fernandez L, Aldous IM, et al. (2017) Shell isolated nanoparticles for enhanced Raman spectroscopy studies in lithium–oxygen cells. Faraday Discuss 205: 469–490. doi: 10.1039/C7FD00151G
    [204] Park Y, Kim Y, Kim SM, et al. (2017) Reaction at the electrolyte–electrode interface in a Li-ion battery studied by in situ Raman spectroscopy. Bull Korean Chem Soc 38: 511–513. doi: 10.1002/bkcs.11117
    [205] Yamanaka T, Nakagawa H, Tsubouchi S, et al. (2017) In situ Raman spectroscopic studies on concentration of electrolyte salt in lithium ion batteries by using ultrafine multifiber probes. ChemSusChem 10: 855–861. doi: 10.1002/cssc.201601473
    [206] Lewis IR, Griffiths PR (1996) Raman spectrometry with fiber-optic sampling. Appl Spectrosc 50: 12A–30A.
    [207] Yamanaka T, Nakagawa H, Ochida M, et al. (2016) Ultrafine fiber Raman probe with high spatial resolution and fluorescence noise reduction. J Phys Chem C 120: 2585–2591. doi: 10.1021/acs.jpcc.5b11894
    [208] Yamanaka T, Nakagawa H, Tsubouchi S, et al. (2017) In situ Raman spectroscopic studies on concentration change of ions in the electrolyte solution in separator regions in a lithium ion battery by using multi-microprobes. Electrochem Commun 77: 32–35. doi: 10.1016/j.elecom.2017.01.020
    [209] Yamanaka T, Nakagawa H, Tsubouchi S, et al. (2017) Modification of the solid electrolyte interphase by chronoamperometric pretreatment and its effect on the concentration change of electrolyte salt in lithium ion batteries studied by in situ microprobe Raman spectroscopy. J Electrochem Soc 164: A2355–A2359. doi: 10.1149/2.0601712jes
  • This article has been cited by:

    1. Toshiro Yamanaka, Ken-ichi Okazaki, Takeshi Abe, Koji Nishio, Zempachi Ogumi, Evolution of Reactions of a Fluoride Shuttle Battery at the Surfaces of BiF3 Microclusters Studied by In Situ Raman Microscopy, 2019, 12, 18645631, 527, 10.1002/cssc.201802209
    2. Toshiro Yamanaka, Ken-ichi Okazaki, Zempachi Ogumi, Takeshi Abe, Reactivity and Mechanisms in Fluoride Shuttle Battery Reactions: Difference between Orthorhombic and Cubic BiF3 Single Microparticles, 2019, 2, 2574-0962, 8801, 10.1021/acsaem.9b01803
    3. Nataliia Mozhzhukhina, Eibar Flores, Robin Lundström, Ville Nyström, Paul G. Kitz, Kristina Edström, Erik J. Berg, Direct Operando Observation of Double Layer Charging and Early Solid Electrolyte Interphase Formation in Li-Ion Battery Electrolytes, 2020, 11, 1948-7185, 4119, 10.1021/acs.jpclett.0c01089
    4. Yasutaka Matsuda, Naoaki Kuwata, Tatsunori Okawa, Arunkumar Dorai, Osamu Kamishima, Junichi Kawamura, In situ Raman spectroscopy of Li CoO2 cathode in Li/Li3PO4/LiCoO2 all-solid-state thin-film lithium battery, 2019, 335, 01672738, 7, 10.1016/j.ssi.2019.02.010
    5. Madison R. Tuttle, Shiyu Zhang, Bisthiazolyl Quinones: Stabilizing Organic Electrode Materials with Sulfur-Rich Thiazyl Motifs, 2020, 32, 0897-4756, 255, 10.1021/acs.chemmater.9b03678
    6. Eibar Flores, Petr Novák, Ulrich Aschauer, Erik J. Berg, Cation Ordering and Redox Chemistry of Layered Ni-Rich LixNi1–2yCoyMnyO2: An Operando Raman Spectroscopy Study, 2020, 32, 0897-4756, 186, 10.1021/acs.chemmater.9b03202
    7. Ashraf Abdel-Ghany, Ahmed M. Hashem, Alain Mauger, Christian M. Julien, Lithium-Rich Cobalt-Free Manganese-Based Layered Cathode Materials for Li-Ion Batteries: Suppressing the Voltage Fading, 2020, 13, 1996-1073, 3487, 10.3390/en13133487
    8. Imanol Landa-Medrano, Aitor Eguia-Barrio, Susan Sananes-Israel, Silvia Lijó-Pando, Iker Boyano, Francisco Alcaide, Idoia Urdampilleta, Iratxe de Meatza, In Situ Analysis of NMC∣graphite Li-Ion Batteries by Means of Complementary Electrochemical Methods, 2020, 167, 1945-7111, 090528, 10.1149/1945-7111/ab8b99
    9. Dmitry V. Pelegov, Anastasia A. Koshkina, Boris N. Slautin, Vadim S. Gorshkov, Statistical Raman spectroscopy characterization of carbon additive in low‐C composites: Toward industrial quality control, 2019, 0377-0486, 10.1002/jrs.5604
    10. Timothy E. Rosser, Juliana P. S. Sousa, Yasmine Ziouani, Oleksandr Bondarchuk, Dmitri Y. Petrovykh, Xian-Kui Wei, Jo J. L. Humphrey, Marc Heggen, Yury V. Kolen'ko, Andrew J. Wain, Enhanced oxygen evolution catalysis by aluminium-doped cobalt phosphide through in situ surface area increase, 2020, 10, 2044-4753, 2398, 10.1039/D0CY00123F
    11. Christian M. Julien, Alain Mauger, Functional behavior of AlF3 coatings for high-performance cathode materials for lithium-ion batteries, 2019, 6, 2372-0484, 406, 10.3934/matersci.2019.3.406
    12. C. Heubner, U. Langklotz, C. Lämmel, M. Schneider, A. Michaelis, Electrochemical single-particle measurements of electrode materials for Li-ion batteries: Possibilities, insights and implications for future development, 2020, 330, 00134686, 135160, 10.1016/j.electacta.2019.135160
    13. M.A. Cabañero, M Hagen, E. Quiroga-González, In-operando Raman study of lithium plating on graphite electrodes of lithium ion batteries, 2021, 374, 00134686, 137487, 10.1016/j.electacta.2020.137487
    14. Yuhuan Zhou, Hongfu Cui, Bao Qiu, Yuanhua Xia, Chong Yin, Liyang Wan, Zhepu Shi, Zhaoping Liu, Sufficient Oxygen Redox Activation against Voltage Decay in Li-Rich Layered Oxide Cathode Materials, 2021, 2639-4979, 433, 10.1021/acsmaterialslett.1c00088
    15. Ankur Baliyan, Hideto Imai, Machine Learning based Analytical Framework for Automatic Hyperspectral Raman Analysis of Lithium-ion Battery Electrodes, 2019, 9, 2045-2322, 10.1038/s41598-019-54770-2
    16. Toshiro Yamanaka, Asuman Celik Kucuk, Zempachi Ogumi, Takeshi Abe, Evolution of Fluoride Shuttle Battery Reactions of BiF3 Microparticles in a CsF/LiBOB/Tetraglyme Electrolyte: Dependence on Structure, Size, and Shape, 2020, 3, 2574-0962, 9390, 10.1021/acsaem.0c01753
    17. Nataliia Mozhzhukhina, Jolla Kullgren, Christian Baur, Olof Gustafsson, William R. Brant, Maximilian Fichtner, Daniel Brandell, Short‐range ordering in the Li‐rich disordered rock salt cathode material Li 2 VO 2 F revealed by Raman spectroscopy , 2020, 51, 0377-0486, 2095, 10.1002/jrs.5953
    18. Ali Reza Kamali, Dongwei Qiao, Zhongning Shi, Dexi Wang, Green molten salt modification of cobalt oxide for lithium ion battery anode application, 2021, 02540584, 124585, 10.1016/j.matchemphys.2021.124585
    19. Ankush Bhatia, Yosra Dridi Zrelli, Jean-Pierre Pereira-Ramos, Rita Baddour-Hadjean, Detailed redox mechanism and self-discharge diagnostic of 4.9 V LiMn1.5Ni0.5O4 spinel cathode revealed by Raman spectroscopy, 2021, 2050-7488, 10.1039/D1TA00989C
    20. Toshiro Yamanaka, Zempachi Ogumi, Takeshi Abe, Mechanisms of and three-dimensional morphology changes in fluoride shuttle battery reactions of PbF2 microparticles, 2021, 9, 2050-7488, 22544, 10.1039/D1TA06192E
    21. E. Salagre, S. Quílez, R. de Benito, M. Jaafar, H. P. van der Meulen, E. Vasco, R. Cid, E. J. Fuller, A. A. Talin, P. Segovia, E. G. Michel, C. Polop, A multi-technique approach to understanding delithiation damage in LiCoO2 thin films, 2021, 11, 2045-2322, 10.1038/s41598-021-91051-3
    22. Anna Szczęsna‐Chrzan, Tomasz Trzeciak, Magdalena Zybert, Hubert Ronduda, Andrzej Ostrowski, Maciej Trzaskowski, Marcin Drozd, Maciej Smoliński, Grażyna Zofia Żukowska, Wioletta Raróg‐Pilecka, Władysław Wieczorek, Alexander Buckel, Reza Younesi, Marek Marcinek, Systematic Studies on Liquid Sodium 4,5‐dicyano‐2‐(trifluoromethyl)imidazolate (NaTDI)‐Based Electrolytes and Its Impact on the Cycling Behaviour Against Wet Impregnated WI‐NaNMC and Prussian White Cathodes, 2022, 9, 2196-7350, 2102012, 10.1002/admi.202102012
    23. Jacopo Celè, Sylvain Franger, Yann Lamy, Sami Oukassi, Minimal Architecture Lithium Batteries: Toward High Energy Density Storage Solutions, 2023, 1613-6810, 2207657, 10.1002/smll.202207657
    24. Ang Fu, Zhengfeng Zhang, Jiande Lin, Yue Zou, Changdong Qin, Chuanjing Xu, Pengfei Yan, Ke Zhou, Jialiang Hao, Xuerui Yang, Yong Cheng, De-Yin Wu, Yong Yang, Ming-Sheng Wang, Jianming Zheng, Highly stable operation of LiCoO2 at cut-off ≥ 4.6 V enabled by synergistic structural and interfacial manipulation, 2022, 46, 24058297, 406, 10.1016/j.ensm.2022.01.033
    25. Ghulam Abbas, Zahid Ali Zafar, Farjana J. Sonia, Karel Knížek, Jana Houdková, Petr Jiříček, Martin Kalbáč, Jiří Červenka, Otakar Frank, The Effects of Ultrasound Treatment of Graphite on the Reversibility of the (De)Intercalation of an Anion from Aqueous Electrolyte Solution, 2022, 12, 2079-4991, 3932, 10.3390/nano12223932
    26. Xiaokang Ju, Xu Hou, Zhongqing Liu, Leilei Du, Li Zhang, Tangtang Xie, Elie Paillard, Taihong Wang, Martin Winter, Jie Li, Revealing the Effect of High Ni Content in Li‐Rich Cathode Materials: Mitigating Voltage Decay or Increasing Intrinsic Reactivity, 2023, 1613-6810, 2207328, 10.1002/smll.202207328
    27. Hua Wang, Ahmed M. Hashem, Ashraf E. Abdel-Ghany, Somia M. Abbas, Rasha S. El-Tawil, Tianyi Li, Xintong Li, Hazim El-Mounayri, Andres Tovar, Likun Zhu, Alain Mauger, Christian M. Julien, Effect of Cationic (Na+) and Anionic (F−) Co-Doping on the Structural and Electrochemical Properties of LiNi1/3Mn1/3Co1/3O2 Cathode Material for Lithium-Ion Batteries, 2022, 23, 1422-0067, 6755, 10.3390/ijms23126755
    28. Ulrike Boesenberg, Christian Henriksen, Kaare Lund Rasmussen, Yet-Ming Chiang, Jan Garrevoet, Dorthe B. Ravnsbæk, State of LiFePO4 Li-Ion Battery Electrodes after 6533 Deep-Discharge Cycles Characterized by Combined Micro-XRF and Micro-XRD, 2022, 5, 2574-0962, 4358, 10.1021/acsaem.1c03966
    29. Seunghwa Lee, You-Chiuan Chu, Lichen Bai, Hao Ming Chen, Xile Hu, Operando identification of a side-on nickel superoxide intermediate and the mechanism of oxygen evolution on nickel oxyhydroxide, 2023, 3, 26671093, 100475, 10.1016/j.checat.2022.11.014
    30. Ortal Lidor‐Shalev, Nicole Leifer, Michal Ejgenberg, Hagit Aviv, Ilana Perelshtein, Gil Goobes, Malachi Noked, , Molecular Layer Deposition of Alucone Thin Film on LiCoO 2 to Enable High Voltage Operation , 2021, 4, 2566-6223, 1739, 10.1002/batt.202100152
    31. Zhi Yi Leong, Jintao Zhang, Sareh Vafakhah, Meng Ding, Lu Guo, Hui Ying Yang, Electrochemically activated layered manganese oxide for selective removal of calcium and magnesium ions in hybrid capacitive deionization, 2021, 520, 00119164, 115374, 10.1016/j.desal.2021.115374
    32. Toshiro Yamanaka, Zempachi Ogumi, Takeshi Abe, Evolution of fluoride shuttle battery reactions and three-dimensional morphology changes of BiF3 microparticles in an ethylene carbonate-based liquid electrolyte, 2022, 1, 2753-1457, 632, 10.1039/D2YA00184E
    33. Ruth Pulido, Nelson Naveas, Raúl J. Martin-Palma, Fernando Agulló-Rueda, Victor R. Ferró, Jacobo Hernández-Montelongo, Gonzalo Recio-Sánchez, Ivan Brito, Miguel Manso-Silván, Phonon Structure, Infra-Red and Raman Spectra of Li2MnO3 by First-Principles Calculations, 2022, 15, 1996-1944, 6237, 10.3390/ma15186237
    34. Elhoucine Elmaataouy, Abdelwahed Chari, Marwa Tayoury, Jones Alami, Mouad Dahbi, 2021, Sol-Gel Synthesis of Li3VO4 as High-Capacity and High-Capability Anode for Lithium-Ion Batteries, 978-1-6654-1319-0, 1, 10.1109/IRSEC53969.2021.9741106
    35. Lydia Meyer, Najmus Saqib, Jason Porter, Review—Operando Optical Spectroscopy Studies of Batteries, 2021, 168, 0013-4651, 090561, 10.1149/1945-7111/ac2088
    36. Bruno Beccard, Shaileshkumar N. Karavadra, Sudhir Dahal, Lithium-Ion Battery Manufacturing and Quality Control: Raman Spectroscopy, an Analytical Technique of Choice, 2022, 1939-1900, 46, 10.56530/spectroscopy.sx2271c5
    37. Marcel Heber, Kathrin Hofmann, Christian Hess, Raman Diagnostics of Cathode Materials for Li-Ion Batteries Using Multi-Wavelength Excitation, 2022, 8, 2313-0105, 10, 10.3390/batteries8020010
    38. Yasaman Shirazi Moghadam, Sirshendu Dinda, Abdel El Kharbachi, Georgian Melinte, Christian Kübel, Maximilian Fichtner, Structural and Electrochemical Insights from the Fluorination of Disordered Mn-Based Rock Salt Cathode Materials, 2022, 34, 0897-4756, 2268, 10.1021/acs.chemmater.1c04059
    39. Sofia Marchesini, Benjamen P. Reed, Helen Jones, Lidija Matjacic, Timothy E. Rosser, Yundong Zhou, Barry Brennan, Mariavitalia Tiddia, Rhodri Jervis, Melanie. J. Loveridge, Rinaldo Raccichini, Juyeon Park, Andrew J. Wain, Gareth Hinds, Ian S. Gilmore, Alexander G. Shard, Andrew J. Pollard, Surface Analysis of Pristine and Cycled NMC/Graphite Lithium-Ion Battery Electrodes: Addressing the Measurement Challenges, 2022, 14, 1944-8244, 52779, 10.1021/acsami.2c13636
    40. Fabian Alexander Kreth, Lars Henning Hess, Andrea Balducci, In-operando GC-MS: A new tool for the understanding of degradation processes occurring in electrochemical capacitors, 2023, 56, 24058297, 192, 10.1016/j.ensm.2023.01.014
    41. Zehang Sun, Yamin Zhang, Yang Liu, Linrui Hou, Changzhou Yuan, Recent Progress on In Situ/Operando Characterization of Rechargeable Alkali Ion Batteries, 2021, 86, 2192-6506, 1487, 10.1002/cplu.202100393
    42. Taehoon Kim, Luis K. Ono, Yabing Qi, Understanding the nucleation and growth of the degenerated surface structure of the layered transition metal oxide cathodes for lithium-ion batteries by operando Raman spectroscopy, 2022, 915, 15726657, 116340, 10.1016/j.jelechem.2022.116340
    43. Yannik Moryson, Felix Walther, Joachim Sann, Boris Mogwitz, Shamail Ahmed, Simon Burkhardt, Limei Chen, Peter J. Klar, Kerstin Volz, Sarah Fearn, Marcus Rohnke, Jürgen Janek, Analyzing Nanometer-Thin Cathode Particle Coatings for Lithium-Ion Batteries—The Example of TiO2 on NCM622, 2021, 4, 2574-0962, 7168, 10.1021/acsaem.1c01255
    44. Toshiro Yamanaka, Zempachi Ogumi, Takeshi Abe, Fluoride Shuttle Battery Reactions of CuF2: Intermediate Phase for Defluorination, 2022, 126, 1932-7447, 12361, 10.1021/acs.jpcc.2c02772
    45. Anna Milewska, Wojciech Zając, Anita Trenczek-Zając, Janina Molenda, Cobalt-free copper-substituted Na(Ti,Mn,Ni,Cu)O2 layered oxide cathode materials for Na-ion batteries, 2022, 315, 00224596, 123461, 10.1016/j.jssc.2022.123461
    46. Yuvashri Jayamkondan, Philipp Adelhelm, Prasant Kumar Nayak, Integrated Ni and Li‐Rich Layered Oxide Cathode Materials for High Voltage Cycling in Rechargeable Li‐Ion Batteries, 2022, 9, 2196-0216, 10.1002/celc.202200786
    47. Naoaki Kuwata, Yasutaka Matsuda, Tatsunori Okawa, Gen Hasegawa, Osamu Kamishima, Junichi Kawamura, Ion dynamics of the Li Mn2O4 cathode in thin-film solid-state batteries revealed by in situ Raman spectroscopy, 2022, 380, 01672738, 115925, 10.1016/j.ssi.2022.115925
    48. Lukas Neidhart, Katja Fröhlich, Franz Winter, Marcus Jahn, Implementing Binder Gradients in Thick Water-Based NMC811 Cathodes via Multi-Layer Coating, 2023, 9, 2313-0105, 171, 10.3390/batteries9030171
    49. Jie Xiao, Feifei Shi, Tobias Glossmann, Christopher Burnett, Zhao Liu, From laboratory innovations to materials manufacturing for lithium-based batteries, 2023, 2058-7546, 10.1038/s41560-023-01221-y
    50. Yongxin Zhang, Xinghua Tan, Luting Song, Dongdong Mao, Zhengwei Fan, Sai Su, Pian Zhang, Jianping Xie, Zhaoxia Lu, Weiguo Chu, Superb High-Voltage Performance of LiCoO2 with Structure and Surface Highly Stabilized by Co-Doping Trace Mg and F during One-Pot Solid-State Synthesis, 2023, 24686069, 101313, 10.1016/j.mtener.2023.101313
    51. Xintong Li, Kai Chang, Somia M. Abbas, Rasha S. El-Tawil, Ashraf E. Abdel-Ghany, Ahmed M. Hashem, Hua Wang, Amanda L. Coughlin, Shixiong Zhang, Alain Mauger, Likun Zhu, Christian M. Julien, Silver Nanocoating of LiNi0.8Co0.1Mn0.1O2 Cathode Material for Lithium-Ion Batteries, 2023, 14, 2072-666X, 907, 10.3390/mi14050907
    52. M. Radtke, M. Heber, C. Hess, 2023, 9780124095472, 10.1016/B978-0-323-85669-0.00100-8
    53. Quoc-Nam Ha, Noto Susanto Gultom, Mikha Zefanya Silitonga, Tadele Negash Gemeda, Dong-Hau Kuo, Novel core–shell structure of Ni3S2@LiMoNiOx(OH)y nanorod arrays toward efficient high-current–density hydrogen evolution reaction, 2023, 467, 13858947, 143253, 10.1016/j.cej.2023.143253
    54. Zeyu Wu, Zhenhua Wang, Jing Zhang, Zhe Bai, Lina Zhao, Ruilong Li, Zhanfeng Yang, Yu Bai, Kening Sun, Decline Mechanism of Graphite/Lithium Metal Hybrid Anode and Its Stabilization by Inorganic-Rich Solid Electrolyte Interface, 2023, 1944-8244, 10.1021/acsami.3c05630
    55. Ardavan Makvandi, Sandra Lobe, Michael Wolff, Martin Peterlechner, Christoph Gammer, Yaser Hamedi Jouybari, Sven Uhlenbruck, Gerhard Wilde, Al-doped ZnO-coated LiCoO2 thin-film electrode: Understanding the impact of a coating layer on the degradation mechanism, 2023, 580, 03787753, 233451, 10.1016/j.jpowsour.2023.233451
    56. Julia Maibach, Josef Rizell, Aleksandar Matic, Nataliia Mozhzhukhina, Toward Operando Characterization of Interphases in Batteries, 2023, 2639-4979, 2431, 10.1021/acsmaterialslett.3c00207
    57. Jennifer Rozendo, Marco A.S. Garcia, Scarllett L.S. Lima, Nikola Tasić, Birkan Emrem, Jhonatan L. Fiorio, Guillermo Solorzano, André H.B. Dourado, Luís M. Gonçalves, Thiago R.L.C. Paixão, Jan-Ole Joswig, Anderson G.M. da Silva, Pedro Vidinha, How do Gold-Nanocrystal Surface Facets Affect Their Electrocatalytic Activities and The Benzocaine-Oxidation Mechanism?, 2023, 24680230, 103282, 10.1016/j.surfin.2023.103282
    58. Gabriela Soukupová, Martin Jindra, Tomáš Lapka, Zuzana Živcová Vlčková, Marcela Dendisová, Jan Prokeš, Otakar Frank, Fatima Hassouna, Novel silicon nanoparticles-based carbonized polypyrrole nanotube composites as anode materials for Li-ion batteries, 2024, 593, 03787753, 233976, 10.1016/j.jpowsour.2023.233976
    59. Ghizlane Elomari, Loubna Hdidou, Hicham Larhlimi, Mohamed Aqil, Mohammed Makha, Jones Alami, Mouad Dahbi, Sputtered Silicon-Coated Graphite Electrodes as High Cycling Stability and Improved Kinetics Anodes for Lithium Ion Batteries, 2024, 1944-8244, 10.1021/acsami.3c12056
    60. Guangcai Tan, Shun Wan, Jie‐Jie Chen, Han‐Qing Yu, Yan Yu, Reduced Lattice Constant in Al‐Doped LiMn2O4 Nanoparticles for Boosted Electrochemical Lithium Extraction, 2024, 0935-9648, 10.1002/adma.202310657
    61. Jitendra K. Yadav, Brajesh Tiwari, Ambesh Dixit, 2024, Chapter 3, 978-981-99-9008-5, 35, 10.1007/978-981-99-9009-2_3
    62. Ayoub El Bendali, Mohamed Aqil, Loubna Hdidou, Nabil El Halya, Karim El Ouardi, Jones Alami, Davide Boschetto, Mouad Dahbi, The Electrochemical and Structural Changes of Phosphorus-Doped TiO2 through In Situ Raman and In Situ X-Ray Diffraction Analysis, 2024, 2470-1343, 10.1021/acsomega.3c08122
    63. Manuel Grumet, Clara von Scarpatetti, Tomáš Bučko, David A. Egger, Delta Machine Learning for Predicting Dielectric Properties and Raman Spectra, 2024, 1932-7447, 10.1021/acs.jpcc.4c00886
    64. Jiannan Xu, Jieyong Tang, Ying Shi, Jianjun Xie, Fang Lei, Lingcong Fan, Lei Zhang, Investigation of ion desolvation in aqueous electrolyte with carbon electrode via electrochemical in-situ Raman, 2024, 01694332, 160551, 10.1016/j.apsusc.2024.160551
    65. Dominika A. Buchberger, Bartosz Hamankiewicz, Monika Michalska, Alicja Głaszczka, Andrzej Czerwinski, Ex Situ Raman Mapping of LiMn2O4 Electrodes Cycled in Lithium-Ion Batteries, 2024, 2470-1343, 10.1021/acsomega.4c01480
    66. Charlotte Gervillié-Mouravieff, Wurigumula Bao, Daniel A. Steingart, Ying Shirley Meng, Non-destructive characterization techniques for battery performance and life-cycle assessment, 2024, 2948-1201, 10.1038/s44287-024-00069-y
    67. Alexander A. Ryabin, Subin Choi, Yumin Heo, Sebastian Kunze, Dmitry V. Pelegov, Jongwoo Lim, Potential-driven restructuring of lithium cobalt oxide yields an enhanced active phase for the oxygen evolution reaction, 2024, 2050-7488, 10.1039/D4TA02136C
    68. M.J. Ramirez-Peral, J. Díaz-Sánchez, A. Galindo, M. Crespillo, H.P. van der Meulen, C. Morant, C. Polop, E. Vasco, Impact of the Li-loss mechanisms inherent to the physical vapor deposition of LiCoO2 cathode on its electrochemical performance, 2024, 24058297, 103658, 10.1016/j.ensm.2024.103658
    69. Quentin Jacquet, Nataliia Mozhzhukhina, Peter N.O. Gillespie, Gilles Wittmann, Lucia Perez Ramirez, Federico G. Capone, Jean‐Pascal Rueff, Stephanie Belin, Rémi Dedryvère, Lorenzo Stievano, Aleksandar Matic, Emmanuelle Suard, Nicholas B. Brookes, Alessandro Longo, Deborah Prezzi, Sandrine Lyonnard, Antonella Iadecola, A Fundamental Correlative Spectroscopic Study on Li1‐xNiO2 and NaNiO2, 2024, 1614-6832, 10.1002/aenm.202401413
    70. Simon-Johannes Kinkelin, Matthias Steimecke, Michael Bron, Laser-induced Degradation of Carbon Nanotubes During in situ-Raman Spectroscopy at High Electrochemical Potentials, 2024, 00134686, 144991, 10.1016/j.electacta.2024.144991
    71. Arvind Singh, Shamik Chakrabarti, Manjula G Nair, A K Thakur, Controlled single step synthesis, structural and electrode response of honeycomb layered Li3Co2SbO6 , 2024, 99, 0031-8949, 1059d9, 10.1088/1402-4896/ad799b
    72. Bruno Valentim, Teaching Aid Regarding the Application of Advanced Organic Petrography in Recycling End-of-Life Lithium-Ion Batteries, 2024, 10, 2313-0105, 391, 10.3390/batteries10110391
    73. Rawdah Whba, Ebru Doğan, Iqra Moeez, Ali Hussain Umar Bhatti, Muhammad Akbar, Kyung Yoon Chung, Emine Altin, Mehmet Nurullah Ates, Sebahat Altundag, Radostina Stoyanova, Sevda Sahinbay, Serdar Altin, Evaluation of the Effect of Precursor NMC622@TiO2 Core–Shell Powders Using a Prelithiated Anode from Fig Seeds: Spotlight on Li-ion Full-Cell Performance, 2024, 1944-8244, 10.1021/acsami.4c11557
    74. Yayun Xu, Weizhe Wu, Qiuying Yi, Hui Zhang, Qi Lu, Kai Zhang, Tianting Pang, Jialing Song, Dengsong Zhang, LiMn2O4 Nanoparticles In Situ Embedded in Carbon Networks for Lithium Extraction from Brine via Hybrid Capacitive Deionization, 2025, 1944-8244, 10.1021/acsami.4c16382
    75. Yun Seong Byeon, Hyo Bin Lee, Yoojin Hong, Hyun-seung Kim, Young-Jun Kim, Woosuk Cho, Min-Sik Park, Cation-Deficient LixWO3 Surface Coating on Ni-Rich Cathodes Materials for Lithium-Ion Batteries, 2025, 1944-8244, 10.1021/acsami.4c18935
    76. Soumyadip Mitra, Thilini Rathnayaka Mudiyanselage, Xijue Wang, Gaurav Lohar, Deepak Dubal, Chandran Sudakar, Bolstering the Rate Performance of Co‐Free Ni‐Rich Layered Oxide Cathode through a Rapid Heating Method, 2025, 2566-6223, 10.1002/batt.202400782
    77. Ziyi Wang, Yulai Wei, Pinyi Wang, Ruimin Song, Sida Zhang, Changding Wang, Zhiwei Shen, Weigen Chen, Spectroscopic insights into the thermal aging process of lithium-ion battery electrolytes, 2025, 334, 13861425, 125926, 10.1016/j.saa.2025.125926
    78. Gulzat Nuroldayeva, Tanay Umurzak, Aziza Kireyeva, Assylzat Aishova, Orynbassar Mukhan, Sung-Soo Kim, Zhumabay Bakenov, Nurzhan Umirov, A comparative study of bulk and surface W-doped high-Ni cathode materials for lithium-ion batteries, 2025, 2040-3364, 10.1039/D4NR04691A
    79. Alessandra Zanoletti, Antonella Cornelio, Laura Borgese, Giacomo Siviero, Amedeo Cinosi, Elisa Galli, Elza Bontempi, Sample preparation procedures for elemental analysis of critical raw materials in lithium-ion battery black mass: Challenges responding to the supplementary battery recycling regulation, 2025, 380, 03014797, 124973, 10.1016/j.jenvman.2025.124973
    80. Ngoc Tram Phung, Yue Feng, Théodore M. Poupardin, Bon Min Koo, Catherine Henry de Villeneuve, Michel Rosso, François Ozanam, Infrared Operando Study of the Solid-Electrolyte Interphase of Amorphous-Si Electrodes for Li-Ion Batteries: Effects of Methylation and Boron Doping, 2025, 2574-0962, 10.1021/acsaem.4c03245
    81. Itumeleng Seotsanyana‐Mokhosi, Kenneth I. Ozoemena, Enhancing the Electrochemical Performance of Nickel‐Rich Layered Cathode Material (NMC 622) for Lithium‐Ion Batteries by Ultrathin Conductive Ceria‐Surface Coating, 2025, 1434-1948, 10.1002/ejic.202400689
    82. Hakwoo Lee, Jihyun Kim, Suwon Lee, Maxim Avdeev, Wanli Yang, Yong‐Mook Kang, Characterizing Disorders Within Cathode Materials of Lithium‐Ion Batteries, 2025, 1433-7851, 10.1002/anie.202501958
    83. Hakwoo Lee, Jihyun Kim, Suwon Lee, Maxim Avdeev, Wanli Yang, Yong‐Mook Kang, Characterizing Disorders Within Cathode Materials of Lithium‐Ion Batteries, 2025, 0044-8249, 10.1002/ange.202501958
  • Reader Comments
  • © 2018 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(18516) PDF downloads(5230) Cited by(83)

Article outline

Figures and Tables

Figures(20)  /  Tables(1)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog