
Mei Yang, Ruyi Bi, Jiangyan Wang, Ranbo Yu, and Dan Wang, Decoding lithium batteries through advanced in situcharacterization techniques, Int. J. Miner. Metall. Mater., 29(2022), No. 5, pp.965-989. https://dx.doi.org/10.1007/s12613-022-2461-0 |
The ever-growing energy demands of society require the development of more efficient and economical energy storage systems [1–3]. Rechargeable lithium batteries (LBs), given their high theoretical energy density and environmental friendliness, have drawn intense research interest and dominated the market of portable electronic devices. Although LBs have also been adopted for grid-scale energy storage and electric vehicles, their relatively low capacity and high fabrication cost compared with solar, wind, or fossil energy sources pose barriers to their market adoption. The property and performance of LBs are mainly determined by their anode and cathode components. Thus, high-capacity and cost-effective electrode materials must be designed and explored to promote LBs into a wider consumer market.
Various electrode materials, such as Si, Sn, and Li metal for anodes and Ni-rich oxide and S for cathodes, have been explored for LBs. However, despite their higher theoretical capacity than conventional electrode materials, all these emerging materials suffer from grand challenges. For example, although Si possesses more than ten times the theoretical capacity of graphite anodes, its huge volume expansion (~300%) during lithiation causes structural collapse, loss of electrical contact, and accumulated solid electrolyte interface (SEI) formation, which induce a practically low achievable capacity and poor cycling life [4]. Li metal, which possesses the highest theoretical capacity and lowest electrochemical potential, has been considered as the best anode of choice for LBs. However, Li tends to deposit in dendritic form, which causes capacity decay and brings severe safety concerns to batteries [5]. As for S cathodes, the polysulfide dissolution severely harms its cycling life and impedes the commercialization of lithium–sulfur (Li–S) batteries [6–7]. In addition, the recovery and recycling of valuable metal elements in exhausted LBs attract great attention due to resource consumption and cost increase [8–10].
To improve the performance of batteries, scholars should improve their understanding of the chemical and electrochemical reaction processes and the internal degradation mechanisms in the whole battery system from different scales, including the electrode materials, electrodes, and devices. Given such a goal, various material characterization and electrochemical analysis techniques have been developed. Compared with characterization under ex situ conditions, in situ and operando characterization not only can provide the information the same as that from ex situ characterization but can also capture information related to the dynamics of charge/discharge cycles [11].
In the past years, in situ and operando characterization tools have been extensively developed for the monitoring of material structure and surface/interface chemistry evolution under operating conditions. For example, in situ transmission electron microscopy (TEM) is a powerful tool used to visualize morphological and structural evolutions and volume expansion/contraction [12]. In situ X-ray absorption spectroscopy (XAS) shows a unique capability to disclose the atomic and electronic structure of the active host metal in action [13]. In situ atomic force microscopy (AFM) provides direct information on the surface topography; it is especially useful in detecting the SEI layer evolution [14]. In situ Raman spectroscopy is used to analyze new products as the transient states of an electrode surface [15]. Other in situ tools, such as scanning electron microscopy (SEM) [16], X-ray diffraction (XRD) [17], Fourier transform infrared (FTIR) spectroscopy [18], mass spectrometry (MS) [19], nuclear magnetic resonance (NMR) spectroscopy [20], and neutron reflectometry/neutron depth profiling (NDP) [21], have also been developed to study battery failure during battery operation. With the assistance of these techniques, a comprehensive understanding of how the structure and composition of battery components affect their chemical/electrochemical behaviors during battery operation can be obtained. This information can guide people in the optimization and design of suitable electrode and electrolyte materials, thus improving the overall performance of batteries.
In this review, we summarize the recent progress in advanced in situ characterization techniques for LBs, especially focusing on the multi-model coupling of different characterization techniques. We first briefly discuss the advancement of single characterization techniques and the combination of several techniques for the LB system. Then, the characterizations of different battery components, including the cathode, anode, electrode–electrolyte interface, and electrolyte, are discussed. Case studies are used to illustrate how these advanced characterizations are used to study the composition and structural evolution of typical battery systems during cycling and reveal the structure–performance correlation, ranging from the local structure and functional groups at the molecular level to the morphological evolution at the electrode scale and overall electrochemical performance on the device scale. Finally, along with the concluding remarks and remaining challenges, opportunities and possible directions for future development of in situ characterization techniques that will further benefit LBs are discussed.
In the past years, the great progress in in situ characterization techniques facilitated the comprehensive understanding of the reaction mechanism in LBs. The research has greatly benefited the optimization and design of more efficient battery electrodes and electrolyte materials and contributed to the further improvement of the overall battery performance [22]. Fig. 1 summarizes the important and widely applied in situ techniques with their spatial resolution scales and the corresponding detection objectives in batteries. Depending on the relevant length scale of each unit, these techniques can be selectively used to disclose the structure, morphology, chemistry, and kinetics of batteries. Moreover, with the joint help of different characterization techniques, comprehensive studies have been carried out to understand the complicated issues that arise during cycling with different physical/chemical aspects and at multiple length scales. In section 2.1 and section 2.2, we discuss the progress of in situ characterization using single and joint multi-modal techniques, respectively. In section 2.3, we discuss the design of electrochemical cells for in situ characterization.
In this part, the progress of typical in situ characterization technologies for LBs based on X-ray, electron, neutron, optical, magnetism, and scanning probe microscopy (SPM) is discussed.
X-ray techniques offer quantitative and qualitative information on the structure and chemical composition of materials [13–17]. In situ XRD is used to investigate the phase constitution, space group symmetry, lattice parameters, and space-averaged atomic structure during cycling. XAS can provide fine insights into the structural evolution at the atomic level, detecting the electronic transition, valence state changes, sites symmetries, bond length, coordination number, degree of disordering, and chemical identity of the nearest atom. TXM and XRT offer direct visualization of the phase transformation, crack creation and growth, and electrode deformation, which are associated with the degradation of battery performance.
Compared with X-rays, neutron-based techniques, typically including NPD, have unique advantages, such as the complementary scattering cross-sections of neutrons, high sample penetration, and sensitivity to low atomic number elements (e.g., H, Li, and O) [23]. NPD, similar to XRD, can precisely determine structural and dynamic information at the atomic level, but it is difficult to use in amorphous materials. NDP has a nanoscale spatial resolution, which is very sensitive to the Li-ion concentration changes with respect to the depth of battery materials. On this basis, other neutron-based techniques, for example, NPDF, NR, and neutron imaging (NI) have been developed and applied to the in situ analysis of electrodes in LBs [21].
In comparison with X-ray and neutron, an electron beam has a shorter de Broglie wavelength and shallower penetration depth. Therefore, electron-based techniques are theoretically capable of obtaining higher resolution [12,16]. TEM can provide a combination of nanoscale down to atomic level spatial resolution. EELS can examine the oxidation state related to the reaction chemistry. Energy dispersive X-ray analysis (EDX) provides information about the elemental distribution and quantitative concentration. The combination of these techniques, in situ TEM turns into a powerful analytical tool to reveal the electrochemical mechanisms. By contrast, the resolution of SEM is about 10 nm, which is usually employed to probe the morphological change.
Optical techniques, which have non-destructive, non-contact, and non-vacuum requirements, are also widely adopted for in situ characterization of LBs. In situ Raman spectroscopy can be applied to monitor the valence and bond variation in the local structure and oxidation state at the atomic scale. With the adoption of the mapping mode, Raman images provide distinct phase-transformed zones accompanying electrochemical reactions [15,24]. FTIR can identify organic chemical species, including ion molecules, solvents, dipolar species, and some chemical breakdown products [18,25]. Optical microscopy (OM), which possesses a low resolution of about 200 nm, is often used to visualize the macrostructural features of batteries [26–27].
Techniques based on magnetism include NMR, EPR, and Mössbauer spectroscopy. NMR helps in monitoring the local electronic environment around the nucleus over the electrode, electrolyte, and their interfaces [20]. MRI can provide spatial information in an opaque cell body [28]. EPR can track the initialization of microstructural growth and the reaction intermediates of electrolyte breakdown and lithium metal without any spectroscopic interference in the signals [29]. MS enables one to study the influence of the electronic environment on nuclear hyperfine energy levels and probe the structure of materials [30].
The main principle of SPM is based on the detection of physical parameters produced by the distance change between the sample and the tip of a fine scanning probe. In situ AFM is available to study the interface of batteries, such as SEI morphology, Li dendrite growth, and mechanical and electrical properties. New analytical tools are developed by modifications of the scanning probe. For example, a periodic high-frequency voltage bias is applied between the cathode and anode, which results in oscillatory surface displacement. With the use of an electrically active AFM probe, ESM can capture oscillatory signals to study the lithium ionic motion at the nanometer scale [31]. Electrolyte-filled nanocapillaries are adopted as the probe for SICM, which can provide an ionic current map with topography [32]. Table 1 summarizes the current main in situ characterization techniques, including their spatial resolution, applicable scenarios and information collected, and the current challenge of each technique.
In situ technique | Spatial resolution | Applicable scenes and obtained information | Facing challenge | |
X-ray technique | XRD | ~μm scale | Quantitatively and qualitatively provide composition, distribution, and conversion of crystallographic structure analyses; study local environment of an element of electrodes. | Amorphous and low atomic number elements in LIBs. |
XAS | 1–100 nm | |||
TXM, XRT | ≥15 nm | |||
XPS | ~μm scale | |||
Neutron-based technique | NPD | High penetration and sensitive to low atomic number elements. Precisely determine structural and dynamic information of electrodes. | Amorphous and high atomic number elements in LIBs. | |
NR | nm | |||
NI | ~μm scale | |||
NDP | nm | |||
Electron-based technique | TEM | ≥0.1 nm | Provide morphological information of electrodes and SEI; obtain structural information by coupling accessory. | Shallow penetration; high vacuum; high-energy electron damage. |
SEM | 10 nm | |||
EELS | ||||
Optical technique | Raman spectroscopy | 10 nm–1 μm | Probe the molecule structural features and the chemical composition of electrodes, SEI, and electrolytes. Visualization of macrostructural features. | Specific active component; signal intensity and background interference; bulky or depth analysis. |
FTIR spectroscopy | ~μm scale | |||
OM | ~μm scale | |||
Scanning probe technique | AFM | nm | Monitor morphology and changes of surface/interface in diverse environments, SEI, Li dendrite, etc., as well as mechanical properties and electrical properties. | Low resolution |
ESM | 10 nm–1μm | |||
SICM | 10 nm–50 nm | |||
Magnetic resonance technique | NMR | mm | Apply to an odd atomic mass or odd atomic number element of LIBs; monitor the local electronic environment around the nucleus through electrodes, electrolytes, and their interfaces. | Specific elements; low resolution; |
EPR | ||||
Mössbauer spectroscopy |
Each technique has its advantages and disadvantages, leading to different application scenarios. Hence, different in situ characterization techniques must be combined to conduct comprehensive studies on the understanding of different physical/chemical aspects at multiple length scales. Luo et al. [33] studied the charge compensation in 3d-transition-metal (TM) oxide intercalation cathodes by combining operando MS, soft XAS (SXAS), resonant inelastic X-ray scattering spectroscopy (RIXS), XANES spectroscopy, and Raman spectroscopy (Fig. 2). According to the 18O-labeled MS results, oxygen was extracted from the lattice of the charging cathodes and not from the electrolyte, which almost entirely generated the capacity to extract charges above 4.5 V. SXAS, RIXS, and XANES spectra were used for the joint investigation of the formation of electron-hole states on oxygen and the dominant mechanism of charge compensation, indicating that localized electron holes formed on the O ions coordinated by Mn4+ and Li+, which promoted the localization and not the formation of true
Given that the phenomena pertinent to the functioning of batteries occur over wide spatial and temporal ranges, coupling different in situ characterization technologies with simultaneous measurements at the same spot is favorable to eliminating the information mismatch of time lag. Wang et al. [34] developed an operando TXM–XAS approach to study the dynamic phase transformation evolution and intercalation pathway within multi-particle and individual LiFePO4 particles during battery operation (Fig. 3(a)–(d)). Therein, TXM images provided direct 2D mapping of the inhomogeneous phase spatial distribution, whereas XANES spectra indicated the composition of each region, corresponding to nearly pure LiFePO4 and FePO4 phases, respectively. The single-particle system presented direct evidence of two-phase coexistence for the first time and suggested the two-phase dynamic models of individual LiFePO4 particles by in operando imaging work. Pérez-Villar et al. [35] exploited an in situ microscopic approach combining Raman spectroscopy and FTIR micro-spectroscopy to investigate the interfacial reactions and spectral characterization of the local structure at the same spot on the surface of electrodes (Fig. 3(e)–(f)). Through micro-FTIR spectroscopy, solvation–desolvation interfacial phenomena were detected, whereas the carbonaceous material was characterized by micro-Raman spectroscopy at the same spot on the glassy carbon surface. This research demonstrated the reliability of the combined microscopy technique and its suitability in obtaining complementary information about both the surface species involved in interfacial reactions. Tsuchiya et al. [36] designed a combined ion-beam analysis of high-energy elastic recoil detection (ERD) and Rutherford backscattering spectrometry (RBS) with 9.0 MeV tetravalent oxygen ion (O4+) probe beams to study the distribution of Li and its movement around the interface. These typed coupling techniques are based on the same modal or excitation beam, e.g., X-ray, laser, and ion-beam. Given their similar demand and conditions, they are easy to implement by modification of specific devices and cells. Nevertheless, congeneric information alone can be obtained owing to the limitation of the same excitation source.
Techniques should also be coupled based on different modalities [37–39]. Schmidt et al. [38] developed a novel correlative microscopy device by combining Raman microscopy, SEM, and EDX. As shown in Fig. 4(b), the laser beam for the confocal Raman microscope was linked to the SEM chamber through an optical window. SEM provides high-resolution images of surface topography, EDX offers elemental component information, and Raman microscopy delivers molecular bond information, leading to a comprehensive result. Zhang et al. [39] constructed a novel AFM with an environmental TEM (AFM–ETEM) setup to observe in situ growth and measure the stress changes of individual Li whiskers (Fig. 4(c)). In the in situ electro-chemo-mechanical experiment, an arc-discharged carbon nanotube (CNT) was attached to a conducting AFM tip by electron-beam deposition of carbonaceous materials; this assembly was used as the cathode, and the scratched Li metal on the top of a sharp tungsten needle was used as the anode; Li2CO3 on the Li surface was used as the solid electrolyte. The morphology and growth process of the whisker can be directly recorded by ETEM, whereas the axial compressive stress can be detected by the AFM tip, corresponding to the critical stress and applied voltage. A sub-micrometer whisker grew at overpotentials at room temperature, generating a growth stress of up to 130 MPa.
To study the lithium diffusion behaviors in LB systems, our group successfully designed and developed a multi-model in situ characterization system coupling NMR technique and Raman spectrum (Fig. 5). The laser can enter into the magnetic field and reach the targeted materials across the optical window through an optical fiber and optical extension pipe, whereas the Raman signals can be collected by the objective lens and reach the spectrograph through the optical fiber. The interference can be avoided by controlling the technical parameters of the instrument and selecting the proper ingredients. Using our system, the Raman spectra of electrodes, NMR signals, or MRI of 1H or 7Li can be obtained simultaneously during the discharge/charge cycling. Raman spectra can reveal structural conversion at the molecular level and deduce electrochemical reaction and kinetics process; NMR data can provide the local environment information of hydrogen and lithium element, whereas MRI can reveal the morphological information about electrodes, electrolytes, and their interfaces. Owing to synchronous data acquisition, the difference in temporal resolutions between NMR and Raman techniques is reduced, which favors the accurate and comprehensive understanding of the battery failure mechanism.
The multi-model coupling of characterization techniques, which has a deep significance, proceeds beyond the simple addition of different methods. In practice, different characterizations have varying spatial resolutions from nm to the µm scale while also reaching the mm and cm levels and temporal resolutions from several seconds to minutes, hours, or days. To eliminate the influence of acquisition time, scientists artificially set the testing conditions and parameters to retain the target in equilibrium or stable state. Given that the system of batteries is complex, dynamic, and interrelated, the integration of the information collected from each test is a huge challenge. The concept of “in situ” in multi-model coupling needs to be defined and distinguished and requires the consistency of spatial and temporal scales. The characterization sites or zones are spatially unchanged in a complex system, depending on the spatial resolutions of the characterization technique. The applicable objectives are stable during the acquisition period, relying on temporal resolutions. The multi-model coupling has the following features. First, it is favorable to the simultaneous collection of different information, compensating for the applicable objectives limited to a single technology. Then, the results obtained can cover a broad scale range from the atomic to microcosmic, mesoscale, and macroscopic scales, building up the spatial link for whole systems. More importantly, the multi-model coupling can trace the real and complete dynamic change of a process, eliminating the error of temporal mismatching one at a time. The multi-model coupling will be the growing trend of in situ techniques in the future, but it is still in its infancy. For example, temporal resolutions are not only related to the parameters of each instrument but also the specific materials and process of charge/discharge cycling, which is more complex than spatial-temporal. Finding solutions to the temporal resolutions of different characterization techniques is a huge challenge. Hence, great effort needs to be exerted to address this bottleneck of technology.
To conduct in situ measurements successfully, scholars should carefully optimize the in situ electrochemical cells to properly fit the specific analytical techniques. As shown in Fig. 6, some typical setups and schematics of in situ cells have been reported for different types of in situ experiments [25,40–46]. Some basic principles are followed for their design [47]. First, the design of the in situ cell needs to have an environment similar to that in a commercial or a coin cell. For example, an open-type cell without any covering is adopted for in situ TEM measurement. The excited radial wave can reach the sample region and return to the detector across the packages by various transparent windows. Polyethylene, beryllium, and Kapton film are used as X-ray windows, whereas glass, quartz, sapphire, and CaF2 are suitable for the in situ Raman test. Meanwhile, metallic parts should be used as sparingly as possible in NMR. Third, the interference from inactive cell components to outgoing detected signals should be at the minimum to truly collect information under in situ operating conditions. The interaction of the particle beam with cell components often generates background signals, which need to be precisely removed afterward from the total signals. Importantly, radiation damages and side effects on the material should be reduced, such that the magnetic field orientation in NMR has a direct influence on the paramagnetic cathode materials, and the heat produced by laser during in situ Raman test may lead to the decomposition of electrolytes. Finally, the in situ setup should be easy to assemble/disassemble and highly reproducible.
Cathode materials are one of the most important factors that determine the performance of LBs. Thus, the geometric structure, phase structure, and composition evolution of cathode materials must be traced during battery operation through in situ characterization.
Lithium-rich cathode materials, such as lithium-rich manganese-based cathode materials (xLi2MnO3·(1−x)LiTMO2, TM = Ni, Mn, Co, etc.), have extremely high theoretical specific capacities (>350 mAh·g−1) and reversible specific capacities (>250 mAh·g−1) and are considered to be the most promising next-generation Li-ion battery (LIB) cathode materials [48]. However, Li-rich Mn-based cathode materials also face problems, such as the low initial coulombic efficiency, poor multiplicity performance, and severe capacity and voltage decay, which hinder their commercialization. In situ characterization means can help in revealing some internal mechanisms, which can provide new insights and ideas for the future development of Li-rich cathode materials.
Currently, the mechanisms involved in the activation of high-energy LiNixCoyMn(1−x−y)O2 (NCM) remain intensely debated. In this context, Lanz et al. [49] compared the changes in the Raman spectra of stoichiometric and high-energy NCM during electrochemical cycling under the same conditions for the first time. Ex situ Raman measurements of NCM with increased overlithiation revealed a shift of the A1g band toward the Ag band of Li2MnO3, whereas no such movement was observed after electrochemical cycling, thus lending support to the Li2MnO3 domain model and the irreversibility of Li2MnO3 activation, respectively (Fig. 7(a)). Finally, the in situ Raman spectra of stoichiometric and high-energy NCM showed a new reversible band at ~545 cm−1, which was stable over a larger potential range in the latter compound.
Triggering oxygen redox activity is considered a promising strategy to increase the capacity of cathode materials for LBs. However, the irreversible loss of lattice oxygen exacerbates the structural distortion of the spinel phase, leading to severe voltage and capacity decay. Cao et al. [50] evaluated the reversibility of the structural evolution of Li-excess O2-type layered oxide cathode, Li0.66[Li0.12Ni0.15Mn0.73]O2 (LLNMO) in the first two Li+ (de)intercalation cycles through systematic in situ XRD. The in situ XRD results revealed that the evolution of the peak position at the second charge showed an opposite trend relative to that of the initial discharge. In addition, the layered structure and O2-type oxygen stacking can be maintained after 50 cycles (Fig. 7(b)). The two pseudo-plateaus located at 3.67 and 2.73 V were well preserved, and the voltage decay can be effectively suppressed during long cycling.
Moreover, lithium diffusion is one of the key criteria for achieving high-magnification and high-power batteries. Zhou et al. [51] demonstrated how in situ NMR methods can be adapted to probe the mobility of lithium and the nature of electrode phase transitions in real time during cell cycling (Fig.7(c)). The results showed that the center of mass of the broad static resonance was close to the center of mass of the isotropic resonance observed during rotation. The resonance at about chemical shift (δ) = 250 ppm originated from the lithium metal negative electrode. The spinel resonance shifted to high frequencies and decreased in intensity as lithium was removed from the tetrahedral sites, and Mn3+ was oxidized to Mn4+ during the charging process. The electrochemical trend and shift of the spinel peak position were reversed during discharge.
Nickel-rich ternary materials possessing a nickel content generally greater than 50wt% are one of the cathode materials used for high-energy-density LIBs. Given the high nickel content, nickel-rich ternary cathode materials have a high reversible discharge capacity. However, numerous problems also occur, and they include unstable surface properties, structural defects and lithium–nickel mixing, intergranular cracks, and microstrain, which lead to poor cycle performance.
To solve the above problems, researchers have carried out a number of studies. Seong et al. [52] suggested a simple in situ method to control the residual lithium chemistry of a high nickel lithium layered oxide, Li(Ni0.91Co0.06Mn0.03)O2 (NCM9163), with minimal side effects. Further surface-specific analyses were performed using XPS depth analysis and time-of-flight secondary-ion MS to determine the chemical properties of SO2-treated NCM9163 (Fig. 8(a)). As observed in the XPS curves, the sulfur signal was not detected in the SO2-treated NCM9163 until 10 min after milling but not in the original NCM9163. They observed that if the preferred reaction is the formation of a low alkalinity lithium compound, then it mitigates the formation of the predominantly basic LiOH or Li2CO3.
Otherwise, the dissolution of TMs from the cathode active material and deposition on the anode will cause significant cell aging. Jung et al. [53] investigated the dissolution of TMs from an Li1−xNi0.6Mn0.2Co0.2O2 (NMC622) cathode via in situ and ex situ XRD (Fig. 8(b)). The analysis showed that the aging mechanism for all three metals was the loss of recyclable lithium in graphite SEI. This loss was greater when Mn was present in the electrolyte compared with Ni and Co due to the higher activity of the deposited Mn toward SEI decomposition compared with Ni and Co.
For nickel-rich cathode materials, the role of crystal defects on their structural evolution and subsequent property degradation during electrochemical cycling is unclear. Li et al. [54] studied the structural evolution of Ni-rich LiNi1−x−yMnxCoyO2 (NMC) cathodes using Li0.5La0.5TiO3 (LLTO) as solid-state electrolyte via in situ TEM (Fig. 8(c)). They identified that the antiphase boundary (APB) and twin boundary (TB) separating layered phases played an essential role during phase change. When lithium was depleted, the APB extended across the laminar structure, and the mixing of Li/TM ions in the lamellar phase induced the formation of rock-salt phases along the coherent TB.
With the advantages of high specific energy, low cost, and abundant resources, Li–S batteries have received widespread attention from the research community. On the one hand, compared with current commercial cathode materials, monolithic sulfur provides 8–10 times higher theoretical specific capacity. On the other hand, monomass sulfur is abundantly stored and environmentally friendly. However, numerous unresolved issues exist with Li–S batteries. First, the insulating nature of S and Li2S leads to poor kinetics of the conversion process and low actual capacity utilization. Second, as an intermediate, lithium polysulfides (LiPSs) have a significant “shuttle effect” on most organic liquid electrolytes, where soluble sulfur shuttles between the cathode and anode, deteriorating the reaction interface. In addition, the severe volume change (80%) leads to the degradation of the mechanical properties of the sulfur cathode [55]. The current research on Li–S batteries focuses on suppressing the polysulfide shuttle effect, preventing lithium dendrite growth, and improving the conversion reaction kinetics.
The immobilization of sulfur and LiPSs in various host materials is a common strategy used to overcome some of the problems of the sulfur cathode. However, a phenomenon in which some hosts themselves can be lithiated−delithiated occurs during the cycling process. Liu et al. [56] proposed the concept of dynamic hosts for Li–S batteries for the first time and elucidated the mechanism through which TiS2 acts in such a fashion via in situ XRD (Fig. 9(a)). TiS2 was completely converted to LiTiS2 during the initial discharge process. Upon charging, LiTiS2 was converted to LixTiS2 (0 < x < 1) and disappeared. During the next charging process, the diffraction peak of Li2S also disappeared. Similar behavior was observed in the subsequent cycles.
Another major problem with Li–S batteries is that the mechanistic understanding of the reaction behind the polysulfide shuttle is still incomplete. Conder et al. [57] identified the characteristics of polysulfides adsorbed on the surface of glass fiber diaphragms and monitored their evolution during cycling via operando XRD (Fig. 9(b)). The results showed that only the long-chain polysulfides were adsorbed on the glass fiber surface, and they remained on it during the repeated lithiation/delithiation processes. These findings not only demonstrated that XRD can be used to study liquid–liquid interactions in Li–S batteries but also showed that fumed silica is a promising electrolyte additive that can significantly improve the performance of Li–S batteries.
Sulfur reduction is a stepwise electrochemical process with a series of highly soluble polysulfides as intermediate species. Wu et al. [58] examined the discharge and charge processes of a Li−S cathode in 1 M lithium bis(trifluoromethane sulfonyl)imide (LiTFSI) and tetraethylene glycol dimethyl ether (TEGDME)/1,3-dioxolane (DIOX) (1/1, volume ratio) via in situ Raman spectroscopy (Fig. 9(c)). Raman spectroscopy showed that
Anode materials are another important factor affecting the energy density and cycle life of batteries. Based on the mechanism of lithium-ion storage, anode materials can be divided into intercalation, conversion, and alloying-type anodes [60]. However, the dynamic evolution of the microstructure, morphology, phase, and chemical composition of anode materials is not very well understood. Therefore, the chemical, material, physical, and mechanical information on LIBs can be obtained by using advanced in situ characterization techniques, which are conducive to further material design and performance improvement [61].
Graphite, which is a classical intercalation anode material, is the most commonly used commercial anode material [62]. A series of phase transitions and volume changes is caused by ion intercalation in graphite, and numerous controversial issues about its intercalation mechanism remain. Thus, in situ characterization is beneficial to understanding the working mechanism of graphite [63].
Tao et al. [16] studied the anisotropic deformation and dynamic microstructural evolution of graphite-based commercial electrodes using a combination of in situ SEM digital image correlation. The electrode exhibited an irreversible swelling of 50% in the initial cycle and a reversible anisotropic expansion and contraction deformation process in the subsequent cycle (Fig. 10(a)). The vertical swelling deformation was evident, and the electrode–separator interface was the main area of lithiation and delithiation. In addition, the diffusion path of lithium-ion was obtained based on the evolution of the strain field.
Schweidler et al. [64] measured the phase diagram of LixC6 by in situ XRD and gave the single-phase and coexisting phase regions. The study revealed that when C6 was completely lithiated to LiC6, the total volume expansion was 13.2%, of which about 5.9% occurred in the early dilution stage. The remaining expansion was due to the transition from phase 2 to phase 1. As a result, lithium was thought to intercalate indiscriminately in graphite, but it intercalated only in one layer in every two layers, leaving all graphene in the same chemical state (Fig. 10(b)). This study provided valuable information indicating that graphite undergoes considerable crystalline strain under partial lithiation.
Li et al. [65] fabricated ZrNb14O37 nanowires (NWs) by electrostatic spinning, and they were used as anodes for LIBs. The basic reaction mechanism of lithium/delihiation was studied by in situ XRD. Fig. 10(c) shows the in situ XRD system used for the study; it consisted of an XRD device and a LAND CT2001A battery tester. The discharge process can be divided into three different regions based on the behavior of the main peak of the in situ XRD pattern. State I was the solid solution reaction region, in which 5.25 Li+ insertion occurred in each molecular formula unit of ZrNb14O37 to form the intermediate Li5.25ZrNb14O37. State II was a two-phase reaction in which an additional 2.87 Li+ insertion into the octahedral frame occurred, indicating the transition from Li5.25ZrNb14O37 to Li8.12ZrNb14O37. State III was the second solution reaction, in which 12.84 Li+ insertion into each formula unit occurred to form the Li20.96ZrNb14O37 phase. When the battery was fully charged to 3 V, the characteristic peak almost returned to the original position, indicating that its structure had high stability and reversibility.
Pang et al. [66] used in situ NPD to study the transference of Li+ in lithium titanate (LTO) anodes with different particle sizes during the battery operation (Fig. 10(d)). Li+ migrated from 8a site to 16c site via 32e site during the charging process, accompanied by the rearrangement of oxygen positions in the structure. The increase in the number of Li+ migrations was due to the shortening of the path length at the smaller particle size anode rather than a change in the Li+ migration path. Li replenishment at position 8a was faster than Li transfer to position 16c. Compared with LTO-1, which had a large particle size, LTO-2 had a smaller particle size and contained a larger proportion of lithium sites.
Compared with intercalation anode materials, conversion anode materials, mainly including TM oxides, sulfides, nitrides, and phosphides, have a high theoretical specific capacity.
He et al. [67] employed a strain-sensitive, bright-field STEM to study the physical and chemical mechanism of nano-Fe3O4 in real time. Explicit visualization of the two-step intercalation process of lithiation in Fe3O4 nanocrystals revealed that the initial lithium intercalation formed as two phases of LiFe3O4 (Fig. 11(a)). Further lithiation led to a conversion reaction, which resulted in the coexistence of three different phases in a single nanoparticle (NP).
He et al. [68] first synthesized a 1D pure β-MnO2 phase with a 1×1 tunnel structure and used it as a cathode for a series of in situ measurements to reveal its energy storage mechanism. The bulk phase changes of β-MnO2 crystal structure during the cyclic process were studied by in situ XRD. The instability of the frame structure was caused by the incomplete reversible transition of β-MnO2 to the orthogonal LiMnO2 phase during cyclic lithiation, which led to capacity decay (Fig. 11(b)). In addition, in situ TEM analysis was performed on the lithium storage behavior of β-MnO2 nanorods to understand their microstructural changes at the microscale. The findings of this study can provide better guidance for the structural regulation of such materials to further optimize lithium storage properties.
At present, the in situ characterization techniques for tensile mechanical properties and fracture mechanisms of electrode materials are very limited. Song et al. [69] quantitatively studied the tensile properties of electrochemically modified SnO2 NWs by in situ SEM. The results showed that the transformation of crystal to glass during the lithiation process resulted in large plastic deformation (Fig. 11(c)). The lithiation and delithiation processes, on the other hand, reduced the mechanical properties of SnO2 NWs. Notably, the mechanical properties of delithiation SnO2 NWs were generally higher than those of lithiation. This study promotes the understanding of the mechanical properties of SnO2-based nanomaterials in LIBs.
Kitada et al. [70] observed the electrochemical reaction process of silicon-based materials by in situ NMR. By comparing the electrochemical processes of amorphous lithium silicide, amorphous SiO, and disproportionated silicon monoxide obtained by heat treatment at different temperatures, they pointed out that amorphous silicon oxide has better cyclic properties than crystalline silicon (Fig. 11(d)). This reaction did not follow the typical two-phase reaction but was a gradual solid solution process, thus avoiding the electrode deterioration caused by volume mutation.
Si is a typical representative of alloying-type anode materials, and it has a high theoretical specific capacity (4200 mAh·g−1) and reserve abundance in the earth’s crust; thus, it is considered an ideal choice for the next generation of LIB anode materials [71]. However, the large volume change during the process of lithium intercalation is one of the main reasons leading to poor cycle performance. Given the limited spatial resolution of characterization methods, the microscopic mechanism of improving the performance of Si anode is still unclear.
Zhao et al. [72] developed a simple and controllable methodto embed multiple Si NPs as cores into hollow multishelled structures (HoMS), and in situ TEM technology was used to study the dynamic morphological evolution of Si NPs during lithiation (Fig. 12(a)). The overall structure of the electrode remains intact during Li+ repeated insertion/extraction process. This demonstrated that the inner shell can effectively limit the direction of core expansion, whereas the outer shell maintained a stable interface and overall structural integrity. Therefore, the result showed the excellent electrochemical performance of the electrode material in LIBs. In addition, in situ Raman spectroscopy was used to detect the electrochemical changes in Si NP@2S-CoFe2O4 HoMS at different potentials. The peak intensity of the Si–Si bond (516.2 cm−1) decreased with the development of lithiation. When the discharge reached 0.01 V, the peak disappeared completely, indicating the formation of Si–Li alloy.
Liu et al. [73] investigated the formation of lithium ions in Sn electrodes by in situ NDP and monitored the transfer of lithium ions in the electrode (Fig. 12(b)). In situ NDP enables researchers to understand the transport rate and distribution of Li+, thus guiding the development of effective storage mechanisms for materials. It complements other material characterization methods and helps us realize the complex interrelationships among electrochemistry, reaction dynamics, intercalation, and migration. This technique is not limited to materials formed between metals but can also be used to study a wide range of anode and cathode materials.
Li et al. [74] studied the dynamic morphology and phase change of selenium-doped germanium at the single-particle level by synchrotron-radiation in situ TXM. The experimental results showed that the micron-sized Ge0.9Se0.1 particles had good reversibility at a 1 C rate, which promotes the learning of their better cycling performance and admirable rate capacity (Fig. 12(c)). The in situ operation technology based on single-particle batteries not only avoids the influence of other components in the coin battery on the active material and the damage induced by X-ray but also provides a reference for understanding the dynamic characteristics of battery materials during the particle-level charge and discharge process.
Lithium metal is considered to be an ideal material for high-energy-density secondary batteries owing to its low potential platform (−3.04 V vs. standard hydrogen electrode) and high theoretical capacity (3860 mAh·g−1) [75]. However, lithium dendrite growth and consequent serious safety problems seriously limit its commercial applications.
Li et al. [76] designed and formed a microscopic-ordered 3D electrode structure on the surface of a garnet electrolyte by electron-beam evaporation and ion-beam etching method. In situ NDP technique with high spatial resolution and detection sensitivity was used to quantify lithium deposition. Its detection principle is shown in Fig. 13(a). The particles emitted by different depth paths had different energy losses and carried various energies. Based on this condition, the depth distribution information of lithium in the Z direction can be determined to distinguish the deposition distribution of lithium under the hole and Ti film. The voltage and current information of the battery were collected via in situ charge and discharge, and the energy information of outgoing particles was collected synchronously. The discharge capacity and total integral intensity were consistent, indicating the sensitivity and reliability of collecting information on lithium atoms by this method.
Yu et al. [77] used in situ X-ray imaging to observe the deposition and stripping behavior of Li metal during battery operation and quantitatively characterized the evolution of Li porosity/consolidation during the electroplating process. A single lithium branch growing at a current density of 0.5 mA·cm−2 was larger than that at a higher current density of 10.0 mA·cm−2, whereas the density of lithium branches was smaller (Fig. 13(b)). The effects of lithium-salt concentration, current density, ion strength, different electrolytes, and additives on the morphological evolution of lithium plating/stripping were studied systematically. By adjusting these parameters, the inhibition of Li dendrite formation and formation of uniform lithium deposition were achieved in repeated cycles. This work contributes to understanding the complex lithium plating/stripping mechanism more comprehensively and improving the safety, utilization, and cycle life of lithium metal batteries.
Li et al. [78] captured the first image of atomic lithium dendrites by cryo–electron microscopy (EM). Li dendrites can preserve the original structure and chemical information at low temperatures. In carbon-based electrolytes, lithium dendrites were observed to grow along the <111> (preferred), <110>, or <211> directions, forming single-crystal NWs with a hexahedral structure (Fig. 13(c)). The dendrite growth directions may change at the kink, but no crystal defects were observed. This work provides a simple way to preserve and image the original state of beam-sensitive battery materials at the atomic scale, revealing their detailed nanostructures. The relevant data observed from these experiments can provide a complete understanding of the battery failure mechanism.
Gunnarsdóttir et al. [79] studied the plating, stripping, and corrosion behaviors of lithium ions on a copper collector by using in situ 7Li NMR technology. The method allows the tracking of inactive or dead lithium formation during the plating/stripping of lithium in lithium metal batteries. The formation of dead lithium and SEI can be quantified using NMR, and their relative formation rates in carbonate and ether electrolytes can be compared. Fig. 13(d) shows the in situ 7Li NMR spectra obtained after plating and stripping. The 7Li metal peak in the Cu–LiFePO4 (LFP) cells appeared at about 275 ppm and shifted to a lower direction in the subsequent cycle. Using in situ NMR, the authors demonstrated that polymer coating and surface chemical modification of the Cu current collector contributed to the stabilization of the lithium metal surface.
In general, a stable and uniform SEI layer prevents the continuous parasitic reaction between electrodes and the electrolyte, which is critical for achieving a high coulombic efficiency. The stable cathode–electrolyte interface (CEI) layer can prevent adverse phase transformation by surface reconstruction of the cathode material [80–81]. Therefore, the development of advanced characterization techniques to systematically study the formation and properties of SEI/CEI layers is critical to assist in the development of optimized anode, cathode, and electrolyte combinations.
Liu et al. [82] characterized the SEI formation process between graphite anode and electrolyte using a collaborative strategy, combining the quantitative, in situ, and on-site characteristics of electrochemical quartz crystal microbalance (EQCM) and the atomic accuracy of AFM. The EQCM can accurately measure the accumulation or loss of substances on the graphite electrode during the lithiation process, and the balance can identify lithium fluoride and lithium alkyl carbonate as the dominant chemical compositions at different potentials (Fig. 14(a)). Furthermore, SEI morphology during the first lithiation process was measured by in situ AFM. These quantitative observations can provide useful guidelines for designing and customizing better interphases for new battery chemistry.
Hou et al. [83] in situ observed the formation, growth, and destruction of SEI during lithiation and delithiation under a high current density using a sub-nanoscale mass-sensitive Cs-corrected STEM. Fig. 14(b) shows a schematic of the electrochemical device set for in situ STEM observation of the lithium-ion liquid microbattery. The evolution of SEI film was observed under high angle annular dark-field (HAADF) and annual bright-field (ABF) imaging modes, revealing the dynamic evolution of organic and inorganic layers, respectively. The experimental results showed that the failure of the SEI film was related to the rapid dissolution of the inorganic layer in contact with the electrolyte in the broken SEI film. In addition, the crack in SEI film was caused by the change in electrode volume during the process of delithiation.
Through in situ optical imaging technology, Song et al. [84] observed that the dissolved polysulfide was decomposed, and gases were generated inside the solid-state electrolytes (SSEs) as the cycle number increased, resulting in the severe deformation of the sulfur cathode structure. In addition, the structural evolution and dynamic behavior of composite electrolytes were observed in real time at the nano/microscale through in situ AFM. The results showed that the interaction between the polymer networks and functional filler particles weakened after polysulfide dissolution in composite electrolytes, which significantly affected the mechanical stability of the polymer networks and further intensified the deformation of composite electrolytes (Fig. 14(c)).
Chen et al. [85] reported the dynamic evolution of CEI on the surface of LiNi0.33Co0.33Mn0.33O2 (LNMC) by using a carefully designed in situ surface-enhanced Raman spectroscopy (SERS). To enhance Raman scattering, the scientists deposited gold NPs with enhanced Raman spectral activity on the surface of the LNMC electrode without a binder in monolayer form (Fig. 14(d)). Therefore, in ethylene carbonate/dimethyl carbonate (EC/DMC)-based electrolytes, the chemical vibration patterns generated on the surface of the LNMC electrode had a high sensitivity for Raman detection. Owing to the strong SERS effect of gold monolayer NPs, ester and ether chains were observed in the CEI from the experiment. Through the analysis of dynamic spectral characteristics of the battery in the cycle process, the surface of the LNMC electrode usually formed ether and ester substances under low potentials, which proved that CEI contained organic compounds with ether and ester functional groups. In the process of continuous charge and discharge, the Raman spectrum of CEI presented a very evident dynamic change, which was consistent with the change in the LNMC band.
Understanding the reaction and state of electrolytes during electrochemical cycles is always the key to exploring and designing devices. In recent years, numerous in situ characterization techniques, such as in situ Raman microscopy [86–87], in situ FTIR spectroscopy [88], in situ small-angle neutron scattering, in situ AFM-based imaging, and NMR spectroscopy [89], have been developed and applied to the study of electrolytes during electrochemical cycling.
In common electrochemical equipment, the whole operation process is accompanied by the transfer of Li ions. Electrochemical devices can be designed or studied by measuring the spatial Li-ion concentration. Some studies have achieved this goal with confocal Raman microscopy. Li ions lack optical properties that are easy to capture and can be analyzed by measuring the vibration signatures of closely related molecules. A distinct linear relationship exists between LiClO4 concentration and the fraction of intensity on the sideband (Fig. 15(a)). By combining this high-resolution spatial technology with a simple microfluidic device, the diffusion coefficient of lithium ions in DMC in two different concentration regimes can be determined. Confocal Raman microscopy is intuitive and effective in the determination of lithium-ion transport. Moreover, the value of this approach can be observed in other concentration-dependent spectral electrolyte systems [90].
The research of the evolution of surface chemistry and electrode microstructure during battery cycles is also desirable. Jafta et al. [91] suggested operando small-angle neutron scattering (SANS) to study the effect of different concentrations of electrolytes in SEI formation, pore filling, and carbon framework expansion on the ordered-mesoporous carbon electrode surface (Fig. 15(b)). In the 4 M electrolyte system, Li-rich reduction products formed at high potentials in the micropores, mesopores formed lithium salts quickly before the formation of carbonaceous products. The expansion of the carbon framework was affected by micropore filling and co-intercalation.
A half-cell device containing AFM probe/SSE/Li (Fig. 15(c)) was used to explain the mechanism of SSE degradation from the nanoscale ions and electron transport [92]. Using nm-scale operando imaging, the researchers demonstrated highly nonuniform ion transport, and electron leakage resulted in the degradation of SSE. SSE degradation can be inhibited by covering both sides of SSE with a thin polymer film because the reduction of ions was effectively alleviated despite the slight increase in electronic leakage.
SEI plays a key role in determining the performance of batteries [93]. In addition, during the formation of SEI, the electrochemical reaction of the electrolyte solvent leads to substantial gas evolution, which also largely affects the electrochemical performance. The challenges of volume expansion, explosion, and other safety issues or performance degradation of batteries are largely caused by gas evolution. Thus, studying the composition and proportion of gases can solve a number of the current issues in batteries. To obtain more intuitive and accurate test results, scientists have developed numerous in situ characterization methods for evolving gases, for example, in situ pressure analysis [94], in situ DEMS [95], in situ gas-chromatography (GC) analysis, and in situ Raman spectroscopy [96].
Teng et al. [97] reported a system of GC analysis and thermal conductivity detector that was connected to a set of 900 mAh pouch cells and used for in situ testing of gas generation during battery cycles (Fig. 16(a)). The electrochemical stability of γ-butyrolactone (GBL) was remarkably better than those of other electrolytes used in this work. The electrolyte composed of GBL and ethyl methyl carbonate showed the least gas release in the first charging process of LIBs, which improved the safety and stability of the battery. In addition to quantitative analysis, the system determined the composition of the gases. In GBL-containing electrolytes, the CO and C2H4 generated from EC decomposition were evidently less than those in the EC-containing electrolytes.
In situ DEMS was used to study the effect of various battery parameters on CO2 evolution during the cycling process [98]. High temperatures and battery voltages increased CO2 production (Fig. 16(b)). Additives and lithium carbonate in the cathode electrode affected the release of CO2, and the effect of lithium carbonate only occurred in the first cycle. The type of cathode active material was the main influencing factor for the amount of CO2 evolved. Based on the DEMS method, Zeng et al. [99] evaluated the oxidation stability of the Layered lithium-rich manganese oxide in tris(2,2,2-trifluoroethyl) phosphate (TFEP)-based electrolytes, in which no CO2 evolved throughout the charge/discharge process. MS has high sensitivity and specializes in distinguishing various components by mass to charge (m/z) ratio. They also reported the direct observation of polysulfides in a Li–S battery by combining in situ electrochemistry and MS (Fig. 16(c)) [100]. Several key polysulfide intermediates during sulfur redox have been identified, and their abundance distributions at different potentials enhanced the understanding of the shuttle effect, which largely determines the performance of Li–S batteries.
In summary, numerous leading research groups have conducted pioneering work in the developments and applications of in situ characterization techniques in the past decades, including the design of in situ cells. On the basis of single in situ technology, the multi-modal coupling characterization of different techniques was developed and is expected to eliminate the temporal mismatch of various information. Nevertheless, breaking through the limitation of temporal resolution is a great challenge. Although it is still in its infancy, multi-modal coupling will become a growing trend in the future. In situ characterization reveals indispensable information on how electrode materials behave and how their structures evolute under working conditions. These insights enable researchers to correlate the electrochemical process with the composition, phase structure, geometric structure, etc., at the macro- and micro-scales, which are essential for understanding the failure mechanisms of batteries. Moreover, these insights are also helpful in electrode recycling and sheds light on the utilization of spent LIBs.
Despite the significant progress achieved in the mechanism analysis and electrode material design through in situ techniques, numerous questions remain to be answered in the pursuit of new in situ characterization methods. In the future, more efforts need to be devoted to the below directions, which will promote in situ characterization to further contribute to LBs.
First, although single characterization techniques can also provide abundant information, the combination of different characterization techniques always shows a greater power in obtaining a more comprehensive image of electrode material structure–performance correlation and battery failure mechanism. For example, although in situ MS characterization of the gas evolution can provide some understanding of the battery failure mechanism, researchers can reach a deeper understanding of the ongoing anode, cathode, and electrolyte material structure and surface/interface chemistry evolution in the battery system under operation conditions by correlating the information obtained from different techniques, thus offering better guidance for the design of electrode and electrolyte materials to improve battery applications. Second, scholars should aim for the development of a multi-modal coupling system that collects different modal information of various applicable objectives at multiple length scales, thus decoding diverse physical/chemical phenomena in batteries. For instance, bench-top XAS can be combined with our developed NMR–Raman coupling system, that is, XAS–NMR–Raman. Thus, XAS can be used to analyze the phase structure evolution of cathodes and provide fine insights into the local structure, whereas NMR and Raman can reveal the evolution information of the anode and electrolyte, respectively. Certainly, the interference in different modal techniques should be avoided by carefully designing the assembly site, adopted material, etc., thus ensuring that the obtained information is reliable. In addition, in situ electrochemical cells can be optimized to meet the requirement of multi-modal coupling characterizations, including the selection of transmission window, encapsulation and connection of cell package, and so on. Third, spatial–temporal matching information must be obtained for multi-model coupling characterization. To address the challenge of temporal resolutions, scholars should accurately control the characterization site and signal acquisition time of different techniques. Moreover, a compatible operating system and a data processing system need to be developed for thorough analysis. Currently, most in situ characterizations focus on thermodynamical properties in static studies. Hence, more attention can be paid to the kinetics analysis. The probing of non-equilibrium states can provide hints on the battery reaction feasibility. By improving the data-collection rate, more understanding of the electrode kinetics can be obtained through multi-modal coupling characterizations. Fourth, another interesting direction would be the in situ characterization of batteries under external stimuli, such as optical, thermal, and magnetic. Finally, investigations on batteries will certainly provide insights into other scientific fields, such as physical chemistry and analytical chemistry, which will, in turn, promote the battery field. Moreover, Advances in these techniques and knowledge can promote the development of other electrochemical energy storage systems.
A more exciting understanding of the battery failure mechanism and additional guidance on a better battery design can be obtained with the continuous efforts devoted to in situ characterization technique field. We hope that this review will stimulate exciting investigations and help to accelerate the commercialization of next-generation batteries.
This work was financially supported by the National Natural Science Foundation of China (Nos. 21820102002, 21931012, 22111530178, 51932001, 51872024, and 51972305), the Cooperation Fund of the Dalian National Laboratory for Clean Energy(DNL), Chinese Academy of Science (CAS) (No. DNL202020), the National Key Research and Development Program of China (No. 2018YFA0703503), and the Scientific Instrument Developing Project of the Chinese Academy of Sciences (No. YZ201623).
The authors have no conflicts to declare.
[1] |
E. Pomerantseva, F. Bonaccorso, X.L. Feng, Y. Cui, and Y. Gogotsi, Energy storage: The future enabled by nanomaterials, Science, 366(2019), No. 6468, p. 969. |
[2] |
Z.Y. Gu, J.Z. Guo, X.X. Zhao, et al., High-ionicity fluorophosphate lattice via aliovalent substitution as advanced cathode materials in sodium-ion batteries, InfoMat, 3(2021), No. 6, p. 694. DOI: 10.1002/inf2.12184 |
[3] |
X.X. Luo, W.H. Li, H.J. Liang, et al., Covalent organic framework with highly accessible carbonyls and π-cation effect for advanced potassium-ion batteries, Angew. Chem. Int. Ed., 61(2022), No. 10, art. No. e202117661. |
[4] |
J. Yang, Y.H. Lin, B.S. Guo, et al., Enhanced electrochemical performance of Si/C electrode through surface modification using SrF2 particle, Int. J. Miner. Metall. Mater., 28(2021), No. 10, p. 1621. DOI: 10.1007/s12613-021-2270-x |
[5] |
D.C. Lin, Y.Y. Liu, and Y. Cui, Reviving the lithium metal anode for high-energy batteries, Nat. Nanotechnol., 12(2017), No. 3, p. 194. DOI: 10.1038/nnano.2017.16 |
[6] |
E.H.M. Salhabi, J.L. Zhao, J.Y. Wang, M. Yang, B. Wang, and D. Wang, Hollow multi-shelled structural TiO2–x with multiple spatial confinement for long-life lithium–sulfur batteries, Angew. Chem. Int. Ed., 58(2019), No. 27, p. 9078. DOI: 10.1002/anie.201903295 |
[7] |
Q. Jiang, W.Q. Zhang, J.C. Zhao, P. H. Rao, and J.F. Mao, Superior sodium and lithium storage in strongly coupled amorphous Sb2S3 spheres and carbon nanotubes, Int. J. Miner. Metall. Mater., 28(2021), No. 7, p. 1194. DOI: 10.1007/s12613-021-2259-5 |
[8] |
Y.F. Meng, H.J. Liang, C.D. Zhao, et al., Concurrent recycling chemistry for cathode/anode in spent graphite/LiFePO4 batteries: Designing a unique cation/anion-co-workable dual-ion battery, J. Energy Chem., 64(2022), p. 166. DOI: 10.1016/j.jechem.2021.04.047 |
[9] |
T. Fujita, H. Chen, K.T. Wang, et al., Reduction, reuse and recycle of spent Li-ion batteries for automobiles: A review, Int. J. Miner. Metall. Mater., 28(2021), No. 2, p. 179. DOI: 10.1007/s12613-020-2127-8 |
[10] |
L.Y. Sun, B.R. Liu, T. Wu, et al., Hydrometallurgical recycling of valuable metals from spent lithium-ion batteries by reductive leaching with stannous chloride, Int. J. Miner. Metall. Mater., 28(2021), No. 6, p. 991. DOI: 10.1007/s12613-020-2115-z |
[11] |
J. Lu, T. Wu, and K. Amine, State-of-the-art characterization techniques for advanced lithium-ion batteries, Nat. Energy, 2(2017), art. No. 17011. DOI: 10.1038/nenergy.2017.11 |
[12] |
J. Cui, H.K. Zheng, and K. He, In situ TEM study on conversion-type electrodes for rechargeable ion batteries, Adv. Mater., 33(2021), No. 6, art. No. e2000699. DOI: 10.1002/adma.202000699 |
[13] |
F. Lin, Y.J. Liu, X.Q. Yu, et al., Synchrotron X-ray analytical techniques for studying materials electrochemistry in rechargeable batteries, Chem. Rev., 117(2017), No. 21, p. 13123. DOI: 10.1021/acs.chemrev.7b00007 |
[14] |
S.Y. Lang, Y. Shi, Y.G. Guo, D. Wang, R. Wen, and L.J. Wan, Insight into the interfacial process and mechanism in lithium-sulfur batteries: An in situ AFM study, Angew. Chem. Int. Ed., 55(2016), No. 51, p. 15835. DOI: 10.1002/anie.201608730 |
[15] |
R. Baddour-Hadjean and J.P. Pereira-Ramos, Raman microspectrometry applied to the study of electrode materials for lithium batteries, Chem. Rev., 110(2010), No. 3, p. 1278. DOI: 10.1021/cr800344k |
[16] |
R. Tao, J.G. Zhu, Y.F. Zhang, W.L. Song, H.S. Chen, and D.N. Fang, Quantifying the 2D anisotropic displacement and strain fields in graphite-based electrode via in situ scanning electron microscopy and digital image correlation, Extreme Mech. Lett., 35(2020), art. No. 100635. DOI: 10.1016/j.eml.2020.100635 |
[17] |
S.M. Bak, Z. Shadike, R.Q. Lin, X.Q. Yu, and X.Q. Yang, In situ/operando synchrotron-based X-ray techniques for lithium-ion battery research, NPG Asia Mater., 10(2018), No. 7, p. 563. DOI: 10.1038/s41427-018-0056-z |
[18] |
V.R. Rikka, S.R. Sahu, A. Chatterjee, et al., In situ/ex situ investigations on the formation of the mosaic solid electrolyte interface layer on graphite anode for lithium-ion batteries, J. Phys. Chem. C, 122(2018), No. 50, p. 28717. DOI: 10.1021/acs.jpcc.8b09210 |
[19] |
Y. Yamagishi, H. Morita, Y. Nomura, and E. Igaki, Visualizing lithium distribution and degradation of composite electrodes in sulfide-based all-solid-state batteries using operando time-of-flight secondary ion mass spectrometry, ACS Appl. Mater. Interfaces, 13(2021), No. 1, p. 580. DOI: 10.1021/acsami.0c18505 |
[20] |
A.I. Freytag, A.D. Pauric, S.A. Krachkovskiy, and G.R. Goward, In situ magic-angle spinning 7Li NMR analysis of a full electrochemical lithium-ion battery using a jelly roll cell design, J. Am. Chem. Soc., 141(2019), No. 35, p. 13758. DOI: 10.1021/jacs.9b06885 |
[21] |
E.Y. Zhao, Z.G. Zhang, X.Y. Li, L.H. He, X.Q. Yu, H. Li, and F.W. Wang, Neutron-based characterization techniques for lithium-ion battery research, Chin. Phys. B, 29(2020), No. 1, art. No. 018201. DOI: 10.1088/1674-1056/ab5d07 |
[22] |
Z. Deng, X. Lin, Z.Y. Huang, J.T. Meng, Y. Zhong, G.T. Ma, Y. Zhou, Y. Shen, H. Ding, and Y.H. Huang, Recent progress on advanced imaging techniques for lithium-ion batteries, Adv. Energy Mater., 11(2021), No. 2, art. No. 2000806. DOI: 10.1002/aenm.202000806 |
[23] |
B.H. Song, G.M. Veith, J. Park, M. Yoon, P.S. Whitfield, M.J. Kirkham, J. Liu, and A. Huq, Metastable Li1+δMn2O4 (0≤δ≤1) spinel phases revealed by in operando neutron diffraction and first-principles calculations, Chem. Mater., 31(2019), No. 1, p. 124. DOI: 10.1021/acs.chemmater.8b03199 |
[24] |
Y.P. Han, D.J. Chen, S. Ali, C. Feng, F.P. Meng, M. Waqas, and W.D. He, Hierarchical self-supported carbon nanostructure enables superior stability of highly nitrogen-doped anodes, ChemElectroChem, 7(2020), No. 18, p. 3883. DOI: 10.1002/celc.202001005 |
[25] |
K. Hongyou, T. Hattori, Y. Nagai, T. Tanaka, H. Nii, and K. Shoda, Dynamic in situ Fourier transform infrared measurements of chemical bonds of electrolyte solvents during the initial charging process in a Li ion battery, J. Power Sources, 243(2013), p. 72. DOI: 10.1016/j.jpowsour.2013.05.192 |
[26] |
F. Rittweger, C. Modrzynski, V. Roscher, D.L. Danilov, P.H.L. Notten, and K.R. Riemschneider, Investigation of charge carrier dynamics in positive lithium-ion battery electrodes via optical in situ observation, J. Power Sources, 482(2021), art. No. 228943. DOI: 10.1016/j.jpowsour.2020.228943 |
[27] |
H. Wu, D. Zhuo, D. Kong, and Y. Cui, Improving battery safety by early detection of internal shorting with a bifunctional separator, Nat. Commun., 5(2014), art. No. 5193. DOI: 10.1038/ncomms6193 |
[28] |
Y.C. Hsieh, J.H. Thienenkamp, C.J. Huang et al., Revealing the impact of film-forming electrolyte additives on lithium metal batteries via solid-state NMR/MRI analysis, J. Phys. Chem. C, 125(2021), No. 1, p. 252. DOI: 10.1021/acs.jpcc.0c09771 |
[29] |
J. Wandt, P. Jakes, J. Granwehr, R.A. Eichel, and H.A. Gasteiger, Quantitative and time-resolved detection of lithium plating on graphite anodes in lithium ion batteries, Mater. Today, 21(2018), No. 3, p. 231. DOI: 10.1016/j.mattod.2017.11.001 |
[30] |
R. Sakuma, H. Hashimoto, T. Fujii, J. Takada, N. Hayashi, and M. Takano, In situ Mössbauer analysis of bacterial iron-oxide nano-particles for lithium-ion battery, Hyperfine Interact., 240(2019), No. 1, art. No. 80. DOI: 10.1007/s10751-019-1639-y |
[31] |
N. Balke, S. Kalnaus, N.J. Dudney, C. Daniel, S. Jesse, and S.V. Kalinin, Local detection of activation energy for ionic transport in lithium cobalt oxide, Nano Lett., 12(2012), No. 7, p. 3399. DOI: 10.1021/nl300219g |
[32] |
A.L. Lipson, R.S. Ginder, and M.C. Hersam, Nanoscale in situ characterization of Li-ion battery electrochemistry via scanning ion conductance microscopy, Adv. Mater., 23(2011), No. 47, p. 5613. DOI: 10.1002/adma.201103094 |
[33] |
K. Luo, M.R. Roberts, R. Hao, et al., Charge-compensation in 3d-transition-metal-oxide intercalation cathodes through the generation of localized electron holes on oxygen, Nat. Chem., 8(2016), No. 7, p. 684. DOI: 10.1038/nchem.2471 |
[34] |
J.J. Wang, Y.C.K. Chen-Wiegart, and J. Wang, In operando tracking phase transformation evolution of lithium iron phosphate with hard X-ray microscopy, Nat. Commun., 5(2014), art. No. 4570. DOI: 10.1038/ncomms5570 |
[35] |
S. Pérez-Villar, P. Lanz, H. Schneider, and P. Novák, Characterization of a model solid electrolyte interphase/carbon interface by combined in situ Raman/Fourier transform infrared microscopy, Electrochim. Acta, 106(2013), p. 506. DOI: 10.1016/j.electacta.2013.05.124 |
[36] |
B. Tsuchiya, J. Ohnishi, Y. Sasaki, et al., In situ direct lithium distribution analysis around interfaces in an all-solid-state rechargeable lithium battery by combined ion-beam method, Adv. Mater. Interfaces, 6(2019), No. 14, art. No. 1900100. DOI: 10.1002/admi.201900100 |
[37] |
J.K. Zhu, H.H. Shen, X.B. Shi, et al., Revealing the chemical and structural evolution of V2O5 nanoribbons in lithium-ion batteries using in situ transmission electron microscopy, Anal. Chem., 91(2019), No. 17, p. 11055. DOI: 10.1021/acs.analchem.9b01571 |
[38] |
R. Schmidt, H. Fitzek, M. Nachtnebel, C. Mayrhofer, H. Schroettner, and A. Zankel, The combination of electron microscopy, Raman microscopy and energy dispersive X-ray spectroscopy for the investigation of polymeric materials, Macromol. Symp., 384(2019), No. 1, art. No. 1800237. DOI: 10.1002/masy.201800237 |
[39] |
L.Q. Zhang, T.T. Yang, C.C. Du, et al., Lithium whisker growth and stress generation in an in situ atomic force microscope–environmental transmission electron microscope set-up, Nat. Nanotechnol., 15(2020), No. 2, p. 94. DOI: 10.1038/s41565-019-0604-x |
[40] |
O.J. Borkiewicz, B. Shyam, K.M. Wiaderek, C. Kurtz, P.J. Chupas, and K.W. Chapman, The AMPIX electrochemical cell: A versatile apparatus for in situ X-ray scattering and spectroscopic measurements, J. Appl. Crystallogr., 45(2012), No. 6, p. 1261. DOI: 10.1107/S0021889812042720 |
[41] |
L. Vitoux, M. Reichardt, S. Sallard, P. Novák, D. Sheptyakov, and C. Villevieille, A cylindrical cell for operando neutron diffraction of Li-ion battery electrode materials, Front. Energy Res., 6(2018), art. No. 76. DOI: 10.3389/fenrg.2018.00076 |
[42] |
C. Ghanty, B. Markovsky, E.M. Erickson, et al., 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(2015), No. 10, p. 1479. DOI: 10.1002/celc.201500160 |
[43] |
F. Poli, J.S. Kshetrimayum, L. Monconduit, and M. Letellier, New cell design for in situ NMR studies of lithium-ion batteries, Electrochem. Commun., 13(2011), No. 12, p. 1293. DOI: 10.1016/j.elecom.2011.07.019 |
[44] |
M. Gu, L.R. Parent, B.L. Mehdi, et al., Demonstration of an electrochemical liquid cell for operando transmission electron microscopy observation of the lithiation/delithiation behavior of Si nanowire battery anodes, Nano Lett., 13(2013), No. 12, p. 6106. DOI: 10.1021/nl403402q |
[45] |
X. Zhou, T. Li, Y. Cui, Y. Fu, Y. Liu, and L. Zhu, In situ focused ion beam scanning electron microscope study of microstructural evolution of single tin particle anode for Li-ion batteries, ACS Appl. Mater. Interfaces, 11(2019), No. 2, p. 1733. DOI: 10.1021/acsami.8b13981 |
[46] |
J. Steiger, D. Kramer, and R. Mönig, Microscopic observations of the formation, growth and shrinkage of lithium moss during electrodeposition and dissolution, Electrochim. Acta, 136(2014), p. 529. DOI: 10.1016/j.electacta.2014.05.120 |
[47] |
A.M. Tripathi, W.N. Su, and B.J. Hwang, In situ analytical techniques for battery interface analysis, Chem. Soc. Rev., 47(2018), No. 3, p. 736. DOI: 10.1039/C7CS00180K |
[48] |
M. Sathiya, G. Rousse, K. Ramesha, et al., Reversible anionic redox chemistry in high-capacity layered-oxide electrodes, Nat. Mater., 12(2013), No. 9, p. 827. DOI: 10.1038/nmat3699 |
[49] |
P. Lanz, C. Villevieille, and P. Novák, 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(2014), p. 206. DOI: 10.1016/j.electacta.2014.03.004 |
[50] |
X. Cao, H.F. Li, Y. Qiao, M. Jia, P. He, J. Cabana, and H.S. Zhou, Achieving stable anionic redox chemistry in Li-excess O2-type layered oxide cathode via chemical ion-exchange strategy, Energy Storage Mater., 38(2021), p. 1. DOI: 10.1016/j.ensm.2021.02.047 |
[51] |
L.N. Zhou, M. Leskes, T. Liu, and C.P. Grey, Probing dynamic processes in lithium-ion batteries by in situ NMR spectroscopy: Application to Li1.08Mn1.92O4 electrodes, Angew. Chem. Int. Ed., 54(2015), No. 49, p. 14782. DOI: 10.1002/anie.201507632 |
[52] |
W.M. Seong, K.H. Cho, J.W. Park, et al., Controlling residual lithium in high-nickel (>90 %) lithium layered oxides for cathodes in lithium-ion batteries, Angew. Chem. Int. Ed., 59(2020), No. 42, p. 18662. DOI: 10.1002/anie.202007436 |
[53] |
R. Jung, F. Linsenmann, R. Thomas, et al., Nickel, manganese, and cobalt dissolution from Ni-rich NMC and their effects on NMC622-graphite cells, J. Electrochem. Soc., 166(2019), No. 2, art. No. A378. DOI: 10.1149/2.1151902jes |
[54] |
S. Li, Z.P. Yao, J.M. Zheng, et al., Direct observation of defect-aided structural evolution in a nickel-rich layered cathode, Angew. Chem. Int. Ed., 59(2020), No. 49, p. 22092. DOI: 10.1002/anie.202008144 |
[55] |
L.J. Jia, J. Wang, S.Y. Ren, et al., Unraveling shuttle effect and suppression strategy in lithium/sulfur cells by in situ/operando X-ray absorption spectroscopic characterization, Energy Environ. Mater., 4(2021), No. 2, p. 222. DOI: 10.1002/eem2.12152 |
[56] |
X.C. Liu, Y. Yang, J.J. Wu, et al., Dynamic hosts for high-performance Li–S batteries studied by cryogenic transmission electron microscopy and in situ X-ray diffraction, ACS Energy Lett., 3(2018), No. 6, p. 1325. DOI: 10.1021/acsenergylett.8b00561 |
[57] |
J. Conder, R. Bouchet, S. Trabesinger, C. Marino, L. Gubler, and C. Villevieille, Direct observation of lithium polysulfides in lithium–sulfur batteries using operando X-ray diffraction, Nat. Energy, 2(2017), No. 6, art. No. 17069. DOI: 10.1038/nenergy.2017.69 |
[58] |
H.L. Wu, L.A. Huff, and A.A. Gewirth, In situ Raman spectroscopy of sulfur speciation in lithium-sulfur batteries, ACS Appl. Mater. Interfaces, 7(2015), No. 3, p. 1709. DOI: 10.1021/am5072942 |
[59] |
Z.F. Wang, Y.F. Tang, L.Q. Zhang, M. Li, Z.W. Shan, and J.Y. Huang, In situ TEM observations of discharging/charging of solid-state lithium-sulfur batteries at high temperatures, Small, 16(2020), No. 28, art. No. 2001899. DOI: 10.1002/smll.202001899 |
[60] |
J.Y. Wang, H.J. Tang, H. Wang, R.B. Yu, and D. Wang, Multi-shelled hollow micro-/ nanostructures: Promising platforms for lithium-ion batteries, Mater. Chem. Front., 1(2017), No. 3, p. 414. DOI: 10.1039/C6QM00273K |
[61] |
D.Q. Liu, Z. Shadike, R.Q. Lin, et al., Review of recent development of in situ/operando characterization techniques for lithium battery research, Adv. Mater., 31(2019), No. 28, art. No. 1806620. DOI: 10.1002/adma.201806620 |
[62] |
D.W. Li, Y.K. Wang, B. Lu, and J.Q. Zhang, Real-time measurements of electro-mechanical coupled deformation and mechanical properties of commercial graphite electrodes, Carbon, 169(2020), p. 258. DOI: 10.1016/j.carbon.2020.07.072 |
[63] |
R.D. Ding, Y.L. Huang, G.X. Li, Q. Liao, T. Wei, Y. Liu, Y.J. Huang, and H. He, Carbon anode materials for rechargeable alkali metal ion batteries and in situ characterization techniques, Front. Chem., 8(2020), art. No. 607504. DOI: 10.3389/fchem.2020.607504 |
[64] |
S. Schweidler, L.d. Biasi, A. Schiele, P. Hartmann, T. Brezesinski, and J. Janek, Volume changes of graphite anodes revisited: A combined operando X-ray diffraction and in situ pressure analysis study, J. Phys. Chem. C, 122(2018), No. 16, p. 8829. DOI: 10.1021/acs.jpcc.8b01873 |
[65] |
Y.H. Li, R.T. Zheng, H.X. Yu, et al., Observation of ZrNb14O37 nanowires as a lithium container via in situ and ex situ techniques for high-performance lithium-ion batteries, ACS Appl. Mater. Interfaces, 11(2019), No. 25, p. 22429. DOI: 10.1021/acsami.9b05841 |
[66] |
W.K. Pang, V.K. Peterson, N. Sharma, J.J. Shiu, and S.H. Wu, Lithium migration in Li4Ti5O12 studied using in situ neutron powder diffraction, Chem. Mater., 26(2014), No. 7, p. 2318. DOI: 10.1021/cm5002779 |
[67] |
K. He, S. Zhang, J. Li, et al., Visualizing non-equilibrium lithiation of spinel oxide via in situ transmission electron microscopy, Nat. Commun., 7(2016), art. No. 11441. DOI: 10.1038/ncomms11441 |
[68] |
K. He, Y.F. Yuan, W.T. Yao, et al., Atomistic insights of irreversible Li+ intercalation in MnO2 electrode, Angew. Chem. Int. Ed., 61(2022), No. 2, art. No. e202113420. |
[69] |
B. Song, P. Loya, L.L. Shen, et al., Quantitative in situ fracture testing of tin oxide nanowires for lithium ion battery applications, Nano Energy, 53(2018), p. 277. DOI: 10.1016/j.nanoen.2018.08.057 |
[70] |
K. Kitada, O. Pecher, P.C.M.M. Magusin, M.F. Groh, R.S. Weatherup, and C.P. Grey, Unraveling the reaction mechanisms of SiO anodes for Li-ion batteries by combining in situ 7Li and ex situ 7Li/29Si solid-state NMR spectroscopy, J. Am. Chem. Soc., 141(2019), No. 17, p. 7014. DOI: 10.1021/jacs.9b01589 |
[71] |
Z.Y. Feng, W.J. Peng, Z.X. Wang, H.J. Guo, X.H. Li, G.C. Yan, and J.X. Wang, Review of silicon-based alloys for lithium-ion battery anodes, Int. J. Miner. Metall. Mater., 28(2021), No. 10, p. 1549. DOI: 10.1007/s12613-021-2335-x |
[72] |
J.L. Zhao, J.Y. Wang, R.Y. Bi, et al., General synthesis of multiple-cores@multiple-shells hollow composites and their application to lithium-ion batteries, Angew. Chem. Int. Ed., 60(2021), No. 49, p. 25719. DOI: 10.1002/anie.202110982 |
[73] |
D.X. Liu, J.H. Wang, K. Pan, et al., In situ quantification and visualization of lithium transport with neutrons, Angew. Chem., 126(2014), No. 36, p. 9652. DOI: 10.1002/ange.201404197 |
[74] |
T.Y. Li, X.W. Zhou, Y. Cui, et al., In-situ characterization of dynamic morphological and phase changes of selenium-doped germanium using a single particle cell and synchrotron transmission X-ray microscopy, ChemSusChem, 14(2021), No. 5, p. 1370. DOI: 10.1002/cssc.202002776 |
[75] |
J.H. Um and S.H. Yu, Unraveling the mechanisms of lithium metal plating/stripping via in situ/operando analytical techniques, Adv. Energy Mater., 11(2021), No. 27, art. No. 2003004. DOI: 10.1002/aenm.202003004 |
[76] |
Q. Li, T.C. Yi, X.L. Wang, et al., In-situ visualization of lithium plating in all-solid-state lithium-metal battery, Nano Energy, 63(2019), art. No. 103895. DOI: 10.1016/j.nanoen.2019.103895 |
[77] |
S.H. Yu, X. Huang, J.D. Brock, and H.D. Abruña, Regulating key variables and visualizing lithium dendrite growth: An operando X-ray study, J. Am. Chem. Soc., 141(2019), No. 21, p. 8441. DOI: 10.1021/jacs.8b13297 |
[78] |
Y.Z. Li, Y.B. Li, A. Pei, et al., Atomic structure of sensitive battery materials and interfaces revealed by cryo–electron microscopy, Science, 358(2017), No. 6362, p. 506. DOI: 10.1126/science.aam6014 |
[79] |
A.B. Gunnarsdóttir, C.V. Amanchukwu, S. Menkin, and C.P. Grey, Noninvasive in situ NMR study of “dead lithium” formation and lithium corrosion in full-cell lithium metal batteries, J. Am. Chem. Soc., 142(2020), No. 49, p. 20814. DOI: 10.1021/jacs.0c10258 |
[80] |
X.J. Zeng, D.Q. Liu, S.W. Wang, S. Liu, X.K. Cai, L.H. Zhang, R. Zhao, B.H. Li, and F.Y. Kang, In situ observation of interface evolution on a graphite anode by scanning electrochemical microscopy, ACS Appl. Mater. Interfaces, 12(2020), No. 33, p. 37047. DOI: 10.1021/acsami.0c07250 |
[81] |
L.Y. Wang, L.F. Wang, R. Wang, R. Xu, C. Zhan, W. Yang, and G.C. Liu, Solid electrolyte-electrode interface based on buffer therapy in solid-state lithium batteries, Int. J. Miner. Metall. Mater., 28(2021), No. 10, p. 1584. DOI: 10.1007/s12613-021-2278-2 |
[82] |
T.C. Liu, L.P. Lin, X.X. Bi, et al., In situ quantification of interphasial chemistry in Li-ion battery, Nat. Nanotechnol., 14(2019), No. 1, p. 50. DOI: 10.1038/s41565-018-0284-y |
[83] |
C. Hou, J.H. Han, P. Liu, et al., Operando observations of SEI film evolution by mass-sensitive scanning transmission electron microscopy, Adv. Energy Mater., 9(2019), No. 45, art. No. 1902675. DOI: 10.1002/aenm.201902675 |
[84] |
Y.X. Song, Y. Shi, J. Wan, B. Liu, L.J. Wan, and R. Wen, Dynamic visualization of cathode/electrolyte evolution in quasi-solid-state lithium batteries, Adv. Energy Mater., 10(2020), No. 25, art. No. 2000465. DOI: 10.1002/aenm.202000465 |
[85] |
D.C. Chen, M.A. Mahmoud, J.H. Wang, et al., Operando investigation into dynamic evolution of cathode–electrolyte interfaces in a Li-ion battery, Nano Lett., 19(2019), No. 3, p. 2037. DOI: 10.1021/acs.nanolett.9b00179 |
[86] |
L.M. Suo, Z. Fang, Y.S. Hu, and L.Q. Chen, FT-Raman spectroscopy study of solvent-in-salt electrolytes, Chin. Phys. B, 25(2016), No. 1, art. No. 016101. DOI: 10.1088/1674-1056/25/1/016101 |
[87] |
C.Y. Li, Y. Yu, C. Wang, et al., Surface changes of LiNixMnyCo1–x–yO2 in Li-ion batteries using in situ surface-enhanced Raman spectroscopy, J. Phys. Chem. C, 124(2020), No. 7, p. 4024. DOI: 10.1021/acs.jpcc.9b11677 |
[88] |
Y.R. Zhang, Y. Katayama, R. Tatara, et al., Revealing electrolyte oxidation via carbonate dehydrogenation on Ni-based oxides in Li-ion batteries by in situ Fourier transform infrared spectroscopy, Energy Environ. Sci., 13(2020), No. 1, p. 183. DOI: 10.1039/C9EE02543J |
[89] |
F. Wohde, R. Bhandary, J.M. Moldrickx, J. Sundermeyer, M. Schönhoff, and B. Roling, Li+ ion transport in ionic liquid-based electrolytes and the influence of sulfonate-based zwitterion additives, Solid State Ion., 284(2016), p. 37. DOI: 10.1016/j.ssi.2015.11.017 |
[90] |
J.D. Forster, S.J. Harris, and J.J. Urban, Mapping Li+ concentration and transport via in situ confocal Raman microscopy, J. Phys. Chem. Lett., 5(2014), No. 11, p. 2007. DOI: 10.1021/jz500608e |
[91] |
C.J. Jafta, X.G. Sun, G.M. Veith, et al., Probing microstructure and electrolyte concentration dependent cell chemistry via operando small angle neutron scattering, Energy Environ. Sci., 12(2019), No. 6, p. 1866. DOI: 10.1039/C8EE02703J |
[92] |
C.S. Jiang, Y. Yin, H. Guthrey, K. Park, S.H. Lee, and M.M. Al-Jassim, Local electrical degradations of solid-state electrolyte by nm-scale operando imaging of ionic and electronic transports, J. Power Sources, 481(2021), art. No. 229138. DOI: 10.1016/j.jpowsour.2020.229138 |
[93] |
H.J. Liang, B.H. Hou, W.H. Li, et al., Staging Na/K-ion de-/ intercalation of graphite retrieved from spent Li-ion batteries: In operando X-ray diffraction studies and an advanced anode material for Na/K-ion batteries, Energy Environ. Sci., 12(2019), No. 12, p. 3575. DOI: 10.1039/C9EE02759A |
[94] |
A. Schiele, T. Hatsukade, B.B. Berkes, P. Hartmann, T. Brezesinski, and J. Janek, High-throughput in situ pressure analysis of lithium-ion batteries, Anal. Chem., 89(2017), No. 15, p. 8122. DOI: 10.1021/acs.analchem.7b01760 |
[95] |
B. Michalak, B.B. Berkes, H. Sommer, T. Bergfeldt, T. Brezesinski, and J. Janek, Gas evolution in LiNi0.5Mn1.5O4/graphite cells studied in operando by a combination of differential electrochemical mass spectrometry, neutron imaging, and pressure measurements, Anal. Chem., 88(2016), No. 5, p. 2877. DOI: 10.1021/acs.analchem.5b04696 |
[96] |
B. Gerelt-Od, J. Kim, E. Shin, et al., In situ Raman investigation of resting thermal effects on gas emission in charged commercial 18650 lithium ion batteries, J. Ind. Eng. Chem., 96(2021), p. 339. DOI: 10.1016/j.jiec.2021.01.039 |
[97] |
X. Teng, C. Zhan, Y. Bai, et al., In situ analysis of gas generation in lithium-ion batteries with different carbonate-based electrolytes, ACS Appl. Mater. Interfaces, 7(2015), No. 41, p. 22751. DOI: 10.1021/acsami.5b08399 |
[98] |
J. Vetter, M. Holzapfel, A. Wuersig, W. Scheifele, J. Ufheil, and P. Novák, In situ study on CO2 evolution at lithium-ion battery cathodes, J. Power Sources, 159(2006), No. 1, p. 277. DOI: 10.1016/j.jpowsour.2006.04.087 |
[99] |
Z.Q. Zeng, X.W. Liu, X.Y. Jiang, et al., Enabling an intrinsically safe and high-energy-density 4.5 V-class Li-ion battery with nonflammable electrolyte, InfoMat, 2(2020), No. 5, p. 984. DOI: 10.1002/inf2.12089 |
[100] |
Z.Y. Yu, Y. Shao, L.P. Ma, et al., Revealing the sulfur redox paths in a Li–S battery by an in situ hyphenated technique of electrochemistry and mass spectrometry, Adv. Mater., 34(2022), No. 7, art. No. 2106618. DOI: 10.1002/adma.202106618 |
Xilin Chen, Zhixiang Zhong, Xingkuan Chen, et al. Advancements in cathode materials for aqueous potassium-ion batteries. Energy Storage Materials, 2025, 74: 103948.
![]() | |
Shaorou Ke, Yajing Zhao, Xin Min, et al. Highly mass activity electrocatalysts with ultralow Pt loading on carbon black for hydrogen evolution reaction. International Journal of Minerals, Metallurgy and Materials, 2025, 32(1): 182.
![]() | |
Shangyong Zhao, Yuchen Zhao, Yujia Dai, et al. Recent advances in chemical composition imaging operation based on laser-induced breakdown spectroscopy. Journal of Analytical Atomic Spectrometry, 2025.
![]() |
In situ technique | Spatial resolution | Applicable scenes and obtained information | Facing challenge | |
X-ray technique | XRD | ~μm scale | Quantitatively and qualitatively provide composition, distribution, and conversion of crystallographic structure analyses; study local environment of an element of electrodes. | Amorphous and low atomic number elements in LIBs. |
XAS | 1–100 nm | |||
TXM, XRT | ≥15 nm | |||
XPS | ~μm scale | |||
Neutron-based technique | NPD | High penetration and sensitive to low atomic number elements. Precisely determine structural and dynamic information of electrodes. | Amorphous and high atomic number elements in LIBs. | |
NR | nm | |||
NI | ~μm scale | |||
NDP | nm | |||
Electron-based technique | TEM | ≥0.1 nm | Provide morphological information of electrodes and SEI; obtain structural information by coupling accessory. | Shallow penetration; high vacuum; high-energy electron damage. |
SEM | 10 nm | |||
EELS | ||||
Optical technique | Raman spectroscopy | 10 nm–1 μm | Probe the molecule structural features and the chemical composition of electrodes, SEI, and electrolytes. Visualization of macrostructural features. | Specific active component; signal intensity and background interference; bulky or depth analysis. |
FTIR spectroscopy | ~μm scale | |||
OM | ~μm scale | |||
Scanning probe technique | AFM | nm | Monitor morphology and changes of surface/interface in diverse environments, SEI, Li dendrite, etc., as well as mechanical properties and electrical properties. | Low resolution |
ESM | 10 nm–1μm | |||
SICM | 10 nm–50 nm | |||
Magnetic resonance technique | NMR | mm | Apply to an odd atomic mass or odd atomic number element of LIBs; monitor the local electronic environment around the nucleus through electrodes, electrolytes, and their interfaces. | Specific elements; low resolution; |
EPR | ||||
Mössbauer spectroscopy |