Recent advances in in situ and operando characterization techniques for Li7La3Zr2O12-based solid-state lithium batteries

Lei Zhang ab, Huilin Fan ab, Yuzhen Dang ab, Quanchao Zhuang *c, Hamidreza Arandiyan *de, Yuan Wang f, Ningyan Cheng g, Hongyu Sun h, H. Hugo Pérez Garza h, Runguo Zheng ab, Zhiyuan Wang ab, Sajjad S. Mofarah i, Pramod Koshy i, Suresh K. Bhargava e, Yanhua Cui j, Zongping Shao k and Yanguo Liu *ab
aSchool of Materials Science and Engineering, Northeastern University, Shenyang 110819, China. E-mail: lyg@neuq.edu.cn
bSchool of Resources and Materials, Northeastern University at Qinhuangdao, Qinhuangdao 066004, China
cSchool of Materials and Physics, China University of Mining & Technology, Xuzhou 221116, China. E-mail: 4719@cumt.edu.cn
dLaboratory of Advanced Catalysis for Sustainability, School of Chemistry, University of Sydney, Sydney, NSW 2006, Australia. E-mail: hamid.arandiyan@sydney.edu.au
eCentre for Advanced Materials and Industrial Chemistry (CAMIC), School of Science, RMIT University, Melbourne, VIC 3000, Australia
fInstitute for Frontier Materials, Deakin University, Melbourne, Vic 3125, Australia
gInformation Materials and Intelligent Sensing Laboratory of Anhui Province, Key Laboratory of Structure and Functional Regulation of Hybrid Materials of Ministry of Education, Institutes of Physical Science and Information Technology, Anhui University, Hefei 230601, China
hDENSsolutions B.V., Informaticalaan 12, 2628 ZD Delft, The Netherlands
iSchool of Materials Science and Engineering, UNSW Sydney, Sydney, NSW 2052, Australia
jInstitute of Electronic Engineering, China Academy of Engineering Physics, Mianyang 621900, China
kWA School of Mines: Minerals, Energy and Chemical Engineering, Curtin University, Perth, WA 6845, Australia

Received 1st February 2023 , Accepted 15th March 2023

First published on 23rd March 2023


Abstract

Li7La3Zr2O12 (LLZO)-based solid-state Li batteries (SSLBs) have emerged as one of the most promising energy storage systems due to the potential advantages of solid-state electrolytes (SSEs), such as ionic conductivity, mechanical strength, chemical stability and electrochemical stability. However, there remain several scientific and technical obstacles that need to be tackled before they can be commercialised. The main issues include the degradation and deterioration of SSEs and electrode materials, ambiguity in the Li+ migration routes in SSEs, and interface compatibility between SSEs and electrodes during the charging and discharging processes. Using conventional ex situ characterization techniques to unravel the reasons that lead to these adverse results often requires disassembly of the battery after operation. The sample may be contaminated during the disassembly process, resulting in changes in the material properties within the battery. In contrast, in situ/operando characterization techniques can capture dynamic information during cycling, enabling real-time monitoring of batteries. Therefore, in this review, we briefly illustrate the key challenges currently faced by LLZO-based SSLBs, review recent efforts to study LLZO-based SSLBs using various in situ/operando microscopy and spectroscopy techniques, and elaborate on the capabilities and limitations of these in situ/operando techniques. This review paper not only presents the current challenges but also outlines future developmental prospects for the practical implementation of LLZO-based SSLBs. By identifying and addressing the remaining challenges, this review aims to enhance the comprehensive understanding of LLZO-based SSLBs. Additionally, in situ/operando characterization techniques are highlighted as a promising avenue for future research. The findings presented here can serve as a reference for battery research and provide valuable insights for the development of different types of solid-state batteries.


image file: d3mh00135k-p1.tif

Lei Zhang

Lei Zhang is currently a PhD candidate under the supervision of Prof. Yanguo Liu at Northeast University. He studies solid-state lithium batteries in energy storage and conversion, including oxide, sulfide, and composite electrolyte systems. Meanwhile, he is interested in the impact of interface interactions (anode/electrolyte interfaces and cathode/electrolyte interfaces) on ions and electron transport and is committed to the development of solid-state lithium batteries with stable cycling ability.

image file: d3mh00135k-p2.tif

Huilin Fan

Huilin Fan is currently a PhD candidate under the supervision of Prof. Yanguo Liu at Northeast University. She focuses on the applications of metal chalcogenides, including metal sulfide, selenide, and telluride, in energy storage and conversion. Her research also includes the development of alkali metal ion batteries with ultra-fast charging capabilities by studying the effect of interface interactions (homogeneous interfaces and heterogeneous interfaces) on ion and electron transport.

image file: d3mh00135k-p3.tif

Quanchao Zhuang

Quanchao Zhuang is a Professor of Materials Science and Engineering at the China University of Mining and Technology. He obtained his PhD from Xiamen University in 2007. He is mainly devoted to electrochemical impedance spectroscopy research on the electrode interface reaction mechanism of Li-ion batteries and special-purpose Li primary batteries. He has hosted over 20 projects, including equipment pre-research projects, NSAF joint fund key projects, and NSAF joint fund incubation projects. He has also edited an English monograph, held over 20 authorized invention patents, and published over 130 SCI papers.

image file: d3mh00135k-p4.tif

Hamidreza Arandiyan

Hamidreza Arandiyan received his PhD from Tsinghua University in 2014. He was awarded a Vice-Chancellor's Research Fellowship from UNSW Sydney in 2015. He was awarded the University of Sydney Senior Fellowship from the School of Chemistry in 2018. Currently, he is a Senior Research Fellow in Applied Chemistry & Environmental Science at RMIT University, Melbourne, expanding his work on functional battery materials. His research interests lie in heterogeneous catalysis for environmental remediation and energy applications. He holds the title of Fellow of the Royal Society of Chemistry (FRSC) and is also a member of the Royal Australian Chemical Institute (MRACI CChem).

image file: d3mh00135k-p5.tif

Yuan Wang

Yuan Wang (Helena) received her PhD degree from the School of Chemical Engineering, The University of New South Wales (UNSW), Sydney, Australia, in 2019. She was appointed as a Research Assistant at the National Research Center for Geoanalysis, Chinese Academy of Geological Science, in 2013. She was a Postdoctoral Research Associate at the School of Chemistry, UNSW, Sydney, and a Research Fellow in Applied Chemistry & Environmental Science at RMIT University, Melbourne. Currently, she is an Alfred Deakin Research Fellow and ARC DECRA Fellow at the Institute for Frontier Materials (IFM) at Deakin University, Melbourne. Her research interests focus on developing and characterising nanoporous materials for metal-air batteries and electrocatalytic applications.

image file: d3mh00135k-p6.tif

Hongyu Sun

Hongyu Sun received his PhD degree from the State Key Laboratory of Metastable Materials Science and Technology, Yanshan University, in 2010. He then worked in the Department of Materials Science and Engineering (2010–2012), Beijing National Centre for Electron Microscopy (2012–2015), at Tsinghua University (with Prof. Jing Zhu), and the Department of Micro- and Nanotechnology at the Technical University of Denmark (2015–2018, with Prof. Kristian Mølhave). Now he is a senior application scientist at DENSsolution B.V. He is the recipient of the Robert P. Apkarian Award (Physical Sciences, 2018) of the Microscopy Society of America. His research interests include controllable synthesis of functional structures for energy storage and conversion, liquid cell transmission electron microscopy, and microfabrication.

image file: d3mh00135k-p7.tif

Suresh K. Bhargava

Suresh Bhargava is a world-renowned interdisciplinary scientist with decades of leadership in academia and industry. With over 600 highly cited publications including 2 books and 16 book chapters, 12 patents (6 transferred to industry) with citations >23[thin space (1/6-em)]000. Recipient of many awards, distinguished Prof. Bhargava's academic excellence and remarkable aptitude in integrating fundamental science with engineering sets him apart as a global leader of translational research. The contributions of Prof. Suresh Bhargava exceed far beyond academic knowledge into the realm of new discoveries, invention of new technologies and successful implementation of industrial processes. He is a fellow of seven academies around the world, including ATSE, AAAS and TWAS.

image file: d3mh00135k-p8.tif

Zongping Shao

Zongping Shao is a Professor of Chemical Engineering at Curtin University, Australia, and Nanjing Tech University, China. He obtained his PhD from the Dalian Institute of Chemical Physics, China, in 2000. He worked as a Visiting Scholar at the Institut de Researches sur la Catalyse, CNRS, France, and as a Postdoctoral Research Fellow at the California Institute of Technology, USA, from 2000 to 2005. His research interests include solid oxide fuel cells, Li-ion batteries, supercapacitors, and low-temperature energy-conversion devices. He has been recognised as a highly cited researcher in the energy section by Elsevier since 2015. He was selected as a World Highly Cited Researcher by (Thomson Reuters) Clairvate Analytics in 2014, 2017, 2018, and 2019.

image file: d3mh00135k-p9.tif

Yanguo Liu

Yanguo Liu received his PhD degree in Materials Physics and Chemistry from Yanshan University in 2009. He is currently a professor at Northeast University. His research interests include functional nanostructures and energy science. He has published over 50 papers in peer-reviewed journals and holds 15 patents on energy storage materials.



New concepts

Li-ion batteries have been widely used in small- and medium-sized battery applications such as mobile devices, emergency power systems, and hybrid electric vehicles. However, commercial Li-ion batteries with liquid electrolytes have disadvantages such as leakage, flammability, and incompatibility with Li-metal. High-capacity solid-state lithium batteries (SSLBs), in which liquid electrolytes are replaced with solid-state electrolytes, are expected to completely address the safety of batteries while simultaneously addressing the low energy density and short lifespan of conventional Li-ion batteries. Among different SSEs, Li7La3Zr2O12 (LLZO) garnet-based electrolytes possess advantages of ionic conductivity, chemical stability, electrochemical stability, environmental stability, and safety. They are therefore considered one of the most promising and vital electrolytes for achieving high-energy-density and high-safety SSLBs. However, several scientific and technical obstacles still need to be tackled before LLZO-based SSLBs can be commercialised. Herein, this review briefly describes the key challenges currently faced by LLZO-based SSLBs, reviews recent efforts to study them using various in situ/operando microscopy and spectroscopy techniques, and illustrates the capabilities and limitations of these in situ/operando techniques, thus enhancing the comprehensive understanding of LLZO-based SSLBs, paving the way for LLZO-based SSLBs, and providing valuable insights for the development of different types of SSLBs.

1. Introduction

Nowadays, Li-ion batteries (LIBs) are gaining significant prominence in various fields such as electric vehicles, laptops, smartphones, medical devices, and military weaponry.1 With the development of industry and the demand for high-quality life, LIBs are expected to provide higher energy density and power density.2 However, organic liquid electrolytes for LIBs are highly flammable and exhibit narrow electrochemical windows.3 More importantly, liquid electrolytes are vulnerable to Li dendrites, when in contact with Li metal (3860 mA h g−1, −3.04 V vs. standard hydrogen electrode) and are incompatible with high-voltage cathodes.4 Such critical issues are impeding the further enhancement of energy density and power density for LIBs. SSEs with a wide electrochemical window, non-flammable nature, stability to Li metal, and compatibility with high-voltage cathodes are appropriate candidates instead of liquid electrolytes.5 Among SSEs, LLZO garnet-type electrolytes are regarded as one of the most promising and crucial solid-state electrolytes (SSEs) due to their superior ionic conductivity, (electro)chemical stability, air stability, thermal stability, and safety.6,7 Therefore, LLZO-based SSLBs possess the highest potential for commercial applications, which provides a tremendous opportunity for the development of high energy density and power density solid-state batteries.6,8

However, achieving high energy/power density solid-state batteries is still challenging. During charging and discharging processes, LLZO-based SSLBs exhibit significant volume changes, lower capacity, poor cycling performance, and rapid capacity degradation.9 These findings highlight the need to urgently address the different scientific and technical issues affecting their performance. These include:

(i) Poor ionic conductivity of LLZO-based SSEs compared to that of liquid electrolytes: to enhance the ionic conductivity, better comprehension of both Li+ migration mechanisms inside the LLZO crystalline lattice and effective mechanisms of doped elements and phase conversion mechanisms in synthesis procedures are essential. The preparation of several tailored-LLZO-based SSEs and correlated parameters also need to be clarified. With regard to electrochemical response, it is crucial to elucidate the degradation, deterioration and failure of LLZO-based SSEs and the specific migration pathway of Li+ on the microscale.

(ii) Incompatibility of LLZO-based SSEs with Li or active cathode materials, considering factors such as interfacial contact (such as wettability and volume variation), interfacial (electro)chemical reactions, interfacial degradation, Li dendrite formation (such as the morphology, structure, and formation mechanism) and space charge layer properties: to address these challenges, various strategies have been proposed, such as developing composite anodes or cathodes, removing the LLZO surface contamination layer, and introducing an artificial interface-modified layer. However, the root cause of issues and the effective mechanism involved must be identified. In addition, the degradation/deterioration phenomena of active cathode materials and ionic/electronic conductive pathways inside the cathode must be elaborated. Furthermore, the reaction principles when matching SSLBs with different cathode active materials, limiting the steps of the complete kinetic process, volume expansion/contraction during electrochemical cycling, and thermal runaway are critical factors that need to be thoroughly investigated. Therefore, it is imperative to acquire a comprehensive understanding of the principles and mechanisms behind the aforementioned issues to enable the practical implementation of high energy density and safe LLZO-based SSLBs. To this end, appropriate characterization techniques play a critical role.

As shown in Fig. 1, various in situ/operando characterization techniques have been developed and employed to reveal the issues faced by LLZO-based SSLBs from multiple perspectives. Almost all characterization experiments involve a compromise between establishing practical cell operating conditions and extracting high-quality data. The analysis region, time, and environment are the three important parameters to consider for each characterization technique.10 The analysis region is the most critical parameter for any characterization technique since any macroscopic variations in LLZO-based SSLBs can be determined from the reactions or interactions on an atomic scale. For a cell, a range of relevant features such as cell assembly, composite cathodes, dendrite/cracks, microstructural features, interphase formation, and lattice structures are all seen over a dimension of 10−2–10−10 m. Large region characterization techniques can locate the reaction/failure mechanism, while small-scale characterization techniques can analyze the root cause. Therefore, characterization techniques of multi-scale resolution or integration of various resolutions are essential. Time is another important parameter since different electrochemical reactions such as cell charge and discharge, chemico-mechanical deformation, diffusion, charge transfer, interphase formation and ion hopping occur in times ranging from 105–10−9 s.


image file: d3mh00135k-f1.tif
Fig. 1 Schematic diagram highlighting the challenges in various components in an LLZO-based SSLB. The spatial distribution of features and processes occurring in LLZO-based SSLBs and spatial resolution correspond to various in situ/operando characterization techniques.

The final important parameter is the environment. Characterization approaches generally include ex situ, in situ, and operando, referring to an investigation conducted post-experiment, at the experiment site, or during electrochemical operation, respectively.11Ex situ characterization techniques are valuable in offering an extensive overview of the system state, function, and failure mechanism. However, this requires partial or complete disassembly of the cell prior to experiment, failing to acquire dynamic critical information during cell operation. Besides, cell disassembly can cause damage and contamination and thus affect the characterization results. In situ/operando characterization techniques avoid the negative effects induced by cell disassembly and allow for real-time cell monitoring. The distinction between “in situ” and “operando” is that in situ experiments are measured in “relaxed” states, where Li plating/stripping is paused before data acquisition. In contrast, operando experiments with sufficient time resolution are conducted during plating/stripping continuously. Due to the limitations of instruments, experimental setups, and conditions, in situ/operando experiments cannot perfectly replicate the practical cell operating environment. Accordingly, the effect of several key components, such as contaminants, operating environment, applied pressure, and current density are disregarded. This presents a huge challenge in bridging the divide between experiments and actual application. Nevertheless, in situ/operando characterization techniques are beneficial to unravel the rationale and mechanisms involved in LLZO-based SSLB operation and failure.

This review provides an overview of the critical challenges associated with LLZO-based SSLBs from four perspectives: LLZO-based solid-state electrolytes (SSEs), composite cathodes/Li anodes, electrode/LLZO-based SSE interfaces, and a full-cell battery. These perspectives are first briefly introduced. Then, recent developments and progress of various in situ/operando characterization techniques in the field of LLZO-based SSLBs are examined, including optical, electron, X-ray, scanning probe, magnetism, and neutron-based, respectively and other techniques. Each technique is illustrated from multiple perspectives, considering the working principle, experimental setup, operating conditions, cell structure, advantages, limitations, and applications. An in-depth comprehension of failure mechanisms from various in situ/operando characterization techniques is vital to successfully develop high energy density and safe LLZO-based SSLBs. Eventually, the opportunities and challenges for the future are presented. Each technique is not restricted to its own development but is also integrated with other fields. This review is expected to help address the identified challenges and develop a more comprehensive understanding of LLZO-based SSLBs. Furthermore, it is also expected to illustrate the understanding that can be gained from in situ/operando characterization techniques and serve as a reference for improved understanding in other related electrochemical fields.

2. LLZO-based SSLB challenges

Solid-state batteries rely on the ability to work with high energy density Li anodes and high-voltage cathodes to realize technical and economic feasibility. However, several critical challenges hinder the achievement of high energy density, stable, and safe LLZO-based SSLBs. In this regard, the electrolyte function, electrode structure design, interface chemistry, and the correlations between these factors have to be understood in depth. This section will briefly illustrate the current critical issues and/or challenges from four perspectives, as shown in Fig. 1, LLZO-based SSEs, composite cathodes/Li anodes, electrodes/LLZO-based SSE interfaces, and full-cell batteries.

2.1 LLZO-based SSEs

The ionic conductivity of LLZO-based SSEs is relatively low compared to that of liquid electrolytes. Further enhancing the ionic conductivity is a critical issue.12 LLZO contains two structures in the tetragonal phase (t-LLZO) and cubic phase (c-LLZO), with Ia[3 with combining macron]d and I[4 with combining macron][3 with combining macron]d space groups, respectively. In general, the ionic conductivity of t-LLZO is two orders of magnitude lower than that of c-LLZO. Therefore, elucidating the migration mechanism of Li+ within the tetragonal or cubic LLZO lattice is essential for understanding the variability of ionic conductivity. Additionally, t-LLZO, in contrast to c-LLZO, is stable at room temperature. At higher temperatures, t-LLZO undergoes a phase transition to c-LLZO with high Li+ conductivity. Based on this, understanding the mechanism behind the phase conversion between t-LLZO and c-LLZO is essential to ensure the stable formation of the cubic phase. Numerous works have demonstrated that introducing dopants, including Li-site doping, La-site doping, Zr-site doping, and anion doping, can stabilize the cubic phase at room temperature.9 Due to differences in radii and charges, introducing dopant atoms at any site inevitably induces changes in bond length, Li+ concentration, and Li+ migration channel size, which results in changes in the ionic conductivity. Therefore, thoroughly understanding the effects of dopant atoms is critical to achieve further improvements in the ionic conductivity.

Ideally, LLZO-based SSEs should possess a combination of favorable properties, such as high ionic conductivity, low thickness, wide electrochemical stability window, excellent mechanical properties, and favorable thermal stability. However, achieving a balance between these properties simultaneously in LLZO-based SSEs is challenging. For example, although LLZO-based composite electrolytes (CEs) feature advantages of low thickness and outstanding mechanical properties, the ionic conductivity is relatively low. The distribution of individual components and Li+ transport pathways (either along a single-phase matrix or two-phase interface in CEs) significantly affect the ionic conductivity. Moreover, the preparation of several LLZO-based SSEs with special morphologies and properties along with identification of correlated parameters has been achieved and this includes the preparation of LLZO crystal films with stoichiometric ratios or LLZO porous structures with homogeneous pore distributions.13 During electrochemical cycling, structural defects such as pores, grain boundaries, and cracks in LLZO ceramic pellets can be Li dendrite nucleation sites, which need to be minimized.14 On exposure to high humidity atmospheres (containing high amounts of H2O and CO2), ionic conductivity of LLZO-based SSEs decreases dramatically, owing to the formation of a contamination layer (mainly Li2CO3).15 This significantly affects the electrochemical properties of LLZO-based SSLBs. Hence, it is vital to investigate the effect of air exposure on LLZO structure stability, ionic mobility, and surface conformation.

2.2 Composite cathodes/Li anodes

To ensure utilization of all active cathode particles, each particle must transport ions and electrons to the SSE and current collector. However, the electronic and ionic conductivities of cathode materials tend to be relatively low, indicating that composite cathodes comprising active materials, solid electrolytes, and conductive carbon additives must be employed.16 Optimally, every cathode particle is surrounded by SSE/carbon and ions/electrons can only penetrate through one active material particle. This enhances capacity utilization and high-rate capability due to lower cell overpotential, which allows more capacity to be available before the voltage limit is reached. In fact, several active material particles in composite cathodes are isolated from SSE/carbon and SSE particles cannot flow in a manner similar to liquid electrolytes. As a result, the issue of achieving the required composite cathode structure for stabilizing three-phase boundaries is mainly unresolved. The design and preservation of a perfect Li+/e transport channel in the interior of composite cathodes is also an immediate issue to be addressed. In addition, during charging and discharging, cathode particles experience considerable volume changes, leading to loss of contact with SSE/carbon.17 Cathode particles are also subject to structural transformation or phase conversion, which causes electronic and ionic conductivity changes.18,19 In depth awareness of such processes and clarity of principles are beneficial for providing the appropriate solutions. On the other hand, the effects of volume expansion in Li metal anodes should not be neglected. Due to restrictions in terms of the cell structure being fixed, volume expansion causes an increase in stack pressure that is exerted on the cell, inducing a mechanical short-circuit.20 However, quantifying the degree of volume expansion is a challenge even though controlling the volume expansion of Li anodes is essential during Li plating/stripping. One possible solution to overcome the mechanical short-circuits caused by high pressure is to develop a composite or three-dimensional (3D) Li anode. Such anodes would have a higher melting point but their creep tendency would be much lower than that of pure Li metal.21 Thus such modified anodes can withstand higher operating pressures during lithiation due to more favorable mechanical properties. Consequently, a detailed analysis of composite or 3D Li anode creep and flow behaviors is required to tune and enhance the system performance.

2.3 Electrodes/LLZO-based SSE interfaces

A notable interface incompatibility is seen between LLZO-based SSEs and Li anodes or high-voltage cathodes. One of the significant inevitable issues is high impedance at electrode/electrolyte interfaces. Due to solid–solid contact and absence of liquid electrolyte wetting, the interface has a point-to-point contact. Interface reconstruction, such as coating of a melted composite anode, introduction of artificial interface modification layers or coating of cathode material surfaces can usually assist in achieving intimate contact and lowering interface impedance.9 Accordingly, the composition, structure of reconstructed interfaces and means of achieving intimate contact (such as chemical bonding, conformal contact or intermolecular forces) require further elucidation. Importantly, the kinetic transport behavior at both original and reconstructed interfaces is mainly unexplored, which is still a significant challenge limiting solid-state battery applications.

Another critical issue is that Li dendrites penetrate electrolytes, resulting in cell failure.22 Li dendrite formation can severely limit the rate performance, power density, and Coulombic efficiency of the solid-state batteries. Thus a thorough understanding of the morphology, structure and formation mechanisms of Li dendrites is vital to suppress their growth. The conductivity of the interface (including ion-conductive, electron-conductive, or mixed ion-/electron-conductive) affects Li dendrite growth but the mechanisms are still unclear. In addition to high interface impedance and Li dendrite growth, side reactions or elemental interdiffusion between anodes/cathodes and LLZO-based SSEs can lead to the formation ionic insulation by-products and electrochemically isolated Li (dead), both of which contribute to unrecoverable capacity loss.9 The mechanism of side reactions, root cause of the diffusion behavior, specific composition of products, and distribution of dead Li require further comprehensive investigation. Moreover, high rate electro-dissolution from Li anodes can induce interfacial pore formation, leading to the onset of failure.23 Moreover, integrating high-voltage cathodes with SSEs can contribute to space charge layer formation that obstructs Li+/e motion through the electrode/electrolyte interface.24 Therefore, it is imperative to determine the underlying cause of such issues.

2.4 Full-cell batteries

Most importantly, for full-cell batteries, ∼5 mA h cm−2 reversible cycle capacity at 5 mA cm−2 current density with high Coulombic efficiency is far from realization.25 Unstable transformation at various solid/solid interfaces within LLZO-based SSLBs causes non-optimal material utilization and poor ion transport. Critical issues include volumetric changes in cathodes/anodes, Li dendrite growth, and structural defects in electrolyte, all of which cause electrochemical–mechanical degradation of electrolytes. Continuous degradation of the electrolyte causes deterioration in cell performance, leading to failure. Hence, thorough understanding of the kinetics of charge carrier transport, chemo-mechanical response of SSLBs to cycling, and dynamic external conditions (pressure/temperature) is necessary. The full-cell batteries could exhibit distinctly dissimilar reaction mechanisms, when LLZO-based SSEs are matched with different active cathode materials (such as intercalation or conversion). The limiting steps of the whole kinetic process could also vary with the system composition. In addition, although there is no obvious heat release from LLZO in contact with Li, short-circuited SSLBs can reach temperatures significantly higher than those of conventional LIBs, which could result in fire when in contact with flammable packaging or nearby materials.7 To be specific, compared with liquid batteries, the heat generation from ohmic polarization, activation polarization, and concentration polarization in SSLBs is expected to be aggravated due to high interface resistance. When heat transportation occurs within SSLBs, the anisotropic thermal diffusion could trigger uneven temperature distribution or localized hotspots. This will accelerate Li dendrite growth and initiate a short-circuit inside full-cell batteries, causing a rapid temperature increase. The continuous heat accumulation leads to inner structural damage, rapid performance degradation, and even thermal runaway (fire or explosion). Therefore, real-time and precise thermal monitoring of solid-state batteries is essential. Several phenomena, such as cell swelling and voltage fluctuation, are commonly observed prior to cell failure. However, accurately predicting the short-circuit failure of the cell remains an unresolved issue and can significantly reduce security incidents.

In summary, to reasonably design high energy density, stable, and safe LLZO-based SSLBs, it is necessary to comprehend the abovementioned issues or challenges. Accurately characterizing and diagnosing such critical issues by utilizing various in situ/operando techniques is essential for designing high-performance energy storage systems.

3. In situ/operando characterization and analytical techniques for LLZO-based SSLBs

As noted in the earlier section, LLZO-based SSLBs exhibit a variety of scientific and technical challenges under both quiescent and operating conditions. LLZO-based SSEs, especially CE, consist of a range of materials with varying physical and chemical properties, making experimental characterization challenging. The region which limits the performance of LLZO-based SSLBs is generally the electrode/electrolyte interface. This is referred to as the buried interface, which is hard to probe by conventional ex situ techniques. The internal composition, morphology, structure, and Li+/e transport channels of the composite cathode/Li anode are also hard to analyse with conventional investigation techniques and characterization tools. Therefore, utilizing in situ/operando techniques to characterize and diagnose such challenges is essential. Notably, almost all characterization techniques operate at a limited temporal and spatial resolution. As shown in Fig. 1, length-scales covered by LLZO-based SSLBs include the nano-scale (lattice structures and interphase formation), micro-scale (microstructural features and dendrite cracks) and meso-scale (composite cathodes and cell assembly). The individual components in SSLBs can have different crystal structures, densities, refractive indices, optical transmittances, and electronic conductivity. These differences dictate the effectiveness of a particular characterization technique. Constraints regarding sample geometry and working environments also dictate what techniques can be used to study a particular phenomenon. For example, imaging techniques are used to directly map material density and morphology. Reciprocal space techniques show the crystal structure, strain and phase of a material. Spectroscopic techniques measure the electrochemical state of a sample or probed region.

In recent years, researchers have developed a series of in situ/operando characterization techniques to detect SSLBs (Fig. 2). These characterization tools are intuitive, real-time, dynamic, damage-free, and enable direct real-time cell monitoring without disassembly. Table 1 summarizes the main material features/issues addressed by optical-, electron-, X-ray-, scanning probe-, magnetism-, neutron-based, and other in situ/operando characterization techniques in LLZO-based SSEs, composite cathodes/Li anodes, electrodes/LLZO-based SSE interfaces, and full-cell batteries. This section will discuss the recent developments and advances in these in situ/operando characterization techniques with relevance to LLZO-based SSLBs.


image file: d3mh00135k-f2.tif
Fig. 2 Recent timeline for in situ/operando characterization techniques for the development of LLZO based SSLBs.
Table 1 Summary of the main issues addressed by various in situ/operando characterization techniques in LLZO-based SSEs, composite cathodes/Li anodes, electrode/LLZO-based SSE interfaces, and full-cell batteries
Techniques LLZO-based SSEs Composite cathode/Li anode Electrode/LLZO-based SSE interface Full-cell batteries
Optical based techniques
OM Not yet applied Not yet applied Li dendrite growth; interlayer modification; interface reaction Not yet applied
Raman spectroscopy Thin film preparation Not yet applied Interlayer modification Electrochemical reaction
FTIR spectroscopy Not yet applied Not yet applied Not yet applied Not yet applied
Electron based techniques
SEM Not yet applied Not yet applied Li dendrite growth; interlayer modification Not yet applied
TEM Not yet applied Structure evolution Li dendrite growth; interface reaction Not yet applied
X-Ray based techniques
XRD Phase conversion mechanism Phase conversion mechanism Not yet applied Not yet applied
X-Ray imaging Porous structure; component distribution Not yet applied Li dendrite growth Thermal behavior
XPS Surface/bulk composition Surface composition Interface composition and reaction; interlayer modification Not yet applied
XAS Not yet applied Surface composition Interface composition and reaction Not yet applied
Scanning probe based techniques
AFM Li+ migration pathway Not yet applied Interface reaction; Li dendrite growth; interlayer modification Not yet applied
SECM Not yet applied Not yet applied Not yet applied Not yet applied
Magnetism based techniques
NMR Ionic transport characteristics; Li+ migration pathway Not yet applied Li dendrite growth Not yet applied
EPR Surface state Not yet applied Not yet applied Not yet applied
Neutron based techniques
NDP Li concentration distribution Not yet applied Li dendrite growth; interlayer modification; interface reaction Short-circuit mechanism
NI Not yet applied Not yet applied Li dendrite growth Not yet applied
Other techniques
ToF-SIMS Component distribution Not yet applied Interlayer modification; interface reaction Not yet applied
EIS Not yet applied Not yet applied Interlayer modification; interface reaction Not yet applied
Acoustic characterizati-on Not yet applied Not yet applied Li dendrite growth Not yet applied
Chemo-mechanical measurement Not yet applied Not yet applied Not yet applied Cell failure prediction; mechanical degradation


3.1 In situ/operando optical based techniques

3.1.1 In situ/operando optical microscopy (OM). Initially, the use of OM, an optical instrument, was widespread only in biology. It utilizes optical principles (such as reflection of light and amplification of convex lenses) to enlarge and image micron-sized objects that are not distinguishable to the human eye. OM exhibits the advantages of low cost, ready experimental operation, straightforward sample preparation, non-destructive test nature, and availability for measurement under non-vacuum conditions. Therefore, OM is widely applied in lithium battery investigations, such as monitoring Li dendrite growth or interface evolution.26 As shown in Fig. 3a, in situ OM operating systems generally comprise an OM, a computer, an electrochemical workstation, and a camera.27 Notably, as shown in Fig. 3b, OM typically has two configurations: in-plane and through-plane. The electric field lines of these two configurations may vary considerably and affect the electrochemical behaviour of cells.22 When the electrode thickness is lower than the electrolyte thickness/diameter, in-plane and through-plane configurations display approximate field concentrations at the electrode edges. When the electrode size is comparable to the electrolyte diameter, through-plane configurations with planar field lines show minimal edge effects. The inhomogeneity of the edge effects causes localization of the failure features around the viewing edge. Therefore, in contrast to real applications, thick electrolyte layers are normally employed in in situ operations to obtain reasonable signals.
image file: d3mh00135k-f3.tif
Fig. 3 (a) Schematic illustration of the in situ OM setup. (b) Schematic diagrams and field effects for in-plane and through-plane configurations. Reproduced from ref. 22 with permission from Elsevier. Schematic diagram of a micro-cell structure in (c) in-plane and (d and e) through-plane configurations for in situ OM viewing. Reproduced from ref. 22, 28 and 29 with permission from Elsevier.

Analogous to normal cells, the one designed for in situ OM visualization is composed of a working electrode, an electrolyte, and a counter electrode. The distinction is that it needs a window on one side of the lens, which is constructed of an optically transparent material such as quartz glass or polypropylene. Fig. 3c illustrates a schematic diagram of a micro-cell structure in an in-plane configuration for in situ OM viewing.28 Two Li electrodes are placed parallel at ∼4 mm distance to the SSE surface. Electron beam evaporation is used to deposit a 100 nm Cu layer as a current collector on an optical glass plate. The glass plate with the Cu layer is covered on the Li electrode to ensure good electronic contact. The micro-cell is sealed with epoxy resin. The two electrodes are connected to an electrochemical workstation for charging and discharging. OM in real-time observes the area between the two electrodes. Fig. 3d displays a schematic diagram of a micro-cell structure in a through-plane configuration.29 The SSE is sandwiched between the Li electrode and the electrode to be observed. Ni screens are placed on the Li electrode side to buffer the applied stacking pressure. Also, two electrodes are connected to an electrochemical workstation for charging and discharging tests. Fig. 3e presents a schematic diagram of another micro-cell structure in a through-plane configuration.22 Minor distinctions are that areas apart from the active LLZO surface are masked with Kapton tape to prevent air contamination and Ni foil is adhered to both sides. Notably, the materials for the OM test are transparent or translucent. Therefore, OM is a highly suitable technique for in situ analysis.

In situ OM has been used to view the Li dendrite growth in LLZO-based SSLBs. For LLZO pellets, Guo et al.28 first prepared transparent Ta-doped LLZO (LLZTO) pellets, and then viewed the Li growth behavior inside the cell directly by in situ OM. As shown in Fig. 4a, Li dendrites always penetrate from one electrode to the other at low currents (100 μA) or high currents (500 μA). The penetration of Li dendrites changes from a single direction to multiple directions with increasing current to maintain a large exchange current at the interface. Accordingly, the shape of Li dendrites varied greatly at different currents. The Li dendrites formed at high currents were similar to those in liquid cells. In addition, during the plating/stripping process, dead Li appeared inside LLZTO. In subsequent cycles, Li was preferentially deposited in dead Li regions.


image file: d3mh00135k-f4.tif
Fig. 4 (a) Li dendrite evolution at 100 μA and 500 μA. The scale bar is 1 mm. Reproduced from ref. 28 with permission from Elsevier. (b) Schematic diagram for different types of morphologies of Li penetration. Reproduced from ref. 22 with permission from Elsevier. (c) The optical evolution photographs of bare Li and SbF3@Li during Li deposition at about 4 mA cm−2 for 240 min. Reproduced from ref. 29 with permission from Elsevier.

Kazyak et al.22 revealed that Li penetration includes four morphologies, straight, branching, spalling and diffused (Fig. 4b), indicating that a single mechanism cannot solely explain Li penetration. However, before the cell short-circuit, these morphologies of Li can be reversibly cycled. Moreover, a comparable Li penetration behavior was also found in glassy Li3PS4 (LPS) SSE. Furthermore, the Li filament propagation rate was proportional to the current density. Under deep discharge at high current density, void formation, dewetting and reduction of the contact area were seen at the Li electrode. Importantly, as temperature increases, the interface resistance decreases rapidly and the critical current density (the highest current density that the cell can withstand before short-circuiting) increases. Consequently, higher temperatures suppress Li dendrite growth at high current densities.30 The working temperature of the cell is 195 °C which is just above the Li melting point, and Li changes from the solid to molten state. Meanwhile, the mechanical properties of Li change in a step-wise manner, resulting in a step increase in the critical current density. The internal Li anode displays pressure relaxation, suppressing Li dendrite formation.31

The Li dendrite growth mechanism was also investigated with in situ OM. Both single crystals and polycrystalline LLZO invariably contain surface defects. Initially, due to surface defects, Li deposits at local locations at the interface between LLZO and Li metal electrodes. Then individual points of the surface experience high local pressure, resulting in mechanical cracking. Li fills the cracks and spreads gradually to form Li filaments, ultimately penetrating the SSE and contributing to a cell short circuit. Furthermore, current is also concentrated on this low resistance path to transport, generating high localised heat. Owing to the Joule heating effect, partial Li filaments are melted into Li spheres that disperse in and outside the SSE.32 Therefore, the failure mechanism of brittle LLZO pellets is mainly Griffith-like rather than the shear-modulus criterion proposed by Monroe and Newman.33 It is essential to minimize LLZO surface defects to avoid Li dendrite growth. The glassy LPS SSE displays similar Li dendrite penetration mechanisms.34 To alleviate the Li dendrite formation owing to surface defects, Kim et al.35 revealed the function of artificial interlayer modifications via in situ OM. The kinetics of alloying and precipitation reactions using interlayer species remarkably alter Li distribution and deposition morphology. In this way, the interlayer performs a dual role of a dynamic buffer layer for Li redistribution and a matrix layer for facile Li precipitation.

The lithiation/delithiation process of SSLBs constructed with LLZO-based CEs was viewed with in situ OM. In tetraethylene glycol dimethyl ether (G4)-based gel electrolyte solidification, LLZO particles were deposited on the underside forming a hierarchical layered structure by gravity.36 In Li-symmetric cell structures, as Li+ was deposited on the underside, Li anodes maintained a smooth surface without dendrite formation over the whole cycling process. This benefit arose from LLZO particle layer protection, resulting in homogeneous Li deposition and inhibition of side reactions at the interface. In contrast, when Li plating is on the top side, the active deposition sites rapidly develop, causing heterogeneous Li deposition. Moreover, a self-adaptive structure electrolyte (SASE) structure can also maintain intimate interfacial contact during cycling.37 It is well-known that Li dendrite growth in cells assembled with poly(ethylene oxide) (PEO)-based polymer electrolytes promptly contributes to cell failure and short circuits. It is also hard to suppress Li dendrite generation by adding LLZO particles. Fortunately, Wang et al.29 sprayed SbF3 on Li foil surfaces to in situ form a Li/Li3Sb interlayer, improving the SSE/Li interface. As shown in Fig. 4c, it was revealed by in situ OM that Li was initially deposited in particle interstices of Li/Li3Sb interlayers, which resulted in higher local current density in interlayers. Subsequently, Li was homogeneously and densely deposited on the interlayer surface, which can be attributed to the certain electronic conductivity, high chemical diffusion coefficient, and high Li absorption energy of Li3Sb, as well as to the high ionic conductivity of LiF.

Another characteristic of OM that distinguishes it from electron microscopy techniques is its ability to recognize colors in the visible range. This can open up directions for researching specific electrode layers for lithiation, relevant interface reactions and heterogeneous electrochemical reactions via color variations. Cross-sectional imaging with in situ OM was performed to investigate the interface behavior between an active sulphur cathode and CE (LLZTO as a filler, PEO as a polymer and lithium bis(trifluoromethylsulfonyl) imide (LiTFSI) as a Li salt) at 65 °C.38 Originally at an open circuit voltage (OCV) of 2.60 V, the CE remains bright-white when viewed from the optical glass window and on discharging to 2.34 V, the brightness of CE color diminishes markedly. As the potential shifts to 2.25 V, the CE color transforms from white to light brown, and color evolution spreads from the cathode side to the whole CE. Such a behavior was attributed to the polysulfides (PSs) dissolved in CE. In a fully discharged state at 1.50 V, CE color transforms completely to deep brown. Subsequently, on charging back to 2.60 V, the brown CE color is still maintained, indicating that PSs stay in CE instead of returning to the S cathode. The irreversible reaction induces rapid capacity degradation and results in low Coulomb efficiency of SSLBs. The CE suffers severe deformation with continuous cycling, resulting in S cathode fracture.39 In addition, temperature variations strongly influence PS shuttling, CE irreversible volume changes, and Li metal volume expansion.38 Interestingly, CE containing LiTFSI-lithium bis(fluorosulfonyl) imide (LiFSI) binary salts can achieve self-regulation in terms of interface compatibility and side reactions of dissolved PSs.40

Nevertheless, in situ optical microscopy (OM) still faces several challenges. The most critical challenge is the limited spatial resolution, which cannot reveal the reaction mechanisms on a nano or an atomic scale. For such high magnification imaging, electron microscopy techniques are employed (discussed in Section 3.2). OM is unable to visualize opaque materials, allowing it to be used mostly in surface/interface research. Encouragingly, optical interferometric scattering microscopy (iSCAT) has been developed as a rapid and low-cost imaging technique and feature positions (such as phase boundaries) with “precision” less than 5 nm can be identified in iSCAT images. Unfortunately, iSCAT has not been utilized for analysing LLZO-based SSLBs. However, scientists have visualized and determined the ion diffusion kinetics in single LiCoO2 (LCO) particles by iSCAT.41 During charging and discharging processes, LCO particles showed heterogeneous iSCAT intensity distributions (determined from local dielectric properties of the samples), indicating uneven Li+ diffusion and phase transition.42

3.1.2 In situ/operando Raman spectroscopy. The Raman effect originates from inelastic scattering as monochromatic probe light interacts with materials. The resulting Raman spectroscopy provides time- and spatial-resolved information on vibrations, rotations, and other low-energy molecular/crystal bond modes on the material surface (Fig. 5a).43 Accordingly, detailed data about the sample's chemical structure, morphology, phases, crystallinity, and molecular interactions are available. Since this is a non-destructive analysis method, Raman spectroscopy allows a continuous in situ study of the sample. Importantly, compared to XRD, in situ Raman spectroscopy allows the analysis of amorphous or weakened crystalline materials, especially C, O, H and polysulfide species. However, in situ Raman fails to detect metal signals (such as Li metal) directly as it is sensitive to non-polar bonds and relies on energy level vibrations and rotations in molecular/crystal bonds. Fortunately, information about Li+ has been frequently researched indirectly with the aid of other species (such as C) peak position shifts.44
image file: d3mh00135k-f5.tif
Fig. 5 (a) Schematic illustration of Raman spectroscopy studies. (b) Schematic illustration of an in situ Raman testing cell. Reproduced from ref. 45 with permission from Springer. (c) Schematic illustration of an operando Raman microscopy cell for measuring the cross-section of the sample. Reproduced from ref. 46 with permission from the American Chemical Society. (d) Schematic illustration of in situ Raman spectroscopy of a thin-film battery. Reproduced from ref. 47 with permission from Elsevier. (e) Schematic illustration of a 3D-printed in situ symmetric cell device attached with a micro-heater. Reproduced from ref. 48 with permission from the American Chemical Society.

Fig. 5b illustrates a schematic diagram of an in situ Raman spectroscopy cell used to test the sample positioned on the top.45 The cell consists of a working electrode, an electrolyte, and a counter electrode. In situ Raman cell design requires an extra special unit, a window, to allow detected light to enter without obstruction. The window, generally a quartz glass, is placed on top of the cell structure and is sealed using glue at the point of contact with an Al plastic package. To observe the top region, a current collector with a small hole or a mesh current collector with pores must be used to expose the observation area. The Al tab and Ni tab are connected to an electrochemical workstation for charging and discharging tests. Fig. 5c shows a schematic diagram of an operando Raman spectroscopy cell for viewing the cross-section of the sample46 with an O-ring used to perform sealing. Fig. 5d presents a schematic diagram of an in situ cell for Raman experiments on cathodes in thin film cells (SSE thickness of 3–4 μm).47 Indium tin oxide (ITO) electrodes can be utilized as current collectors due to their transparency to laser light. As electrodes undergo large volume changes during the charging and discharging process, in situ/operando Raman spectroscopy cells should be highly air-tight to ensure a high pressure. In theory, in situ/operando cells should feature a high confining pressure similar to normal cells. Fig. 5e illustrates a 3D-printed in situ symmetric cell device attached to a micro-heater.48 For Raman microscopy, a flat observation area is necessary to obtain clear images of the samples. Thicker samples can be finely polished with SiC paper, while thinner samples or composite electrodes are needed for Ar ion milling.49 When conducting ion milling, the power of ion beam should be controlled to prevent damage to the samples. Therefore, electrodes, electrolytes, and the interfaces between them are accessible for investigation using Raman spectroscopy.

Initially, the annealing process of LLZO films was studied with in situ Raman spectroscopy. The Raman profiles indicate that the temperature range of 23°−700 °C was adequate for cubic LLZO generation since this system benefits from Li3N Li reservoirs that compensate for Li loss during low-temperature vacuum deposition.13 However, high background noise levels result owing to the long distance between the sample in the heater and the Raman microscope since only 50× lenses are used in this setup. It is appropriate to investigate the synthesis temperature of cubic LLZO utilizing in situ Raman spectroscopy. As shown in Fig. 6a, LLZTO pellets after cycling exhibit compressive stress regions and stress-free regions (originally tensile stress regions) when viewed by microscopic two-dimensional (2D) and 3D stress mapping methods utilizing Raman spectroscopy.50 The high tensile stress region with high overpotential induces Li deposition preferentially. Excess Li is deposited along the grain boundaries in LLZTO pellets and accumulates progressively, contributing to the emergence of compressive stress regions. While the development of stress-free regions is attributed to cracks inside LLZTO pellets and voids at interfaces, Li deposition along cracks and grain boundaries connects through the cathode and anode, resulting in cell short-circuits. Introducing an artificial interlayer at the SSE/Li anode interface can effectively homogenize Li+ deposition and suppress Li dendrite formation.9In situ Raman spectroscopy revealed that h-BN interlayer-modified LLZO-based SSLBs produce neither structural nor chemical variations during cycling.48 In contrast, unmodified LLZO-based SSLBs display structural phase transition from cubic to tetragonal LLZO, leading to enhanced interfacial resistance and cell degradation. It is well known that the Li2CO3 layer is generated rapidly on the LLZO pellet surface in air.51 However, the quantitative analysis of the Li2CO3 generation rate has not been adequately conducted. Thus semi-quantitative rate analysis using in situ Raman spectroscopy can be an excellent solution.


image file: d3mh00135k-f6.tif
Fig. 6 (a) 2D and 3D stress mapping for LLZTO before and after electrochemical cycling. Reproduced from ref. 50 with permission from Elsevier. (b) In situ Raman spectra of I2@KB cathode at different discharge and charge states for the first (left) and second (right) cycle. Reproduced from ref. 45 with permission from Springer.

The charging and discharging processes of CE-assembled SSLBs have been further researched with in situ Raman spectroscopy. In Li-LiFePO4 (LFP) cell systems, LLZTO encapsulated in polyacrylonitrile (PAN) eliminates decomposition of succinonitrile (SN) additives effectively.52 Conversely, in unencapsulated systems, SN decomposes severely and spreads throughout CE at higher voltages (>3.5 V), which generates a thick Li+-blocking layer (C[double bond, length as m-dash]N–C layer) at the interface. In Li–S cell systems, the re-emergence of S8 or S82− Raman peaks at 151, 217 and 471 cm−1 at 2.65 V confirms the occurrence of a reversible reaction mechanism.39 Notably, peak intensity post-cycling is weaker than that seen pre-cycling, indicating that S cathodes form PSs partially and dissolve in CE (as verified by in situ OM). In Li–I2 cell systems, the I3 Raman peak at 116 cm−1 apparently disappears in the first cycle, while the I5 Raman peak at 162 cm−1 becomes visible (Fig. 6b, left). This phenomenon suggests that the first cycle is irreversible while the in situ Raman spectroscopy study of the second cycle shows reversible conversion (Fig. 6b, right). Accordingly, a two-step reaction mechanism in all-solid-state Li–I2 cells with a cutoff voltage of 4.0 V has been proposed: (1) 3I5 + 2e ⇌ 5I3 and (2) 5I3 + 10e ⇌ 15I.45 Overall, these results show that studying the charge and discharge mechanism in a specific cell is practical using in situ Raman spectroscopy.

Besides LLZO-based SSEs, the sulfide SSE family has also been studied by in situ Raman spectroscopy. In Li |LPS| LCO cells, an unreacted LCO region occurs in the composite cathode.53 The poor contact between SSE and cathode active material generates low utilization and a heterogeneous distribution. Hence, finding the right strategy to ensure perfect contact between SSE and cathode active material is necessary. Moreover, the c-axis lattice parameter of LCO increases during cycling via peak intensity evolution from Raman spectroscopy.47 Importantly, Li+ diffusion in the cathode requires time, which results in a time-delay between electrochemical charge/discharge and Raman measurements. This is a difficult issue to avoid and address currently. In Se |LPS| C systems, Raman spectroscopy reveals the formation of a PS4−xSex3 interface phase.54 During the total charging and discharging process, the electrochemical reactions of Sen chains and PS4−xSex3 were discovered to be reversible despite the varying extent of lithiation. In addition, operando Raman spectroscopy reveals that the PS43− unit emerges inside LiNi0.8Co0.1Mn0.1O2 (NCM811) particles in NCM811 |Li6PS5Cl (LPSCl)| Li cells during long term cycling, suggesting SSE decomposition and diffusion and cathode particle deterioration.55

As discussed above, most research studies have focused on varying ion concentrations. However, conventional Raman spectroscopy features low spatial and temporal resolution, resulting in real-time failure to monitor ion concentrations.56 Stimulated Raman scattering (SRS), a nonlinear Raman technique, improves signal resolution and accuracy. Due to the utilization of two types of spatiotemporally synchronized picosecond laser pulse sequences, such technique achieves an effect 108 times stronger than conventional Raman spectroscopy.57In situ SRS successfully studied the variation of Li+ concentration and dendrite growth on the Li metal surface in liquid cells, allowed the 3D visualization of Li deposition and revealed the Li deposition procedure.58 Regrettably, in situ SRS is seldom used for LLZO-based SSLBs. Researchers are expected to utilize in situ SRS to further reveal the Li deposition procedure in LLZO-based SSLBs for optimal interface engineering. Besides low spatial and temporal resolution, low signal accuracy is an obvious weakness of conventional Raman spectroscopy. Micro-Raman, surface-enhanced Raman, and confocal Raman spectroscopy can enhance signal accuracy and it is expected that all these techniques will be adopted for LLZO-based SSLBs investigation in the future.

3.1.3 In situ/operando Fourier transform infrared spectroscopy (FTIR). FTIR spectroscopy is a vibrational spectroscopy technique. Over a wide spectral range (14[thin space (1/6-em)]300 to 20 cm−1), dipole movements are monitored by exciting molecular vibration, rotation, and lattice modes based on the absorption of infrared beams.59 The resulting absorption peaks correspond to various modes of the sample, which provide information on the composition and structure of molecules.60 FTIR spectroscopy is characterized by high detection sensitivity, high resolution, high measurement accuracy, and fast measurement speeds and it is ideally advantageous for determining the structure and functional groups of polymers and organic compounds. Therefore, in situ FTIR is vital for investigating electrolyte decomposition and solid electrolyte interphase (SEI) film formation in conventional Li-ion and nascent lithium batteries.61 Surprisingly, in situ FTIR is seldom used in the LLZO-based SSLB research area.

On one hand, the interface between the LLZO pellet and cathode/anode is mainly modified using inorganic materials while LLZO pellets are also inorganic in nature. In situ FTIR fails to achieve a desirable result due to its preference to analyze organic instead of inorganic materials. On the other hand, CE studies mainly concentrate on modulating the species, content and morphology of components, while rarely focusing on the CE degradation and SEI film formation at the interface with the cathode/anode. Consequently, researchers seldom use in situ FTIR characterization. Furthermore, FTIR spectroscopy is inadequate for investigating Li dendrite nucleation and growth. As organic modification of interfaces and degradation of CE will become increasingly researched in the future, in situ FTIR will become critical. The cell structure for in situ FTIR characterization is probably similar to that in in situ Raman spectroscopy. According to the experience of liquid cell research, FTIR is susceptible to O2 and CO2 reactions, resulting in low reliability of outcomes. Several infrared spectroscopy modes, including diffuse reflectance infrared transform spectroscopy (DRIFTS), double modulation FTIR (DM-FTIR), subtractive normalized interfacial FTIR (SNI-FTIR), photoacoustic infrared spectroscopy, and transmission infrared spectroscopy will be implemented in future SSLB research studies.59

3.2 In situ/operando electron based techniques

Due to the atomic and nano-scale imaging resolution, electron microscopy (EM) is very suitable for characterising solid-state batteries. EM imaging is performed by detecting a series of measured signals generated from the interaction between the electron beam emitted from an electron gun and the surface atoms of the test sample. The morphology and structure of materials can be determined from further analysis of such measured signals. With the development of high-angle annular dark field (HAADF) imaging, annular bright field (ABF) imaging, X-ray energy spectroscopy, electron energy-loss spectroscopy and electron holography (EH), the information on elemental distributions, elemental valence and local electric field is also available. In particular, in situ EM, a powerful characterization technique, facilitates extensive analysis of the morphological evolution and growth mechanisms in Li metal, the structural transformation of cathode active materials, and the evolution mechanism of electrode/electrolyte interfaces. Due to their ease and broad availability, scanning electron microscopy (SEM) and transmission electron microscopy (TEM) are typically used for these analyses.
3.2.1 In situ/operando scanning electron microscopy (SEM). SEM is one of the most potent tools in material research and is extensively used owing to its higher spatial resolution than OM. As shown in Fig. 7a, this technique provides numerous advantages: (1) high resolution (<1 nm); (2) large depth of field (∼mm), providing information for 3-dimensional imaging; (3) wide and continuously adjustable magnification, enabling examination of most material morphologies; (4) simple sample preparation; and (5) ease of combining with different accessories to achieve varied sample information such as energy dispersive spectroscopy (EDS) and electron backscatter diffraction (EBSD) to determine the elemental distribution and crystal orientation, respectively. To be specific, evolution of interface reactions can be studied by combining variations in particle size and morphology and EDS data before and after charging and discharging processes. As with OM, SEM offers in situ viewing in selected areas in the sample without disassembling the cell. To carry out in situ SEM characterization, a feedthrough leads the anode and cathode of the in situ cell to the outside of the SEM system through a flange hole on the SEM chamber. This is then connected to an electrochemical workstation to apply voltage to the cell. Meanwhile, the electrical signal is received and transmitted to a computer for real-time imaging. It is noteworthy that SEM demands a high vacuum (∼10−3 Pa) to stabilize the electron source to lower the background noise. Based on this, in situ viewing of SSEs systems containing elements with low vapor pressure is preferable to those containing conventional liquid electrolytes.62
image file: d3mh00135k-f7.tif
Fig. 7 (a) Schematic diagram for SEM studies. (b) Schematic diagram of a plan view in situ SEM cell structure. Reproduced from ref. 63 with permission from Springer. (c) Schematic diagram of the in situ SEM setup which is typically equipped with an FIB system. Reproduced from ref. 64 with permission from Wiley. (d) Schematic diagram of an SEM experimental setup equipped with a microelectrode module and two micromanipulators. Reproduced from ref. 62 with permission from Elsevier. (e) Schematic diagram of a cross-section view in situ SEM cell structure. Reproduced from ref. 65 with permission from Wiley.

Distinct in situ cell setups inside the sample room allow different viewing perspectives resulting in variable information. Fig. 7b illustrates a schematic diagram of a plan view in situ SEM cell structure.63 34 μm Li metal electrodes are pressed on both sides of the electrolyte. A smaller Li electrode is utilized on the top to induce dendrite growth to have edge effects. A Cu spring is employed as a contact electrode and to apply pressure on Li films to press them against LLZO electrolyte to achieve good contact. The lower Li electrode, as a contact electrode and with the same area as LLZO, is rested on a flat Al sample holder that pushes the lower Li electrode on the LLZO electrolyte. The electrode is connected to a cycler using an electrical feedthrough installed in one SEM port. The top Li electrodes can be substituted with Au or Pt electrodes.14,64 As shown in Fig. 7c, to observe correlated information inside electrolytes during electrochemical cycles, the in situ SEM setup is typically equipped with focused ion beam (FIB) technique for milling the surface.64 In addition, Fig. 7d shows an SEM experimental setup equipped with a microelectrode module and two micromanipulators.62 The micromanipulator attached to a tungsten needle is directly placed on the LLZO pellet surface as a working electrode, while that attached to Li metal acts as a counter electrode. The micromanipulator enables external operation in x, y and z directions under high vacuum conditions using a control unit. The electrically-shielded connections for microelectrode and micromanipulator control units are guided through a channel outside of the SEM chamber. Adopting a comparable SEM experimental setup, Fig. 7e shows a schematic diagram of a cross-sectional view of the in situ SEM cell configuration.65 The LLZO pellet is fractured manually into two semicircular pieces; one piece is placed vertically in a homemade sample holder and fixed with an electrically conductive grub screw. One side of the fractured LLZO pellet is pasted with a Li electrode acting as a counter electrode and a reference electrode, which is electronically contacted via a grub screw. Another side is directly in contact with the Al frame. A Li|LLZO half-cell is fixed in a customized sample holder for real time visualization. Notably, another side of the fractured LLZO pellet can be attached with an In electrode or Si electrode for research purposes.66,67

Since the experimental effect of electron beam on Li (as compared to Na/K) is relatively low, in situ SEM technique is frequently adopted to characterize the growth morphology and formation mechanism of Li dendrites on a micro-scale. At first, in situ SEM was used to view the LLZO/Li interface in Li symmetric cells during cycling.63 The results suggest that the LLZO/Li interface characteristics are crucial for Li dendrite growth. At the interface, uneven dissolution of Li causes inhomogeneous current distribution and Li deposition. The generated Li dendrites have three morphologies: mossy, needle, and bump (Fig. 8a). Both bump and mossy features were seen mostly in thin Li metal areas due to the low exerted stress. Moreover, combining SEM and EDS data revealed that the formed Li dendrites contain C and O, which is probably related to Li2CO3, LixCy and Li2O. An in situ SEM device with micromanipulators was adopted to conduct and analyse electrodeposition experiments.62 Such a device reveals the inhomogeneous electrodeposition behavior and Li dendrite and whisker growth (Fig. 8b), which were attributed to the microstructural variations (including one- and two-dimensional defects68) of the LLZO substrate. In theory, a stable and clean LLZO/Li interface exhibits ultrafast charge transfer kinetics (<10−1 Ω cm2).65 Ideally, the transport limitations derived from vacancy injection and diffusion are the exclusive kinetic parameters fundamentally restricting the deposition rate capability. However, inherent defects such as porosity, grain boundaries and impurities are inevitable in the practical polycrystalline SSEs. These defects results in rapid Li deposition kinetics and high nucleation trends, resulting in non-uniform Li deposition.14 In general, short, and thick whiskers are developed at low current densities, while long and thin structures are seen at high current densities. These phenomena reflect that the surface morphology, current density, and surface crystal nucleation kinetics completely control Li's nucleation and deposition behavior. Therefore, minimizing inherent defects to achieve a clean and defect-free interface can prevent the development of local nucleation hotspots. In addition, during the preparation of LLZO pellets, it is extremely probable that microcracks are introduced on the surface. As current is applied, inhomogeneous electric field intensity will be generated around the microcracks. As shown in Fig. 8c, cross-sectional in situ SEM reveals that Li dendrites grow along the cracks and ultimately penetrate LLZO.66


image file: d3mh00135k-f8.tif
Fig. 8 (a) SEM images of the cell at various cycle times. Reproduced from ref. 63 with permission from Springer. (b) SEM images of dendritic-like and whisker-like Li-metal deposition. Reproduced from ref. 62 with permission from Elsevier. (c) SEM images during the cycles. The scale bars are 200 μm. Reproduced from ref. 66 with permission from Elsevier. (d) Time-lapse images showing the formation of a bowl-shaped crack. Reproduced from ref. 64 with permission from Wiley.

Furthermore, an in situ SEM device equipped with an FIB system revealed that Li propagates predominantly along transgranular cracks.64 The crack propagation initiates from the interior of LLZO beneath the electrode surface and then propagates by curving toward the surface, forming a bowl-shaped crack (Fig. 8d). The resulting crack resembles hydraulic cracks caused by high fluid pressure, indicating that the Li deposition-induced inner stress is the major driving force for crack initiation and propagation. Chemical etching or laser treatment of SSE surfaces can realize crack-free surfaces, resulting in homogeneous electric field.

In fact, the pure LLZO pellet surface is not completely clean and defect-free. Hence, an artificial interlayer is introduced to achieve homogeneous Li deposition. The capability of the interlayer to suppress Li dendrites was directly visualized by in situ SEM.62 The Au layer can be alloyed with Li, and because of this, Li metal penetration is delayed. However, after generation of the alloy phase, Li metal still nucleates and penetrates at the interface. Due to the inhomogeneous plating characteristics and the high nucleation overpotential of Cu layers, Li metal penetrates immediately. Therefore, the alloyed interlayer will not fundamentally change the kinetics of the Li anode. Li nucleation will not occur if the alloy's vacancy diffusion coefficient can enable Li fluxes to exceed the external applied current density. It is advantageous to introduce a Li metal reservoir layer or an alloy layer with fast Li diffusion properties at the LLZO/Li interface to assemble LLZO-based SSLBs. Moreover, an electronically insulating LiF layer effectively improves the physical interface contact and suppresses the Li dendrite formation, as revealed by in situ SEM.69 Serial block-face scanning electron microscopy (SBFSEM) can image the whole bulk material morphology especially the interior. Imaging of Li and g-C3N4 composite anodes by SBFSEM revealed that Li3N is enriched on the surface.70 It is comparable to introducing a Li3N layer at the LLZO/Li interface, resulting in homogeneous Li deposition. Analogously, introducing a C3N4 interlayer at the Li1.5Al0.5Ge1.5(PO4)3 (LAGP)/Li interface not only homogenizes the Li+ flux and alleviates the stress induced by uneven Li deposition, but also restrains the interface side reactions between LAGP and Li.71In situ SEM demonstrated that the absence of C3N4 interlayer causes successive generation of side reaction layers and mechanical cracks at the interface. In general, utilization of non-Li anodes essentially suppresses Li dendrite formation. The interface compatibility and stability between Si anodes with various thicknesses and LLZO was investigated via in situ SEM.67 Excellent interfacial contact and electrochemical performance were realized for Si anodes with thicknesses less than 180 nm. Unfortunately, as the Si anode thickness increases, the cell performance deteriorates. Recently, the in situ SEM technique has also been critical in cathode failure research. It is verified that the Li3BO3 (LBO) and LCO interfaces in the composite cathode are cracked as the cycle evolution occurs, resulting in higher interface resistance and cell degradation.72

Importantly, adding various equipment in the in situ SEM sample chamber can realize diverse in situ property measurements. For example, coupling a SEM system with mechanical test systems co-located in a large vacuum chamber can help conduct in situ experiments on mechanical properties.73 Based on this system, the variability of mechanical properties during Li dendrite growth can be adequately investigated. When an organic solvent (e.g., CE systems or interface modifications) or low melting point Na/K metals (e.g., construction of composite anodes) are introduced in the investigation system, environmental SEM (ESEM) is feasible since the sample chamber allows working with wet samples. However, the resolution of ESEM is inferior to that of SEM (∼10 nm at a low vacuum of <103 Pa). Notably, numerous artefacts, including surface damage and alkali metal melting, are observed to arise at room temperature during cross-section viewing using FIB techniques. For example, SEI films formed in liquid cells are susceptible to high-energy Ga+, resulting in incorrect detection of the elemental distribution. To address this phenomenon, cryo-FIB/SEM technique has been successfully applied for spectroscopy and imaging on a nano-scale in liquid cells.74 Such a technique can be further employed in LLZO-based SSLB systems. In that case, several surprising results have been seen, such as natural formation of nano-scale interlayers at the interface or on the Li surface that exhibits heterogeneity influencing Li deposition kinetics.

3.2.2 In situ/operando transmission electron microscopy (TEM). TEM is a powerful instrument for imaging on a nano-scale. As shown in Fig. 9a, the working principle of TEM is similar to that of OM. The main distinction is that it uses electron sources rather than light sources. In specific, an accelerated and focused electron beam (energy of 60–300 keV) is projected onto an extremely thin (<150 nm) sample for sub-nanometer analysis. The electrons collide with the sample atoms to alter the orientation, resulting in stereo-angular scattering. Since the magnitude of the scattering angle is affected by the sample density and thickness, images with various brightness and darkness levels can be displayed on imaging devices (such as fluorescent screens). Through the use of a spherical aberration (Cs) corrector and chromatic aberration (Cc) lens, the spatial resolution of TEM is reached to a sub-angstrom level.75 The condenser lens can be adjusted to generate a parallel beam for the normal TEM mode or a point beam for the scanning TEM (STEM) mode. In the TEM mode, the sample structure and crystal structure defects are investigated by observing the transmitted and diffracted electrons, respectively. STEM can be considered as a combination of SEM and TEM. The electron beam is scanned on the sample surface, and images are derived from the partial electron penetration of the sample. Furthermore, the elemental analysis of materials can be performed using energy-dispersive X-ray spectroscopy (EDX) and electron energy loss spectroscopy (EELS). Therefore, TEM is able to analyze critical messages on a nanoscale or even on an atomic-scale in LLZO-based SSLMs, such as microstructural evolution, phase transitions, chemical composition variations, surface/interface evolution, and reaction kinetics.
image file: d3mh00135k-f9.tif
Fig. 9 (a) Schematic diagram for TEM investigations. (b) Schematic diagram for STM-TEM holder and MEMS-TEM holder. Reproduced from ref. 10 with permission from Elsevier. (c) Schematic illustration of the anode-free Cu|LLZO|Li nanobattery setup for in situ TEM probing. Reproduced from ref. 77 with permission from Springer. (d) Schematic illustration of chip-based holders and batteries, an SEM image of the chip, and the step-by-step schematic illustration of the battery fabrication process. Reproduced from ref. 79 with permission from the American Chemical Society. (e) Configuration of TEM holder and micro-battery. Reproduced from ref. 19 with permission from Springer.

For in situ TEM measurements, it is critical to design miniaturized in situ electrochemical cells because of the demands of high vacuum environments and extremely thin samples.59 Various in situ TEM holders have been developed to research electrical, thermal, magnetic, and force properties in various gas and liquid environments. As shown in Fig. 9b, these holders are mainly divided into scanning tunneling microscopy-TEM (STM-TEM) holders and micro-electro-mechanical system-TEM (MEMS-TEM) holders. STM-TEM holders utilize a scanning probe unit at the end of the sample holder and incorporate current–voltage measurements for in situ operation and electrical measurements. Individual nano-wires or nano-particles generally serve as the electrodes under observation, materials with low saturation vapour pressure serve as electrolytes, and cathode materials or Li serve as counter electrodes.76 Since the reaction cell in the holder is always open, the electrolyte and electrodes are not sealed, so the electrons directly interact with the observed part. Fig. 9c illustrates a schematic diagram of an in situ cell setup with an STM-TEM holder.77 LLZO particles are first distributed on the glass substrate. Then, a Cu rod with Li metal attached is pressed onto the substrate, enabling LLZO particles to be attached and submerged in Li metal. The whole part is mounted on one end of the holder. Li metal adhered to a Cu rod acted as a Li source and LLZO particles semi-submerged in Li metal served as SSEs. The other end of the holder is mounted with a metal probe, namely, Cu, sender W or Cu attached with carbon nanotubes (CNTs) (providing different mechanical constraints), as a working electrode. The working electrode is brought into contact with a LLZO particle to form a solid-state nanocell, realizing a cross-sectional view of the interface. A constant bias is applied to the Li metal to initiate the electrochemical process. Accordingly, STM-TEM holders have the advantages of simple sample preparation and atomic-level spatial resolution. Unsurprisingly, there are several issues as well. First, there is a weak contact between the electrolyte and electrode; second, it is essential to apply a large over-potential for driving Li+ transport due to the low ionic conductivity of the electrolyte, and lastly, there is a difference between the working environment and actual batteries and this can lead to incorrect conclusions.

MEMS-TEM holders conduct in situ electrical measurements and other functions by carrying a MEMS chip in the sample holder. They rely on contact between conductive probes and electrical contacts of pre-patterned circuits on silicon wafers. The MEMS chip equipped with a nanometer-thick robust Si3N4 window results in a sealed environment.78 Nanoparticles or active material lamellae (prepared by FIB) connected to a Pt electrode serve as observed electrodes, while Li or Au is the counter electrode. Fig. 9d illustrates a schematic diagram of the holder, chip, and in situ electrochemical cell.79 First, a lamella is cut from a micron-sized LCO particle by using FIB milling. The LCO lamella is connected to a gold wire by Pt deposition on the nano-chip. The gold wire is parallel to the [001] direction of the LCO lamella. Subsequently, an LLZO electrolyte slice and Au anode are cut out. A whole cell is formed using a Pt deposition connection. An ion beam is utilized to cut Pt deposition on the electrode in case a short circuit occurs during the charging process. Notably, one or two FIB cuts are conducted on the LCO cathode to thin the sample, which makes an electron-transparent area for in situ TEM observation. The two ends of the microcell are connected to two electrodes in the nano-chip using silver paste (Fig. 9e).19 Lastly, the nano-chip is placed in the in situ bias sample holder. The two electrodes are connected to an electrochemical workstation via lead wire for charge and discharge tests. Hence, MEMS-TEM holders not only compensate for the disadvantages of STM-TEM holders but also ensure atomic-level resolution. However, in situ electrochemical cells for MEMS-TEM holders are tough to fabricate.

Research on LLZO-based SSLBs has been conducted with a series of in situ electrochemical cells developed using the two holders described previously. When Al-doped LLZO (Al-LLZO) comes in contact with Li metal, an ultrathin (∼6 nm) interphase is viewed depending on the STM-TEM holder.80 The interface was observed to be tetragonal LLZO, as confirmed by STEM mode and EELS analyses. This phenomenon was attributed to the local phase conversion that occurred due to the slight reduction of the Al-LLZO surface in contact with Li metal. Notably, the interface was stable and avoided the complete reduction of Al-LLZO by Li metal and enabled Li+ transportation. In contrast, cracks, deformation or phase conversion are undetected when Al, Nb co-doped LLZO nanoparticles come into contact with Li metal.81 In fact, the chemical/electrochemical stability between LLZO and Li metal is dependent on the LLZO-doped elements.9 Subsequently, the in situ TEM holder can be used to investigate the evolution of the LLZO/Li metal interface.77 Under stronger mechanical restraints and lower charging rates, LLZO particles clearly display a “splitting” or “peeling” form (Fig. 10a). These single-crystal LLZO particles are free of apparent defects, indicating that the local stresses can reach GPa levels during Li “eruption”. Once the mechanical restraints are weakened, LLZO will not be damaged even if the local current density is increased to A cm−2 level. Therefore, the crack initiation at the interface is associated with the route and efficiency of mass/stress release. Rapid mass/stress release can be realized by either accelerated Li+ transport in carbon nanotubes or increasing temperature to facilitate Li0 diffusion. Moreover, Li+ heterogeneous accumulation at the LLZO/Li interface during the charging process is revealed by EH and EELS.82 Even after the discharge process, Li+ is seen to be present. Besides the LLZO/Li interface, the interior of LLZO was also studied using in situ TEM. The grain boundaries with narrow band gaps (∼1–3 eV) in LLZO are potential leakage channels.83 When 10 V bias voltage is applied, the brightness of the triple-junction voids progressively weakens to a level comparable with surrounding grains as a function of time. The decrease in brightness is attributed to the filling of Li into voids. As a result, Li+ is precociously reduced by electrons to form Li filaments at the grain boundaries, and these Li filaments are ultimately interconnected, causing short circuits. In summary, the electronic conductivity of grain boundaries should not be neglected.


image file: d3mh00135k-f10.tif
Fig. 10 (a) Crack initiation and Li penetration in LLZO by Li eruption at the constrained Cu|LLZO interface. Reproduced from ref. 77 with permission from Springer. (b) In situ STEM micrographs of the LCO cathode structure variations. The Li, oxygen, and cobalt ions are green, purple, and cyan balls, respectively. Reproduced from ref. 79 with permission from the American Chemical Society. (c) Changes in the Li maps at 3–18 nA h, B 39–53 nA h, and the open-circuit state for 30 min between the charge and discharge reactions, respectively. The scale bars are 100 nm. Reproduced from ref. 19 with permission from Springer.

MEMS-TEM holders are regularly used to research the cathode material structure evolution in SSLBs during charging and discharging processes. At high voltage delithiation, primary single-crystal LCO cathodes are transformed into nano-sized polycrystalline phases connected by coherent twin boundaries and anti-phase domain boundaries (Fig. 10b).79 Such transformations could be preferable for Li+ migration. For LiNi0.5Mn1.5O4 (LNMO) cathodes, transition metal ions, in particularly Ni ions, migrate to the 4a sites (16c in the Fd[3 with combining macron]m space group), leading to structural stability reduction.18 The 〈100〉, 〈110〉, and 〈111〉 zone axes are converted from ordered to disordered structures. The 〈112〉 zone axis produces inhomogeneous structural evolution forming three regions: a transition metal-rich region, an anti-phase boundary region, and a transition metal migration front region. Fortunately, the analysis revealed that the improvement of stability can be realized by doping low-valence cations on an atomic scale. It is expected that the structural transition of commercially available LFP and LiNixCoyMn1−xyO2 (NCM) cathode materials during charging and discharging will also be investigated in future research. The reaction kinetics of Li+ in LCO cathodes with LAGP SSE was studied by operando EELS and sparse coding (image de-noising, improved imaging time and spatial resolution).19 Importantly, Li+ migrates along the vertical direction to the LCO/LAGP interface and along the parallel direction (Fig. 10c) and also migrates to the interior of LCO even in the open-circuit state. Moreover, in situ differential phase contrast-STEM (DPC-STEM) technique revealed that discontinuously coating BaTiO3 (BTO) nanoparticles on LCO particles can suppress space charge layer (SCL) formation and facilitate Li+ migration.84

Besides imaging at room temperature, several investigations demand in situ imaging at high or low temperature. For instance, the LLZO phase evolution investigation during calcination demands high temperature imaging.85 Since Li/Na metal and SEI films are sensitive to extremely high electron doses in high-resolution imaging, imaging at low temperatures is necessary.86 Accordingly, it is essential to utilize several heating and/or cooling techniques to realize the desired experimental temperature. Current heating technologies involve furnace-type holders,87 MEMS-based heating holders,88 filament wire holders,89 and laser heating holders.90 Unfortunately, these techniques need to address image drift to satisfy the requirement of continuous imaging during heating. MEMS-based heating holders can deliver a drift of less than 1 nm min−1 under optimal working conditions. Based on this, the phase conversion of Ga-doped LLZO (Ga-LLZO) during high-temperature calcination was revealed.85 As shown in Fig. 11, initially, LZO is formed according to a 2-to-1 epitaxial growth process along the crystallographic orientations of (−111)LZO//(−111)ZrO2 and [211]LZO//[101]ZrO2 at 750 °C. Then, Li and Ga are diffused into LZO layer by layer along the [011] orientation to form Ga-LLZO at 900 °C. Cooling technologies include cooling holders (the cold source is liquid nitrogen or liquid helium) and MEMS-based thermoelectric cooling chips (similar to the heating holder).91 The lowest temperature of cooling holders is as low as liquid nitrogen or liquid helium temperature but fails to stabilize at any temperature in this cooling range. The flow or boiling of liquid nitrogen or liquid helium also introduces vibration that causes drift. MEMS-based thermoelectric cooling chips exhibit the advantages of vibration-free and precise temperature control. Thanks to the cooling techniques, cryo-TEM realizes high-resolution visualization on electron beam-sensitive materials. However, electrochemical studies of in situ cryo-TEM are not available due to space constraints of TEM holders, complexity of chip function, and others. Notably, an initial investigation of LLZO-based SSLBs has been carried out with ex situ cryo-TEM technique.92 Results indicate that microcracks which originate from Li dendrite growth are introduced by mechanically polishing the LLZO surface to eliminate the contaminant layer. An interlayer consisting of amorphous carbon and nanocrystalline LiF is formed in situ by the introduction of CFx and molten Li. Such a layer exhibits superb lithiophobic nature and thus the capability to suppress Li dendrite formation, providing an effective strategy for enabling the use of LLZO-based SSLBs in energy storage applications.


image file: d3mh00135k-f11.tif
Fig. 11 In situ atomic-scale investigation of Ga-LLZO Li+-conducting solid electrolyte during calcination-related growth. Reproduced from ref. 85 with permission from Elsevier.

Overall, in situ TEM has been successfully used to investigate the interface, electrolyte and cathode in LLZO-based SSLBs. Notably, several issues remain: (1) in situ electrochemical cells are detached from practical working situations due to the limitation of the sample chamber; (2) extremely thin samples enhance the difficulty of cell preparation; (3) high-vacuum testing conditions increase the difficulty of laboratory manipulation; (4) sensitivity of Li metal to high electron doses produces unreliable results, thus requiring cooling techniques or low electron doses (without sacrificing resolution); and (5) lower imaging speed and higher image drift. It is believed that after addressing these issues, LLZO-based SSLBs can be more thoroughly and comprehensively investigated by utilizing in situ TEM techniques.

3.3 In situ/operando X-ray based techniques

X-Rays have been employed to reveal the variations in the phase, structure and chemical information in materials at various spatial and temporal length scale levels through interactions with the sample (such as absorption, scattering, secondary electrons, and others).59,93 The intensity of conventional X-rays is not high enough to penetrate several encapsulated materials such as coin cell shells. In this case, an additional transparent window is constructed for X-ray penetration. The materials of transparent windows are generally beryllium, Kapton tape, or polymer films. However, such window materials lower the resolution and precision to a certain extent. With the advent of synchrotron radiation (SR) technology, X-rays can have high flux and high intensity over a broad spectrum. Hence, SR X-rays can penetrate various materials without any damage, which prevents the need to modify the in situ cell window. The characteristics of SR X-ray technology include (i) high brilliance and energy tunability, providing high structural resolution; (ii) high temporal resolution, offering rapid capture of dynamic information; (iii) spatial resolution ranging from a nano- to micro-scale, delivering the potential for direct visualization and 3D reconstruction; and (iv) in situ/operando work, enabling the ability to monitor during (electro)chemical reactions continuously. In recent years, SR X-ray techniques involving X-ray diffraction (XRD), X-ray absorption spectroscopy (XAS), X-ray microscopy, and X-ray photoelectron spectroscopy (XPS) have been increasingly popular in battery research. This section will introduce X-ray and SR techniques in LLZO-based SSLB investigations.
3.3.1 In situ/operando X-ray diffraction (XRD). Due to its non-destructive nature, along with high accuracy, high speed, and absence of contamination sources, XRD technique is widely used in the research of material phases, lattices and volume variation (Fig. 12a). In situ XRD can identify some metastable information that is beyond the ability of ex situ XRD. Fig. 12b illustrates a schematic diagram of an in situ XRD cell.94 It consists of three main parts: a stainless-steel cell body, an X-ray-transparent window and a working electrode. Meanwhile, components such as polytetrafluoroethylene sheaths, springs, and gaskets are also needed. In the case of LLZO-based SSLBs, the separator and liquid electrolyte in Fig. 12b are replaced with a SSE. The X-ray-transparent window is the most critical part, which allows X-rays to reach the electrode. The window is normally made of beryllium metal, Mylar membrane, Al metal or quartz. Among all, beryllium metal exhibits excellent conductivity and is thus used the most. However, the beryllium peak is too strong in XRD patterns.
image file: d3mh00135k-f12.tif
Fig. 12 (a) Schematic diagram for XRD studies. (b) Schematic diagram of an in situ XRD cell. Reproduced from ref. 94 with permission from Wiley. (c) Schematic diagram of an operando XRD cell setup and cell inside a diffractometer. Reproduced from ref. 95 with permission from Elsevier. (d) Schematic diagram of an in situ XRD measure setup for phase conversion analysis during high-temperature heating. Reproduced from ref. 96 with permission from Elsevier.

However, beryllium metal is easily oxidized and produces beryllium oxide peaks that disturb the test results. Kapton films suffer from having lower conductivity although they have no sharp XRD peaks. Al metal windows are not frequently adopted due to strong X-ray absorption and the formation of alloy with Li below 0.45 V. Be metal windows are frequently used in LLZO-based SSLBs systems. The thickness of the window has to be optimal to ensure that the smooth transmission of X-rays is not hindered (high thickness) while ensuring that it is not easily fractured during assembly (low thickness) of in situ cells. If using an SR light source, smaller X-ray-transparent windows can be used and in some cases, a window is not even required. Hence, two conditions should be considered for in situ XRD cell design: first the cell should be simple to assemble/disassemble and reusable and second, an ideal window material should enable X-rays to reach electrodes easily and be stable in cell systems. Fig. 12c illustrates a schematic diagram of an operando XRD cell setup and cell inside a diffractometer for solid-state batteries.95 The cell is made of Ti and polyether ketone. A thin Li foil is placed at the bottom and covered with SSE. Then, a self-standing film of electrode material is placed at the top covered with a window (beryllium metal). The window is compressed with a rubber O-ring to seal the cell. The in situ cell structure design discussed above is suitable for the charging and discharging process. However, most current research reports are focused on changes in the LLZO-based SSE bulk or cathode structure during high-temperature heating. Fig. 12d presents a schematic diagram of an in situ XRD measurement setup for phase conversion analysis during high-temperature heating.96 A bulk or powder sample is placed in a crucible, which is inserted in a heating chamber. The sample is heated to a specific temperature, and XRD patterns are recorded before moving to the next temperature. The final in situ XRD pattern during heating is acquired. Therefore, during charging and discharging or high-temperature heating, in situ XRD techniques enable real-time visualization of material phase conversion and structural transformation. In addition, the working principle of SRXRD is analogous to that of conventional XRD employing laboratory X-ray sources. It can help determine crystallographic information, including lattice parameters, site occupancy, microstructure and strain/stress of many materials.93 High resolution, robust signal strength and low test time of SR are more favorable for measurement of dilute phase scattering, residual stress analysis, time-resolved XRD, and for in situ/operando XRD.93

The structural evolution of Sn anodes in initial lithiation/delithiation was investigated using operando XRD.97 During lithiation, Li–Sn alloy peaks appear, matching the expected Li2Sn5 phase. This phase is converted into the LiSn phase before the end of lithiation. In contrast, during delithiation, the LiSn phase disappears and the Li2Sn5 phase reappears. However, the latter does not disappear at the end of delithiation, indicating irreversibility of the cycle. The results indicate that the formation of an irreversible Li2Sn5 phase during charging and discharging process results in poor electrochemical performance. Currently, in situ XRD is mainly adopted to examine the phase conversion process of LLZO or cathode materials in LLZO-based SSLBs research.98 During annealing, the evolution of undoped LLZO, Ga-LLZO and Al-LLZO films is revealed by in situ grazing incidence (GI) XRD.99,100 The in situ cell employed for GI-XRD is placed under a flowing O2 atmosphere. As shown in Fig. 13a, an undoped amorphous LLZO film is formed initially as Li4Zr3O8 and Zr3O at 300 °C, while the tetragonal LLZO phase occurs at 500 °C. At 500°–700 °C, the tetragonal LLZO phase and the cubic LLZO phase are seen to co-exist. As the temperature continuously increases, the cubic LLZO phase gradually becomes dominant. Finally, LLZO is completely converted from the tetragonal phase to the cubic phase at 700 °C. The temperature for complete conversion is lowered to 650 °C through Ga doping.99 At 650°–700 °C, Al-LLZO films also fully showed cubic phase formation.100 Therefore, Li-site element doping effectively decreases the temperature of complete conversion of LLZO from the tetragonal to cubic phase, indicating that doping can stabilize the cubic phase at lower temperatures. Furthermore, phase evolution of LLZO particles during synthesis was also investigated by in situ high-temperature XRD. As shown in Fig. 13b, the calcination process of LLZTO and Ca/W dual-doped LLZO (LLCZWO) are similar.101 The whole procedure consists of three reaction stages, namely, the formation of La(OH)3 (25° to ∼300 °C), La2O2CO3 with unreacted ZrO2 (∼300° to ∼760 °C) and cubic LLZO (∼760° to 820 °C). Notably, the LLZTO sample shows an additional reaction from Ta2O5 to LiTaO3 compared to LLCZWO in the second stage. Operando synchronous XRD demonstrates that Al-LLZO particles are developed in a cubic phase at 900°–1000 °C with remarkable particle coarsening.102 The simulation results reveal that particles with small size and bimodal distributions are preferable for subsequent sintering densification. Besides the use of different doping elements to change the phase conversion behavior during LLZO particle synthesis, the heating technique can also be a factor that introduces changes. The peak signal of La2Zr2O7 (LZO) (2θ = 3.54°) is seen to rapidly increase in intensity and sharpness until its disappearance at 600° to 1065 °C when a reactive flash sintering (RFS) technique was used.96 In contrast, LZO disappeared progressively in conventional heating experiments. This phenomenon suggests that RFS facilitates intermediate LZO growth during the heating process.


image file: d3mh00135k-f13.tif
Fig. 13 (a) Phase evolution of cosputtered film LLZO with Li2O. Reproduced from ref. 99 with permission from the American Chemical Society. (b) Contour plots of 2D in situ high-temperature XRD patterns of LLCZWO and LLZTO. Reproduced from ref. 101 with permission from the American Chemical Society. (c) Contour plots of in situ high-temperature XRD patterns of a composite mixture of NCM523 charged to 3.8 V with Ga-LLZO in the selected 2θ range under air and N2 flowing conditions. Reproduced from ref. 105 with permission from Wiley.

It is well known that LLZO on exposure to air not only forms Li2CO3 and LiOH contaminants on the surface, but also transforms partially into Li7−xHxLa3Zr2O12 in the bulk. In order to recover cubic LLZO, high-temperature heat treatment is an effective way as confirmed by in situ high-temperature XRD.103 Under an air atmosphere, high-temperature heat treatment consists of five stages: (i) Li7−xHxLa3Zr2O12 + xLiOH = LLZO + xH2O; (ii) Li7−xHxLa3Zr2O12 = LLZO + LZO + (x/2) H2O; (iii) LZO + Li2CO3 = LLZO + CO2; (iv) LLZO = LZO + Li2O (volatilization); and (v) LLZO + H2O + CO2 = Li7−xHxLa3Zr2O12 + LiOH + Li2CO3.

Although large amounts of Li7−xHxLa3Zr2O12 are recovered to cubic LLZO, the volatilization of Li2O causes stoichiometric irreversibility during the heat treatment. The degree of stoichiometric irreversibility is dependent on the kind of doping additive. Furthermore, re-protonation reactions occur in the cooling process. For preventing such reactions, heat treatment in an argon atmosphere is demonstrated to be useful by in situ grazing incidence synchrotron XRD (GISXRD).104 The cell performance of LLZO with heat treatment is superior to that seen when air is used. It is noteworthy to avoid re-exposure of heat-treated LLZO to air which induces protonation reactions. In situ high-temperature XRD also revealed the chemical compatibility and structural stability between charged LiNi0.5Co0.2Mn0.3O2 (NCM523) and Ga-LLZO when co-sintered at high temperatures in various atmospheres.105 As shown in Fig. 13c, Li diffuses from Ga-LLZO to NCM523 with increasing temperature under an air atmosphere. Ga-LLZO is decomposed to LZO and charged spinel NCM523 is recovered from being layered. In contrast, Li diffuses from NCM523 to Ga-LLZO under a nitrogen atmosphere. Ga-LLZO is converted from the cubic to tetragonal phase and NCM523 is reduced to Ni metal. Therefore, co-sintering technique is inappropriate for improving the poor contacts between layered cathodes and LLZO.

X-Ray scattering occurs in the lattice of regularly spaced crystalline materials and this can be used to obtain information about the grain size, shape, or pore size.106 However, there is no reported research study about in situ/operando X-ray scattering in the field of LLZO-based SSLBs. In principle, the in situ/operando X-ray scattering cell structure is similar to the in situ/operando XRD cell. Recently, the nucleation and nucleus size evolution of the Li metal at different current densities in liquid cells has been studied through synchrotron-based transmission mode grazing-incidence small-angle X-ray scattering.107 The high overpotential induced small Li particle size and dense Li dendrites, which enlarged the reaction-specific surface area and consumed more electrolytes.

3.3.2 In situ/operando X-ray imaging (XRI). Since X-rays were discovered, XRI has been extensively employed in medical studies and in recent years, this technique has been applied in battery investigations. X-Rays penetrate and irradiate the sample easily and different elements are identified based on their distinct absorption levels. As a result, images with varying contrasts are generated. Absorption contrast imaging is available to evaluate the spatial distribution of system components. However, light elements (such as Li) absorb laboratory X-rays very weakly, resulting in low image contrast, thus making it impossible to detect interior structures. Accordingly, X-rays with high flux (like SR X-rays) can make the difference for low-density material imaging.108,109 Moreover, phase contrast imaging was achieved by directly detecting the sample's phase information through the movement of interference fringes.110 Therefore, in situ XRI techniques are utilized to capture the morphological evolution of bulk materials via 2D radiography and 3D tomography.

XRI techniques mainly consist of transmission X-ray microscopy (TXM) and X-ray tomography (XRT). Fig. 14a illustrates the schematic diagram of measurements for both imaging techniques.5 XRT, also known as computed tomography (CT), adopts a monochromator to select an incident beam for the sample. A scintillator and a charge-coupled device (CCD) camera converts the transmitted photons into an optical image with micrometer spatial resolution. XRT enables visualization of the internal sample information without slicing and is a real-space full-field imaging method and obeys the optical geometry principle. TXM generally adopts hard X-rays with shorter wavelengths and couples a monochromator and a capillary condenser to establish a micro-focused beam. A Fresnel strip plate (which amplifies the imaging), phase ring and CCD camera are used to realize phase contrast by converting the transmitted photons into an optical image with nanometer spatial resolution. In comparison to other imaging techniques, TXM offers advantages of simple manipulation and usage at atmospheric pressure. From above, the resolution of XRI techniques is decided by the optics in device systems. The spatial resolution ranges from ∼15 nm to several mm.109 Furthermore, coupled tomography and far-field high-energy diffraction microscopy (FF-HEDM) have been used to evaluate the grain-level chemo-mechanics.111 As shown in Fig. 14b, compared to XRT, FF-HEDM captures diffraction images on a 2D area detector. The images are processed to provide the mass center, sizes, crystallographic orientations, and lattice strain tensor information for each grain. FF-HEDM can quantify the stress evolution at an individual grain level in response to electrochemical cycling. In addition, coupled calorimetry and operando SR X-ray fast radiography evaluate the thermal behavior of a cell.112Fig. 14c illustrates a plan of a calorimeter and an instrumented cell irradiated with X-ray beam with thermocouple recording (TCR) and pressure recording (PR). The in situ cell of this setup is a commercial 18650 type battery while the calorimeter is made of a stainless-steel tube. The closed system is thin enough to allow high-energy X-rays to be transmitted through the cell. External measurements of temperature and pressure sensors are recorded over the whole extent of the protocol. This setup successfully establishes a link between external measurements and cell internal structures. In general, in situ XRI employs Swagelok-type cells113 and Fig. 14d shows the corresponding in situ Swagelok-type cell structure. 3D images are derived by rotating the in situ cell to acquire photographic images from various angles and reconstructing them with computer software.114 More importantly, with the benefits of high flux, high brightness, high power and low incident energy of SR X-rays, XRI techniques enable in situ/operando measurements on commercially available CR2023 coin-type cells.115 As shown in Fig. 14e, high-energy X-rays are incident along the direction of the sample cross-section to directly observe the internal structure and transmission imaging is acquired. Mapping analysis of transmission images allows real-time detection of changes in the internal cell structure during the charging and discharging reaction. Therefore, in situ XRI techniques can realize 3D visualization of the changes in the local morphology, structure, and chemical composition within solid-state batteries without causing any damage.


image file: d3mh00135k-f14.tif
Fig. 14 (a) Schematic diagram for XRI. (b) End-station detectors for carrying out tomography and FF-HEDM measurements with the corresponding information expected. (c) Plan of a calorimeter and an instrumented cell irradiated with X-ray beam with TCR and PR. Reproduced from ref. 112 with permission from the American Chemical Society. (d) Schematic diagram of an in situ Swagelok-type cell structure. (e) Schematic diagram of a commercially available CR2023 coin-type cell. Reproduced from ref. 115 with permission from AIP Publishing.

Due to the low attenuation coefficient, the pore (air) and Li metal regions are transparent to X-rays. In contrast, the LLZO ceramic electrolyte region intensely attenuates the X-ray displaying opacity, thereby appearing in the image. Owing to the distinct X-ray absorption ability of pores and LLZO, the LLZO porous structure was successfully investigated using XRT technique.116–118 LLZO porous structures were initially prepared by the freeze casting method with tert-butyl alcohol (TBA) as the solvent (the mass fraction of LLZO is 20%).116 3D images captured by SRXRT reveal that the pore size is ∼50 μm and pores are distributed homogeneously in the whole structure. These pore channels have identical orientations and are even bridge-free over the length range of several hundred μm (far exceeding the LLZO thickness of ∼100 μm). In order to fit the practical manufacturing process, LLZO porous structures are prepared by the freeze tape casting method.117 The effect of various LLZO volume fractions on porous structures was investigated using XRT technique. Fig. 15a shows that unblocked, opened, and homogeneous pore channels are formed for the 7.5 and 12.5 vol% LLZO slurries, while randomly distributed macro and closed pores are formed for the 17.5 vol% sample. Therefore, the higher the LLZO volume fraction, the smaller the pore size, the lower the porosity and the greater the number of closed pores. Such phenomenon is attributed to the higher LLZO volume fraction, as interactions between particles and solvent lower the particle dispersion stability. The particles fail to be uniformly excluded during freezing procedure. Compared to the freeze casting method, pore orientation performed by the freeze tape casting method is not vertical to the surface but at an angle. The pore channels are slanted as the tape-casting process causes the temperature gradients to be not perfectly aligned from the top to bottom. The temperature gradient also leads to a structure with small pores at the bottom and large pores at the top. Subsequently, the bilayer structure was prepared with a porous structural layer and dense layer.118 LiNi0.6Co0.2Mn0.2O2 (NCM622) or other cathode slurries were infiltrated into the porous structural layer. Unidirectional and open pore channels favour cathode particle infiltration, shortening of Li+ diffusion path length, and enlargement of the contact area. The Li anode was coupled on the dense layer side and the assembled SSLB exhibited good charge/discharge capability.


image file: d3mh00135k-f15.tif
Fig. 15 (a) Reconstructed thin films and longitudinal subvolume together with the front view images extracted from the tomographic results for the 7.5%, 12.5%, and 17.5% structures, respectively. Reproduced from ref. 117 with permission from the American Chemical Society. (b) XRT reconstruction of the void phase in the interior of LLZO electrolytes sintered at 1050°, 1100°, and 1150 °C before and after failure. Reproduced from ref. 122 with permission from the American Chemical Society. (c) The 3D rendering of the pillar with internal Li protrusions and voids is viewed from three different directions. The LLZO is in transparent light yellow, Li protrusions are in opaque grey or transparent yellow, and voids inside cracks are in opaque red. Reproduced from ref. 123 with permission from Wiley. (d) X-Ray radiography during the thermal runaway of liquid electrolyte Li-ion batteries and all-solid-state batteries. The white bar gives the 2 mm scale. Reproduced from ref. 112 with permission from the American Chemical Society.

Similarly, polymers are also transparent due to non-attenuation of X-rays. Accordingly, large attenuation contrast is seen in the region between the polymer and LLZO, which allows the visualization of LLZO particles directly in CE. The factors affecting the improved CE transport properties were examined using XRT technique based on this principle.119 In theory, the contact surface area with polymer should be enlarged as LLZO nanoparticle sizes decrease when the content increases. However, LLZO particle aggregation is clearly viewed when the content increases from 5 to 50 vol%. The particle clustering significantly lowers the accessible area, resulting in increased particle aggregation size.120 The dielectric relaxation strength will decrease, indicating that the movement of amorphous polymer chains near the LLZO surface is constrained resulting in lower ionic conductivity. Consequently, ionic conductivity is highly dependent on the accessible area instead of the content of inorganic particles and an enhanced transport in CE was observed mainly due to polymer/inorganic interactions. In addition, an ordered distribution of inorganics is the ideal structure for designing CE. An asymmetric structural design was confirmed by XRT.121 An asymmetric CE with a LLZO-rich layer on the anode side and a polymer-rich layer on the cathode side was constructed via the selective adsorption of a glass fiber (GF) separator. This structure not only improves the ionic conductivity and Li+ migration ability but also addresses the SSE/electrode interface issue to a certain extent.

XRI has also been widely adopted to investigate the meso-scale transformation in LLZO-based SSLBs and the microstructural transformation in LLZO pellets. Stresses, cracks, and electrode volume changes all contribute to mechanical degradation in LLZO SSEs. Most studies have focused on characterizing the microstructural changes in LLZO pellets in response to electrochemical cycling. However, high-energy X-rays dramatically lower the attenuation contrast sensitivity of Li metals and voids. Therefore, it is hard to identify Li deposits, Li dendrites, and void regions. Fortunately, XRT can trace the evolution of LLZO phases and void regions. If isolated Li deposits or Li dendrites occur, void regions are significantly varied. Microstructural details can be analyzed by directly comparing XRT images of pristine LLZO pellets and failed samples. As shown in Fig. 15b, LLZO pellets sintered at 1050°, 1100°, and 1150 °C exhibit distinct void distributions before and after failure.122 In comparison to pristine samples, the increase in the number of voids in the failed LLZO demonstrate Li deposition. It suggests that Li metal is deposited in an isolated form or as an agglomerate and expanded in the void regions. Furthermore, samples sintered at higher temperatures (1150 °C) displayed more connected and tortuous void networks. Samples with higher connectivity exhibited more rapid Li dendrite formation and propagation and hence, high-density LLZO pellets are essential. In order to distinguish voids and Li metal, XRT with a high-resolution imaging mode is employed.123 Such an imaging device utilizes a comparatively low X-ray energy of 5.4 keV to enhance Li and X-ray interactions. Inside the cracks in the XRT image, the lighter part corresponds to Li and the darker part corresponded to voids. According to greyscale discrepancies, Li, voids and LLZO can be clearly segmented and rendered as yellow, red and transparent light yellow, respectively (Fig. 15c). On one hand, Li protrusions mainly propagate intergranularly through LLZO, forming a wavy plane whose geometry is templated by grain boundaries. On the other hand, Li protrusions also form flat branches in a transgranular mode, located at complex geometrical features such as triple-junction grain boundaries. This indicates that Li deposition and Li dendrite formation favorably occurred at cracks, voids, and grain boundaries.124 Importantly, coupled tomography and FF-HEDM can evaluate the grain-level chemo-mechanics in LLZO pellets.125 High location overlaps between stress hotspots (grains with maximum stress values) and cold spots (grains with minimum stress values) and second phases were revealed by coupled FF-HEDM and tomography data; Li metal was preferentially deposited near such locations. Results indicate that failure in LLZO pellets was initiated locally since trace secondary phases cause mechanical stresses and result in ion transportation response inhomogeneities. In addition, a stable anode morphology is critical to achieve high electrochemical performance in LLZO-based SSLBs. Nevertheless, pore formation and local hotspot generation in Li metals are visualized by XRT with a high-resolution imaging mode, confirming the non-uniform interface dynamics.126 Advanced image processing techniques revealed that Li metal hotspots correlate to anisotropic microstructures in LLZO. In particular, the failure cores are the local regions with suboptimal effective properties (such as ion transportation and mechanical stresses). Li dendrites that grew parallel to the LLZO/anode interface were detected via CT imaging.127 The morphology of Li dendrites was “bowl-shaped” as observed by in situ SEM and “spalling” observed by in situ OM, indicating that internal stresses induced by Li deposition as the major driving force for propagation. Overall, LLZO shows severe chemo-mechanical degradation during electrochemical cycling. Cracks and secondary phases at the interface resulted in the emergence of local high-stress regions. Li was deposited in such regions preferentially, which led to LLZO failure. Application of stack pressure (1 to 20 MPa) or enhancement of temperature (25° to 100 °C) enabled homogeneous Li+ deposition at the interface.128 Assessing the chemo-mechanical response of LLZO under relevant operating conditions is thus essential. Nonetheless, designing experimental setups that are compatible with X-ray end-stations with desired pressures and temperatures in imaging systems is challenging. It is expected that future work on setup design will enable controlled in situ and operando experiments.

In practical work on LLZO-based SSLBs, battery safety has received scarce attention and has not been assessed significantly until now. Fortunately, the thermal runaway of liquid electrolyte LIBs and SSLBs was visualized by coupled calorimetry and operando SR X-ray fast radiography (Fig. 15d).112 Compared to LIBs, heat, and gas release in SSLBs during thermal runaway decreased by 11 ± 10% and 40 ± 20%, respectively, while reaction kinetics increased by 42 ± 20%. The decreased heat and gas release in SSLBs was attributed to fewer combustibles, such as liquid electrolytes and separators in the system. The increased reaction kinetics were attributed to improved thermal conductivity of LLZO ceramic particles and evaporation of a part of the liquid electrolyte.

In summary, in situ X-ray techniques offer straightforward visualization of microstructural evolution in LLZO-based SSLBs, including electrolytes, electrodes, and electrolyte/electrode interfaces. It is noted that the application of XRI in SSLBs is still at an early stage. XRI with laboratory X-ray sources is low-cost and operator-friendly, while low-intensity restricts resolution and makes it difficult to separate lighter elements. Although SR X-ray sources with high energy and precision capabilities enhance resolution, exorbitant experimental costs limit its use in normal laboratories. Hence, further practical design work will contribute to a comprehensive understanding of both microscopic and chemical structural evolution in SSLBs through XRI techniques.

3.3.3 In situ/operando X-ray photoelectron spectroscopy (XPS). XPS, also known as electron spectroscopy for chemical analysis, is a powerful technique for analyzing solid surface science (Fig. 16a). Utilizing photons in the X-ray range, which cause emission of core electrons in the sample, XPS enables qualitative examination of surface elements by measuring electron binding energies.129 Accordingly, it can detect all elements except H and He and provide the chemical bond level information. Notably, laboratory-scale surface XPS generally only analyzes sample surface chemistry with thicknesses of ∼10 nm. In contrast, XPS with a synchrotron source allows depth sensitivity at 2–50 nm via tuning photon energy while simultaneously enabling faster measurement speed and higher resolution.93 Considering the demand for testing to greater depth (>50 nm), ion sputtering techniques are available to etch sample surface layers and subsequently expose the sub-surface layer. When ion sputtering is coupled with XPS technique for in situ sample characterization, elemental information at various depths can be obtained. Crucially, an ultra-high vacuum (10−8 Pa) is essential for XPS operation. An advantage of LLZO-based SSEs over liquid electrolytes is the ability to load into vacuum chambers directly without evaporation. Therefore, in situ XPS enables extraction of useful information about surface/interface structures and chemical composition variations that are not easily accessible by other techniques (such as in situ OM, SEM, and XRI).
image file: d3mh00135k-f16.tif
Fig. 16 (a) Schematic diagram for XPS measurements. (b) Schematic diagram of an in situ XPS cell and the configuration of in situ electrochemical XPS measurements. Photographs of (c) the sample holder and (d and e) the vacuum suitcase for transferring samples without air exposure. (f) Laboratory-type XPS apparatus. (g) The simplified schematics of the analysis chamber, electrical connections, and the stage with the sample holder. The terminals A and B are connected to a potentiostat via vacuum feed through external coaxial cables. Reproduced from ref. 130 with permission from the American Chemical Society.

Fig. 16b illustrates an in situ XPS cell schematic diagram and the configuration of in situ electrochemical XPS measurements.130 An active electrode material layer was deposited on LLZO, such as an amorphous Si layer, by using Ar radio frequency magnetron sputtering. The sputter-deposited electrode active material layer was coated with a Cu collector by direct current (DC) sputtering. Importantly, during Cu plating, the central part of the electrode active material layer was masked with a stainless-steel stencil mask. The obtained uncoated Cu part of the active electrode material served as an analysis area for XPS measurements. The Li metal layer was thermally deposited on the other side of the LLZO. The Li metal layer of the in situ XPS cell was bonded to the Cu foil loaded on a sample holder (Fig. 16c). The bottom side was electrically isolated from the sample holder via a polyimide film and connected to terminal A. The top side was connected to the sample holder and terminal B. The cell-loaded sample holder was sealed in a vacuum suitcase and attached to the load lock (Fig. 16d, e). This was then transferred into the XPS apparatus without exposure to atmosphere (Fig. 16f). The sample holder was mounted on a stage in the analysis chamber for in situ electrochemical measurements (Fig. 16g). Terminal B and the sample holder were electrically grounded to the hemispherical analyzer of the XPS apparatus and then bias was applied to terminals A and B by using an electrochemical workstation.

LLZO and the doped (Al, Ta, Nb, and Ga) versions typically exhibit a wide electrochemical stability window and superior stability during charging and discharging processes. However, in situ and operando XPS revealed that an oxygen-deficient interfacial layer (ODI) was developed at the Al-LLZO/Li interface.131 As shown in Fig. 17a, ODI formation was attributed to the reduction of Zr4+ to Zr2+ and Zr0 by the introduction of higher energy when Li came into contact on the LLZO surface. Operando XPS measurements reveal that ODI formation is self-limiting, with very little change in the degree of Zr4+ reduction with increased Li deposition time after ∼1.5 min. Similarly, depth profile XPS also revealed the reduction of Zr4+ and generation of Li-O species at the Ta-LLZO/Li interface.132 The formation rate of ODI is highly affected by current density at the interface. At low current densities, ODI has a passivation effect to protect LLZO from sustained reduction and constrain interfacial resistance. In contrast, ODI formation was probably obstructed, and Li dendrite formation at interface was accelerated at high current densities. Interestingly, electrochemical interface response was not changed by the existence of ODI; in other words, interfacial charge transfer was not impeded. The lithiation/delithiation of amorphous Si electrodes on Ta-LLZO in thin film cell structures was investigated by in situ electrochemical XPS.130 As shown in Fig. 17b, during initial lithiation/delithiation, the formation of irreversible species such as Li-silicates, Li2O and Li2CO3 at the interface was observed. Such species contribute to irreversible capacity loss while interface reactions were also found at the cathode/LLZO interface. The first example is the LCO cathode where depth profile XPS revealed that deleterious Li2CO3, LZO and LaCoO3 species were formed at the interface due to Co and La diffusion during annealing (Fig. 17c).133 Owing to the formation of interfacial phases, Li migration ability was decreased, and interface impedance was increased. Results suggest that lower temperatures, shorter time and CO2-free environments are processing conditions to achieve an ideal cathode/LLZO interface. Second, Mn in LiMn2O4 (LMO) also diffuses into LLZO during higher temperature cycling, which causes significant deterioration of cell properties.134 Mn regions at the interface were detected by depth profile XPS and they were found to become wider with increasing cycles. Therefore, understanding the dynamic transformation of the interface and its role in electrochemical degradation is beneficial in stabilizing such interfaces.


image file: d3mh00135k-f17.tif
Fig. 17 (a) Operando XPS measurements of individual core-level spectra during electrochemical deposition of Li showing their evolution with increasing Li deposition time. Zr 3d core-level spectra compare the extent of Zr4+ reduction as a function of the deposition technique. Reproduced from ref. 131 with permission from the American Chemical Society. (b) Reproduced from ref. 130 with permission from the American Chemical Society. (c) Evolution of surface oxide, carbonate and hydroxide species with heat treatment temperature for LLZO solid electrolyte. Reproduced from ref. 133 with permission from the American Chemical Society. (d) XPS depth profile of Li3N- or LiF-rich layers at LLZO/Li anode interfaces. Reproduced from ref. 139 and 140 with permission from the American Chemical Society and Elsevier. (e) XPS depth profile of a freshly dissociated sample after 24 h air exposure: O 1s and Li 1s core spectra for various sputtering time periods from 1 to 15 min. Reproduced from ref. 142 with permission from the American Chemical Society.

Until now, many interface modification approaches have been successfully proved via depth profile XPS. First, introducing minor liquid electrolytes at cathode/LLZO and LLZO/Li anode interfaces can alleviate but not completely solve the issue cell electrochemical performance deterioration. XPS depth profile experiments show a spontaneous chemical reaction between LLZO and liquid electsrolyte.135 The reaction layer formed on the LLZO surface is composed of a Li-rich phase near the surface and Li–La–Zr oxide on the sub-surface. A continuous increase in reaction layer thickness led to lower cell capacity efficiency and retention. For addressing chemical reactions induced by minor liquid electrolytes, nanometer-scale self-assembled monolayers (SAMs) with acidic anchoring groups were adopted.136In situ Li+/H+ exchange or intense interactions between LLZO and liquid electrolyte were weakened by the compact SAM passivated layer, which achieved an interface resistance close to 0. Second, introducing ionic liquid-based gel interlayers on anode and cathode sides dramatically lowered the interface resistance and stabilized the interface.137 The formation of SEI films in the interface region similar to those in conventional liquid cells was revealed by depth profile XPS. The amorphous nano-interlayer was characterized by an outer layer which was rich in organic species while the inner layer (near the LLZO surface) was dominated by inorganic species. Such structural character enables good battery performance. Third, the introduction of alloy layers at the LLZO/Li anode interface is also a promising approach. Depth profile XPS indicated that a ∼10 nm Sb metal layer was transformed to a Li–Sb alloy layer during lithiation/delithiation.138 The alloy layer permitted efficient Li+ and e percolation and mitigated cavity and Li dendrite formation. The batteries assembled with intercalation-type V2O5 cathodes delivered favorable areal capacity and peak current density. Lastly, molten state composite anodes were constructed to achieve close contact and a stable LLZO/Li anode interface. Composite anodes were prepared by incorporating nitrides such as 2D boron nitride nanosheets139 or graphitic carbon nitride (g-C3N4)70 into molten Li resulting in high viscosity and low surface tension. The LLZO/Li anode interface allowed close integration due to the transformation from point contact to surface contact. Simultaneously, a Li3N-rich layer was formed at the interface which featured high ionic conductivity and low electronic conductivity (Fig. 17d, left). Analogously, the incorporation of fluorides such as MXene with surface terminal functional group of –F140 or AlF3141 into molten Li also achieved close contact and a LiF-rich layer at LLZO/Li anode interface (Fig. 17d, right). Such approaches reduced interface resistance and improved the critical current density value, suppressing Li dendrite formation and thus achieving stabilization of electrochemical cycling.

As previously mentioned, a thin and insulating Li2CO3 layer was formed on LLZO pellets upon air exposure resulting in high interface resistance. The formation mechanism of the Li2CO3 layer was investigated by depth profile XPS. Initially, it was suggested that the air-contaminated LLZO structure was composed of three layers, and these were surface, sub-surface and inner layers which were the Li2CO3 layer (from CO2 and LiOH reaction), intermediate oxide or hydroxide layer, and bulk LLZO layer, respectively.15 Later research indicated that the sub-surface layer was essentially comprised of protonated Li7−xHxLa3Zr2O12 and LiOH due to Li+/H+ exchange (Fig. 17e).142 The Li2CO3 layer significantly obstructs Li+ migration at the interface which facilitates massive void generation and ultimately causes cell failure. Heat treatment is a simple approach to eliminate contaminated layers completely. The temperature range for decomposition of Li2CO3 to CO2 is 620°–1000 °C. In contrast, in situ XPS showed that the decomposition of carbonate species on the LLZO surface initiates at ∼150 °C.143 Notably, it is essential to prevent re-exposure of the clean surface to air after heat treatment. Otherwise, it will result in the formation of contaminated layers. Based on the knowledge of contaminated layer composition, this can be converted into an advantageous interlayer through different electrochemical reactions. For instance, the contaminant layer was converted to an aromatic layer via a simple azo reaction at 60 °C.144 XPS depth profile characterization indicated that the aromatic layer was composed of a C–F-rich surface layer and a LiF-rich bottom layer. The aromatic layer not only offers favorable lithophilicity, but also facilitates Li+ migration and homogeneous Li+ deposition. Alternatively, acid treatment was adopted to eliminate Li2CO3 from the surface and grain boundaries while introducing protons.145 Proton doping suppresses formation of conductive by-products and releases residual stress for retaining the intact interface contact, which remarkably enhances stability and energy density of LLZO-based SSLBs. Moreover, the laboratory-used Li foil surface was found to contain impurities such as Li2O, Li2CO3, LiOH or LiF as seen from the XPS depth profile.146 It is critical to provide a clean Li surface for achieving close contact at the LLZO/Li anode interface. The clean Li surface can be realized by heating Li foil to a molten state and then scraping off the contaminated layer which floats on the surface.

LLZO-based SSEs prepared by various synthesis methods generally exhibit distinct structures, properties, and applications. When LLZO particles were ball-milled together with LiBH4, a unique amorphous dual coating was formed on the surface, which consisted of a LiBO2 inner layer and a LiBH4 outer layer and this was revealed by XPS depth profiles.147 The amorphous coating acts as a filler, binder, and bridge, effectively enhancing contact among particles. The cold-press molded, and un-sintered pellets exhibited high densities, high ionic conductivity, high Li+ transference numbers, and low electronic conductivity. The mechanism of phase conversion in LLZO nanofibers prepared by electro-spinning was investigated using XPS depth profiles.148 The structure and chemical composition of LLZO nanofibers sintered at various temperature stages were analyzed in detail. The diameter of nanofibers was less than 500 nm, which is adequate for complete penetration in XPS depth profile experiments. The phase conversion of nanofibers was identical to that of particles, indicating that various LLZO morphologies cannot affect the phase conversion mechanism.

LLZO thin films are crucial for flexible electronic device applications. High ionic-conductivity, high-density, homogeneous, and crack-free tetragonal LLZO thin films were grown on Pt via a novel CO2 laser-assisted chemical vapor deposition (CVD) technique.149 The presence of a Li2PtO3 layer between LLZO and Pt was proven by XPS depth profile analysis. The presence of such a layer affects total impedance but not the ionic conductivity of the LLZO thin film system. Therefore, the thin film grown via CVD is an ideal candidate for solid-state thin-film lithium batteries. Si nanoparticles filled in polymers help absorb and anchor Li dendrites.150 Li–Si–O and LixSi species formed by reactions between Si nanoparticles and Li dendrites were identified by depth profile XPS. Such reactions suppress Li dendrite formation at the interface and swallow tips of Li inside SSEs to avoid longitudinal growth. Under the condition that the fundamental properties of LLZO-based SSEs are not impaired, adding other nanomaterials that spontaneously react with Li should also produce similar effects. Lastly, the depth profile XPS revealed that LLZO, polyvinylidene fluoride-hexafluoropropylene copolymer (PVDF-HFP), PEO, and SN synthesized CE formed hierarchical SEI on the Li anode side, consisting of a LiF-rich outer layer and a Li3N-rich inner layer.151 The assembled cell offered good electrochemical properties, flexibility, and interface adhesion, displaying application prospects in the field of flexible electronics.

Overall, in situ XPS qualitatively examines surface/interface elements via measurement of electron binding energy, which characterizes phase and elemental composition. Compared to in situ XRD measurement, where phase content is only reflected in certain amounts, in situ XPS offers a wider measurement range and higher sensitivity. However, it fails to visualize interface evolution and Li dendrite morphology as an imaging technique in real time. Furthermore, individual detection depth is less than 50 nm and requires integration with ion sputtering techniques to detect the whole composition. Therefore, the development of innovative XPS technology and compatible in situ cells is still a major challenge.

3.3.4 In situ/operando X-ray absorption spectroscopy (XAS). XAS provides local information for selected elements in materials by accurately measuring X-ray absorption coefficients as a function of incident X-ray energy (Fig. 18a). X-Ray energy ranges vary above and below the absorption range of selected elements. Consequently, XAS, an element-specific local structure probe, is extensively utilized in local geometry and electronic structure research.152 X-rays absorbed by a sample excite core electrons to jump in vacant orbitals or continuum states, generating various energies. Based on the relative energy range of absorption edges, XAS has been divided into X-ray absorption near-edge structure (XANES) spectroscopy and extended X-ray absorption fine-structure (EXAFS) spectroscopy. XANES and EXAFS spectroscopy cover the region within 50 eV and approximately 50–1000 eV above the absorption edge of an element, respectively. XANES spectroscopy offers information on chemical valence, electronic structure, site symmetry, and covalent bond strength of a specific element, while EXAFS spectroscopy provides information on the neighboring atomic structure of a measured element including interatomic distance and coordination number.153 In addition, XAS is available in both surface-sensitive total electron yield (TEY) and bulk-sensitive total fluorescence yield (TFY) modes. Detection depths of TEY and TFY modes are ∼10 nm and ∼100 nm, respectively. Incorporation of such two modes enables access to depth-dependent chemical information. Notably, since the in situ/operando XPS cell structure is similar to the in situ/operando XRD cell structure, no further discussion is provided regarding the setup.
image file: d3mh00135k-f18.tif
Fig. 18 (a) Schematic diagram for XAS studies. (b) Ni L-edge and O K-edge XAS spectra of hybrid LLZO and NCM111 samples detected in the TEY mode and TFY mode at different stages. The inset shows the comparison of the Ni L3 edge. (c) Brief schematic illustration for the alteration of surface topography during the ball-milling and cosintering process. Reproduced from ref. 154 with permission from the American Chemical Society. (d) Co K-edge, Co L3-edge, and O K-edge X-ray absorption spectra of hybrid LCO and LLZO samples at different stages. Reproduced from ref. 133 with permission from the American Chemical Society. (e) Operando S K-edge spectra with first derivative mapping, Ni K-edge spectra, and charge/discharge profiles of bare NMC811-LGPS. Reproduced from ref. 155 with permission from the American Chemical Society.

Fig. 18b illustrates Ni L-edge and O K-edge XAS spectra of hybrid LLZO and LiNi1/3Co1/3Mn1/3O2 (NCM111) samples in TEY and TFY modes.154 Results indicated that LLZO spontaneously coats NCM111 particles with a thickness of ∼100 nm after ball milling. The thickness of the LLZO layer was reduced to ∼10 nm and an interlayer of ∼3 nm was seen after cosintering at 600 °C (Fig. 18c). The interlayer was attributed to Ni/La and Ni/Li exchange resulting in the formation of LZO and LaNiO3 species. The structure, chemical information and charge transfer resistance of the LCO/LLZO interface after cosintering at 500 °C were also investigated by in situ XAS.133 As shown in Fig. 18d, Co K-edge, Co L-edge, and O K-edge XAS spectra revealed Co elemental oxidation and LLZO decomposition, causing the formation of La-Co-O species with a thickness of 20 nm at the interface. Hence, temperature, time, and atmosphere control in cosintering are vital to realize a superior cathode/LLZO interface. Besides, operando XANES spectroscopy was conducted to investigate the interface behavior between the NCM811 cathode and Li10GeP2S12 (LGPS) electrolyte.155 S K-edge and Ni K-edge spectra shown in Fig. 18e suggested that NCM811 and LGPS are unstable during the electrochemical process. Partial LGPS is decomposed to a Li2S contaminant phase. Micro-cracks and phase evolution from a layered to rock-salt structure occur on the NCM811 surface. Operando XANES spectroscopy further proved that the LiNbOx coating layer protected the integrity of the NMC811 cathode and stabilized the NCM811/LGPS interface. Therefore, in situ/operando XAS exhibits considerable potential for probing reactive intermediate states, interface products and reaction mechanisms. Notably, the in situ/operando cell setup poses several challenges during experimental manipulation and requires further development and optimization.

3.4 In situ/operando scanning probe based tools

3.4.1 In situ/operando atomic force microscopy (AFM). AFM, a typical force-based technique, is a new scanning probe microscopy (SPM) technique developed on the basis of scanning tunneling microscopy (STM) (Fig. 19a).156 AFM has been extensively employed in surface/interface analysis research due to the advantages of high resolution at an atomic level (1–20 nm in plane and <0.025 nm in height), low intrusion, multi-modal measurements, and application flexibility. As an AFM probe scans the sample surface, a nano-scale cantilever probe is subjected to bending deformation due to the interaction with the sample. Then, 3D morphological imaging of the sample surface is realized by detecting and converting bending deformation signals.157 AFM typically contains three imaging modes: an intermittent contact mode, a contact mode, and a non-contact mode.158 First, in the intermittent contact mode, a cantilever probe tip and sample are in intermittent contact and oscillate close to resonant frequency. When the tip is close to the sample surface, the amplitude of the tip reduces owing to gravitational forces between them. 3D images of sample surface morphology are derived by recording variations of amplitude. Second, the sample surface is scanned in a contact mode in a constant height or force mode. 3D images of sample surface morphology are obtained via recording variations in the applied voltage which is needed to maintain height or force. Finally, in a non-contact mode, long-range forces such as van der Waals or electrostatic forces between the tip and sample are utilized to maintain no contact. Sample surface morphology information was accessed by detecting variations of force.
image file: d3mh00135k-f19.tif
Fig. 19 (a) Schematic diagram showing AFM microscopy technique. (b) Cross-sectional schematic diagram of an EC-AFM cell. Reproduced from ref. 160 with permission from Elsevier. (c) A digital image and schematic diagram of an in situ EC-AFM cell used for cross-sectional measurement. Reproduced from ref. 40 with permission from Elsevier. (d) Schematic illustration of ionic and electronic co-migration or electronic migration imaging devices based on conductive AFM. Reproduced from ref. 161 with permission from Elsevier.

In contrast to OM or EM, AFM provides surface information (including morphology, mechanical properties such as stress or Young's modulus and electrical properties such as electrical conductivity or diffusion resistance) of numerous materials (e.g., conductors, semiconductors, and insulators) in diverse environments (e.g., ambient air, liquids, or high temperatures).159 Electrochemical AFM (EC-AFM), frequently used in the field of battery research, is mainly employed in the contact mode. Incorporating the benefits of high-resolution surface/interface analysis and electrochemical reactors, EC-AFM provides the evolutionary principles and electro-mechanical properties of surfaces/interfaces in real space during the charging and discharging process.

A fundamental architecture of a three-electrode electrochemical AFM liquid cell for surface/interface probing is shown in Fig. 19b.160 High-resolution images are obtained from variations of phase difference between the driving signal and cantilever response. In LLZO-based SSLBs, given the solid–solid interface of buried electrodes/LLZO-based SSEs and high-activity Li metal, it is critical to refine the fundamental architecture and condition of liquid cells for friendly and non-destructive in situ AFM measurements. Based on this, the design of AFM electrochemical in situ LLZO-based SSLBs should be as follows: (1) cell assembly should be placed in an argon-filled glovebox since components are sensitive to air; (2) there must be excellent electrical connections among the electrode, the current collector and the external circuit; (3) distance between the counter electrode and working electrode inside the cell should be minimized to shorten Li+ migration distance, that is, LLZO-based SSEs should be minimized; (4) avoidance of internal short-circuits during cell assembly; (5) cell assembly materials and cantilever probes should be electrochemically inert to electrodes and LLZO-based SSEs; and (6) experiments should be conducted under the protection of an inert gas such as argon.

Fig. 19c illustrates a digital image and schematic diagram of an in situ EC-AFM cell used for cross-sectional measurement.40 The EC-AFM apparatus is placed in an Ar glovebox and the in situ cell is scanned using an insulated silicon AFM tip. The Li anode, electrolyte, and cathode are sliced with a scalpel to obtain a high-quality cross-section. The layered films are compressed with polyamides and stainless-steel plates. The in situ electrochemical cell was assembled on the corresponding spectroscopy and imaging system coupled with a potentiostat to investigate real-time interface reaction processes. Fig. 19d illustrates the schematic illustration of ionic and electronic co-migration or electronic migration imaging devices based on conductive AFM (c-AFM).161 The Young's modulus, adhesion and current can be simultaneously obtained. The half-cell consists of LLZO-PEO electrolyte, a Li metal anode and a conductive AFM probe. The sample table is a current collector to drive Li+ migration with a certain voltage. When a positive voltage is applied to the sample table, Li+ from the Li metal anode passes through electrolyte and transforms into Li atoms by reduction of the AFM probe after reaching the sample surface. Since the probe range is significantly small, reduced Li+ is removed from the surface by the probe to prevent bulk deposition. Meanwhile, e migrate in the opposite direction from the probe to the current collector. As a result, current collected by c-AFM is the sum of electronic and ionic currents under a positive voltage. When a negative voltage is applied to the sample table, Li+ does not migrate from the probe to the Li metal anode through electrolyte due to Li-depleted electrolyte. The current collected by c-AFM is the only electronic current. Although in situ AFM techniques have been used in a limited capacity in LLZO-based SSLBs so far, several investigations have been conducted to assess electrochemical–mechanical transformations in the electrode, electrolyte, or electrode/electrolyte interface during electrochemical reactions.

In the field of Li-based battery research, in situ AFM was originally employed in 1996 to investigate the exfoliation/deposition behavior of various types of Li salts (including LiClO4, LiPF6, and LiAsF6) in propylene carbonate (PC) solvent on the Cu current collector surface.162 In an electrolyte with LiAsF6, Li+ deposition on the Cu current collector surface is seen to form a smooth layer. In the field of LLZO-based SSLBs, the potential differences between NCM811 particles were measured using AFM on the cathode side.163 The average potential of NCM811-Li3PO4-LLZO composite cathode particles (−144.3 mV) was lower than that of NCM811-LLZO (−42.1 mV). The results indicate that Li3PO4 interlayer modification weakens the space charge layer. As a result, charge transfer between LLZO and cathode active material is boosted. On the Li anode side, both growth mechanism and mechanical properties of Li whiskers under mechanical constraints were investigated by in situ AFM coupled with environmental TEM (AFM-ETEM).164 The Li2CO3 layer on the Li metal surface was crucial to stabilize Li whiskers and avoid electron damage during measurement. With an increasing applied voltage, the growth process of Li whiskers consists of three stages: (1) A spherical Li dendrite is generated and the size is increased. (2) Li whiskers develop axially due to accumulation of inner stresses. (3) Li whiskers collapse suddenly by the release of axial stress. Moreover, the growth of individual Li whiskers generates a stress of up to 130 MPa and yield strength as high as 244 MPa at room temperature. By utilizing mechanical characterization of AFM to monitor the stress during Li dendrite growth in real time, it can provide a novel idea for suppression of Li dendrite growth. Afterwards, morphological evolution of Li deposition in Au, Ag, and Si interlayers during cycling was evaluated by AFM.35 As shown in Fig. 20a, Li deposition in Ag interlayers exhibited a smaller and more homogenous Li nucleation distribution and lower roughness compared to that in Au and Si interlayers. The cycle efficiency with the Ag interlayer system was also higher than that with the Au and Si interlayer system. Notably, an Ag interlayer was deformed during Li deposition/exfoliation due to the formation of Li–Ag alloy, which resulted in lower Coulomb efficiency. Overall, the variations in cell performance are determined from varying electrodeposition and stripping kinetics of interlayers. Furthermore, the flexible and electronically-insulated Li-inserted polyacrylic acid (LiPAA) layer automatically adjusted the stress intensity according to strain evolution, thereby suppressing dendrite growth during the Li plating/stripping process.165 These artificial interlayers not only resolve the issue of dendrite growth, which can cause cell failure during cycling, but also effectively mitigate the side reactions between Li metal and air, thereby significantly enhancing safety.


image file: d3mh00135k-f20.tif
Fig. 20 (a) AFM images of the surface of silver and gold interlayers at various charge and discharge states (Rrms = root mean square roughness, and the scale bar in each image = 10 μm). Reproduced from ref. 35 with permission from Wiley. (b) In situ AFM images of CE with LiTFSI, LiFSI and LiTFSI-LiFSI bi-Li salts during initial charging and discharging. The scale bars are 1 μm. Reproduced from ref. 40 with permission from Elsevier. (c) In situ conductive AFM characterization of Li+ migration in CE with 50 and 75 wt% LLZO at 55 °C. Reproduced from ref. 161 with permission from Elsevier.

For CE, the effect of various PEO molecular weights on electrochemical and interface mechanical properties was investigated using AFM.166 CE with 300 K exhibited high adhesion resulting in higher capacity and retention. The joint variation of adhesion and Young's modulus indicated that surface mechanical properties were decided by the sub-surface distribution of ceramic particles in CE. The morphological evolution of CE (PEO and Ta-LLZO) during cycling in Li–S cells was real-time monitored on a nano-scale via in situ EC-AFM.39 With dissolution of PSs into CE, PEO is progressively transformed into an amorphous state. Interactions between the PEO matrix and Ta-LLZO particles were continuously weakened, which caused Ta-LLZO to become unstable and movable. Eventually, the volume of CE was increased as Ta-LLZO particles separated from the PEO matrix. In situ OM and in situ Raman results revealed that PS shuttling had a critical influence on the mechanical stability of CE and cycling stability of cells. In situ EC-AFM further explored the modulating effects of different Li salt anions in CE on PS shuttling and interface stability.40 As shown in Fig. 20b, CE with LiTFSI, LiFSI, and LiTFSI-LiFSI bi-Li salts exhibited differences in structural evolution and interfacial behaviors during charging and discharging. In the LiTFSI system, the reaction kinetics were accelerated by severe dissolution of PSs, resulting in structure collapse within the interior of CE and weakening of interfacial contact. In contrast, the LiFSI system delayed the dissolution of PSs, which prevented structure deformation and generated a LiF protective layer for enhancing strain tolerance. However, FSI reacted with PSs to form Li-sulfite salts, resulting in constant and irreversible consumption of active S. The LiTFSI-LiFSI bi-Li salt system provides high ionic conductivity and wettability through LiTFSI and a LiF protection layer through LiFSI, achieving an optimal balance between interface compatibility and PS solubility. The bi-Li salt system is beneficial for realizing a superior interface and electrochemical properties. Therefore, real-time observation of in situ EC-AFM offers straightforward insights into interface morphology evolution mechanisms and reaction kinetics, guiding the direction for material design and property enhancement in LLZO-based SSLBs. In situ c-AFM was also performed to reveal the effect of different LLZO contents and temperatures on Li+ and e migration in CE.161 The images revealed that at a low temperature of 30 °C, Li+ only migrates along amorphous PEO regardless of the LLZO content. At a high temperature of 55 °C, Li+ migrates mainly along amorphous PEO in CE with 50 wt% LLZO, while Li+ can migrate along LLZO particles in CE with 75 wt% LLZO (Fig. 20c). Results suggest that Li+ migration pathways in CE are jointly affected by LLZO content and temperature. Crucially, in situ c-AFM showed that the introduction of LLZO particles enhanced CE electronic insulation. Hence, identification of ion migration pathways in CE via morphology evolution paves the way for rational design in future LLZO-based SSLBs.

Overall, in situ AFM provides better resolution and a vacuum-free working environment compared to in situ OM or in situ SEM. More importantly, AFM can examine insulative materials and mechanical properties, which are available to instruct native interface enhancement and artificial interlayer design. Although in situ AFM enables visualization of Li dendrite growth and surface/interface structure evolution, it suffers from a slower imaging rate. At high current density, the temporal resolution is inferior, resulting in failure to capture the whole evolution process and omitting vital procedures. Consequently, the improvement in the temporal resolution of in situ AFM is an issue to be addressed.

3.4.2 In situ/operando scanning electrochemical microscopy (SECM). SECM, a scanning probe technique based on electrochemical reactions for measuring local electrochemical activity of a sample in solution, has been utilized to probe SEI film surface characteristics in liquid cells. SECM adopts an ultra-micro-electrode probe to scan and measure Faraday current of the redox medium which interacts with the sample, acquiring electrochemical information from selected regions.167 The tip current increases or decreases based on back response offering positive or negative feedback, respectively. The measured sample can be a conductor, semiconductor or insulator and the measured interface can be a liquid–liquid interface, solid–liquid interface, or gas–liquid interface. The current distribution evolution of the local SEI in liquid cells was investigated by SECM.168 The electron transfer capacity of Li protrusion regions was observed to be stronger than that of the flat surface, and thus the secondary current distribution on the irregular surface of the complex 3D Li metal anode was inhomogeneous. However, only SECM technique was available to research systems due to the constraints of the operating principle. In LLZO-based SSLBs, SECM technique was applied to acquire electrochemical information between a liquid electrolyte interlayer (or a novel organic solution interlayer) and LLZO during electrochemical reactions. Besides, several other scanning probe techniques such as Kelvin probe force microscopy (KPFM) and laser scanning confocal microscopy (LSCM) can also be employed to investigate LLZO-based SSLBs in future.

3.5 In situ/operando magnetism based techniques

According to the specific spin angular momentum and magnetic momentum, molecules, atoms, and ions with unpaired electrons in species exhibit differing behaviours in a magnetic field. Substances containing 7Li show paramagnetism since the magnetic moments of the original disordered orientation undergo directional changes.169 Therefore, magnetism-based techniques reflect an irreplaceable position in the LLZO-based SSLB research field.
3.5.1 In situ/operando nuclear magnetic resonance (NMR). The researched non-zero magnetic moment nucleus resonates to absorb radio frequency radiation at a certain frequency as magnetic field is applied to the system and this is the underlying principle of NMR technology. It is a non-destructive technique and is extensively employed in biology, medicine, materials science, environmental science, chemistry, and other fields. Since electrons surround the nucleus, NMR enables investigation of local magnetic field changes around the nucleus, which allows qualitative analysis of the electronic structure in nearby chemical environments.170 In recent years, various NMR approaches have been developed, including solid state NMR (SSNMR), magic angle spinning (MAS) NMR, pulsed field gradient (PFG) NMR, NMR relaxation time measurement, two-dimensional exchange spectroscopy (2D EXSY) NMR, and magnetic resonance imaging (MRI) techniques.170,171

In the solid-state battery field, SSNMR is a critical approach to characterize ionic dynamics and local structures at the atomic level via detection of certain nuclei such as 7Li or 6Li. MAS NMR is used for a precise local structure and site occupancy of SSE or information about interface species. PFG NMR and NMR relaxation time measurements have been used for ionic movements on various length scales or time scales. PFG NMR is available for a time scale of 10−2 to 10−1 s, corresponding to a length scale of ∼1 μm. In contrast, NMR relaxation time measurements, including spin–lattice relaxation (T1), spin–spin relaxation (T2), and spin–lattice relaxation in a rotating frame (T1ρ), are available on a time scale of 10−10 to 10−3 s.172 Specifically, the time scales of T1, T2 and T are approximately 10−10–10−7, 10−6–10−4, and 10−4–10−3 s, respectively. Ionic dynamics in materials are qualitatively analyzed since the line width of NMR signals is inversely related to T2.173 2D EXSY NMR is employed to explore the transport pathway and transport rate between two different chemical sites (either in a single material such as SSEs or in two materials such as SSEs and cathodes), i.e., ionic transport pathways both in SSEs and at interfaces can be investigated.174 MRI, a non-invasive imaging technique, is utilized to visualize interface regions to provide information on the structure, morphology, and elemental distribution. In situ NMR technique is generally adopted for real-time monitoring of the evolution of structural and compositional information about the electrode, electrolyte, and electrode/electrolyte interface in lithium batteries. Li exhibits diamagnetism either in the electrolyte or interface film while the Li metal anode shows paramagnetism, resulting in distinction in identification due to difference in resonance frequencies. In addition, radio frequency signals from Li metal anodes are constrained at a surface depth of ∼12 μm (called the skin effect), while characteristic size of Li deposits is smaller than such depths.175 The radio frequency signal can completely penetrate Li deposit microstructures; therefore, in situ NMR technique can qualitatively analyze the evolution of Li microstructures at the interface.

Design of in situ cells in NMR experiments is considerably critical. The plastic bag cell illustrated in Fig. 21a is favored by researchers due to its flexible design.176 The two electrodes in an in situ plastic bag cell are connected to Cu mesh and Cu or Al mesh, respectively. The metal mesh helps overcome skin depth issues that are related to radio frequency penetration through the metal foil. In specific, metal foil shields the material inside the cell from radio frequency (RF) pulses. Moreover, the metal mesh can create better adhesion to electrode materials. Between the two electrodes is an electrolyte-soaked separator (for liquid cell) or a solid electrolyte (for solid cell). Then the in situ cell is placed in the coil of an NMR probe and cycled in the magnet. However, when the cell is inserted in an NMR coil, there is lack of stable pressure in the whole cell. The resistance inside the cell is increased and the contact between the particles and the current collector is reduced. Based on this, pressure is applied to the cell by placing polytetrafluoroethylene (PTFE) plates outside the in situ cell and tightening the cell with Teflon tape (a stronger string/tape can be used) or by applying small clamps to plates. A vacuum sealer is available to seal bag cells to produce them with a more consistent pressure. Besides the plastic bag cell, coin cells and cylindrical cells are also employed in in situ NMR experiments.177 Due to sufficient temporal and spatial resolution, in situ MRI provides 3D images of species with specific concentration and NMR activity on the time scale of electrochemical reactions.


image file: d3mh00135k-f21.tif
Fig. 21 (a) Schematic illustration of the in situ NMR setup. Reproduced from ref. 176 with permission from Elsevier. (b) Pictures and schematic illustration of a symmetric Li|LGPS|Li cell placed in a home-made cylindrical cell for MRI. Reproduced from ref. 178 with permission from the American Chemical Society. (c) A fixed gap cell for synchronized operando characterization of acoustic transmission and ssNMR spectroscopy. Reproduced from ref. 179 with permission from Springer.

Fig. 21b illustrates pictures and schematic diagrams of an in situ cell placed in a home-made cylindrical cell for MRI.178 Two pieces of Li foil are pressed closely onto both surfaces of electrolyte (LGPS) to ensure intimate contact. The overall setup is transferred into a home-made cylindrical cell designed for MRI. The main body of the cylindrical cell is fabricated from electrochemically inert polyether ether ketone (PEEK), which allows penetration of radio frequency pulses for NMR and MRI experiments. High purity stainless-steel wires and Cu foil act as current collectors. Rubber O-rings are used to ensure air tightness. The assembled cylindrical cell is placed in an NMR probe to enhance sensitivity and minimize noise for both NMR and MRI acquisitions. Interestingly, a fixed gap cell for synchronized operando characterization of acoustic transmission and SSNMR spectroscopy is designed, as shown in Fig. 21c.179 Acoustic transmission probes change in interface contact mechanics, while ssNMR measures rates of microstructure formation. Such a cell is useful for characterizing mechanics and microstructures of the same cell under operation without noise from cell-to-cell variation. Two wide-band ultrasonic transducers are firmly bonded using epoxy to either side of a machined PTFE cell holder. An acoustic gel coupling agent is placed between the transducer face and PTFE cell surface. The fabricated cell is mounted in a cell holder. A cell-grade Cu mesh current collector replaces Ni foil. Ultrasonic transducers are attached to shielded electrochemical cables and cables are attached to protruding Cu mesh current collectors by using a small amount of Sn solder between the exposed wire and edges of Cu mesh. The cable connected to the bottom electrode is passed through the magnet, while the cable connected to top electrode is passed through top of the magnet. Since no cable crosses the electrolyte area, the signal-to-noise is lowered. Before operando tests, the cell potential is checked with a multimeter to verify that good contact is achieved.

Ex situ NMR technique offers distinct advantages in investigating ionic movement in LLZO-based SSEs. It provides an in-depth comprehension of the microstructure and ionic transport characteristics (including the transport rate and pathway) in SSEs on an atomic scale. The ionic conductivity is a vital parameter in the design of high performance LLZO. It is well known that the ionic conductivity of cubic LLZO is two orders of magnitude higher than that of tetragonal LLZO. The stabilization of the cubic phase at room temperature is needed to dope various elements. NMR spectroscopy can reveal doping mechanisms and site occupancy by doped elements. 27Al or 71Ga NMR spectra indicated that the Al or Ga element is preferentially occupied at tetrahedral 24d sites instead of octahedral 96h or La sites in the LLZO structure.18017O NMR spectra showed distinct and well-resolved resonance signals which were attributed to O bound to the doped elements (Al or Ga). Consequently, the occupation of Al or Ga at Li sites instead of its presence at grain boundaries is mainly responsible for stabilizing cubic LLZO. Introduction of heterovalent cations generally causes charge imbalance and Li redistribution. Doping of high-valent cation reduces the total amount of Li+ in LLZO, while doping of low-valent cation enhances the Li+ content. 6Li NMR spectra displayed that replacing Zr with W reduces Li+ occupancy at octahedral 48g sites.181 Replacing La and Zr with Ca and W, respectively, increased Li+ occupancy at 96h sites.101 Replacing Li, La, and Zr with Ga, Ba and Ta, respectively, enhanced Li+ occupancy at 24d sites and maintained constant occupancy at octahedral 96h sites.182 Such examples demonstrate that doping with ions of various valence states and at different contents of cations can change the Li+ occupancy at Li sites, resulting in lower migration barriers, lower activation energy, and higher ionic conductivity. Notably, doping with different cations modulates the occupancy of Li sites in a distinct manner. The systematic elaboration of the modulation mechanism in the cases of doping with different cations required further investigation. Moreover, Li+/H+ exchange occurring on the LLZO surface also changes the Li+ occupancy at Li sites and total content.183

For anion doping, calculations suggest that F is favored to replace O2 and O3 sites in energy stabilization compared to Cl and Br.184 However, SSNMR only confirmed the presence of F in LLZO. The precise position of F in the structure could not be located. Importantly, the optimal doping content was determined via7Li MAS NMR.185 The smaller the full width at half maximum (FWHM) in spectroscopy, the higher the ionic conductivity. The ionic conductivity is also affected by transport pathways and transport rates. For Li+ transport pathways within LLZO, NMR relaxation time measurement and stimulated echo NMR technique indicated the lowest activation energy for Li+ diffusion following 96h–24d–96h' pathways.186 Moreover, 2D EXSY NMR demonstrated an ionic exchange between 24d and 96h sites and suggested that the Li+ transport pathway is 24d–96h–48g–96h–24d.181 Both transport pathways are essentially identical. The Li+ migration rate is slowest at 48g sites, which is considered as the deciding step to constrain ionic conductivity. Nonetheless, the NMR technique was unable to exclude direct jumps between two neighboring 96h sites with lower frequency and higher activation energy.187 Li+ transport rates in LLZO can be accessed using dynamic NMR at different temperatures.172 The line width of 7Li signals in Ga- or Al-doped LLZO begins to narrow at a temperature of 180 K compared to 280 K for undoped LLZO, indicating rapid ionic movement in Ga- or Al-doped LLZO.

Besides LLZO substances, Li+ transport pathways in CEs consisting of LLZO and organic polymers are not clear. To resolve this issue, NMR technique has been used to reveal Li+ transport pathways in several CEs. Introduction of ionic liquids generally enhances both ionic conductivity and mechanical properties in CE.188,189 Unexpectedly, variable temperature NMR and 2D EXSY NMR techniques indicate that chemical exchange between ionic liquids and LLZO particles is absent.188,189 Enhancement of ionic conductivity is attributed to Li+ migration preferentially in ionic liquids while LLZO particles provide mechanical strength. Consequently, CE consisting of inorganic ceramic particles and ionic liquids is possibly not essential for expensive active ceramics such as LLZO. When LLZO primary particles are uniformly dispersed in PEO, Li+ migration is transferred from PEO polymers to LLZO ceramic particles with increasing LLZO content as confirmed by 6Li NMR and 2D EXSY NMR.190 In specific, Li+ migration occurs mainly through LLZO ceramic particles rather than the LLZO/PEO interface or PEO polymer as LLZO content is 50 wt%. Moreover, a spontaneous Li+ exchange procedure between PEO polymers and LLZO particles resulted in a homogeneous Li+ diffusion and minimized the current gradient in CE.191 The protonated LLZO surface with the LiOH phase also exchanged Li+ with bulk LLZO and PEO polymers.192 Besides PEO polymers, several polymers such as polyvinylidene fluoride (PVDF), PAN and poly(ε-caprolactone-co-trimethylene carbonate) (PCL-PTMC) have been tested for interactions with LLZO particle surfaces for initiating rapid Li+ exchange.193 Since LLZO particles in the CE system exhibit intense alkalinity, partial dehydrofluorination of PVDF matrix was revealed by 1H NMR, which reinforces the interactions between PVDF and LLZO. Analogously, partial dehydrocyanation of PAN polymer skeletons formed local conjugated structures, resulting in continuous Li+ rapid conduction pathways at the LLZO/PAN interface. Interestingly, NMR studies suggest that preferential Li+ transport pathways in CE can be artificially controlled by utilizing several artificial modification approaches. Etching of LLZO particle surfaces to eliminate inherent resistive layers facilitates Li+ transport along locally modified interfaces and neighboring environments.194 When LLZO is present in nanowires rather than particles, Li+ diffusion preferentially occurs in the LLZO/polymer interface region.195 Construction of novel polymer skeletons such as fluoroboron-centered Li-conductive polymer frameworks (LiBFSIE) enabled the Li+ transport to be mainly dependent on LLZO particles and LiBFSIE/LLZO interfaces.196 With the incorporation of additives such as plasticizer tetraethylene glycol dimethyl ether (TEGDME) or organic plastic salt (OPS), Li+ was principally transported via additive-associated phases instead of LLZO or polymers (Fig. 22a).190


image file: d3mh00135k-f22.tif
Fig. 22 (a) 6Li NMR comparison of pristine and cycled LLZO with various contents or additive TEGDME. Reproduced from ref. 190 with permission from the American Chemical Society. (b) Li dendrite growth in LLZO visualized by 7Li NMR CSI. Reproduced from ref. 198 with permission from the American Chemical Society. (c) Voltage profile normalized peak intensities and 7Li chemical shift (ppm) of the Li metal region in operando MRI testing. Reproduced from ref. 179 with permission from Springer.

Incompatibility and electrochemical instability at the electrode/electrolyte interface are the main reasons that have limited practical applications of LLZO-based SSLBs. In order to comprehend the root causes of complex and transient reactions at the interface, in situ or ex situ NMR techniques have been developed and applied. On the anode side, Li deposits with differing structures were identified using NMR resonance displacement197 since the characteristic peak value for dendrite Li is ∼270 ppm, while that for mossy Li is ∼261 ppm. Furthermore, the reduction of peak intensity in the NMR spectrum indicates enhancement of Li deposition. Accordingly, NMR technique demonstrated that mossy Li dominates initially, while dendrite Li growth occurs in later stages.58,197 The LLZO/Li anode interface is generally accepted to be a kinetically stable interface but the major issue with this interface is Li dendrite growth. Owing to different resonance displacements, the growth evolution of Li microstructures in the Li |LLZO| Li symmetric cell was visualized by 7Li NMR chemical shift imaging (CSI).198 As shown in Fig. 22b, although the current density was lower than 0.5 mA cm−2 and the voltage profile was not characterized by short-circuiting, Li dendrites were generated and progressively accumulated to connect both electrodes ultimately. Such a method can be further employed to investigate electrolyte thickness, current density or stacking pressure effects on Li dendrite growth. The effect of stacking pressure on interface contact and Li microstructure growth was examined by SSNMR CSI.179 At a stacking pressure of 2 MPa, continuous void formation at the interface caused significant surface roughening. Above a stacking pressure of 7 MPa, Li microstructure growth was alleviated due to creep-driven interfacial dynamics. Although contact depletion was reversibly recovered at stacking pressure up to 13 MPa, the likelihood of fracture induced by local crack propagation was increased. Subsequent operando MRI illustrates irreversible Li microstructure (7Li resonance peak centered at ∼258 ppm) formation during cycling (Fig. 22c). The stacking pressure of an operando cell is <1 MPa as tested by EIS. Therefore, searching for an appropriate stacking pressure is necessary to avoid cell short circuits. In addition, Li/LGPS interface evolution during cycling was investigated using MRI technique.178 Images show that Li+ is non-uniformly distributed at the Li/LGPS interface after cycling, resulting in increased interface resistance and local hotspot development. PEO-coated LGPS demonstrated a uniform distribution of Li+ at the interface, which lowered the interface resistance and suppressed dendrite growth. On the cathode side, NMR technique revealed that a continuous Li+ transport pathway was formed within the interior of the composite cathode and this consisted of LLZO particles and cathode active material, reducing the Li+ diffusion distance and lowering the interface resistance.19931P MAS NMR spectra showed elemental diffusion in composite cathodes comprisingLi3+xP1−xSixO4 electrolyte and LFP active material during heat treatment.200 Fe in LFP diffused into Li3+xP1−xSixO4 leading to the formation of Li3−zFezP1−zSizO4. The interface transport rate between Li2S active material and Li6PS5Br electrolyte under various preparation conditions was quantitatively analyzed by 7Li 2D EXSY NMR.174 Preparation steps such as nanosizing and ball milling can accelerate Li+ exchange at the interface for improving ionic transport performance. However, electrochemical cycling remarkably changed such an exchange behavior, leading to lower ionic transport. Thus in situ NMR techniques are expected to be useful in investigating ionic transport at LLZO/cathode interfaces during cycling in future studies.

Overall, in situ and ex situ NMR techniques serve as powerful tools to research non-transparent samples, namely, LLZO-based SSLBs. In situ NMR technique with non-invasive character can simultaneously provide variation in ionic concentration at interfaces, Li dendrite formation and microstructure evolution. However, the resolution of NMR technique for solid substances was lower than for liquids due to the wider NMR spectrum.201 In particular, MRI offers lower temporal and spatial resolution and is thus unable to image topographic changes in real-time as in situ EM, X-ray imaging or AFM. Improving the resolution will require a larger magnetic field gradient or the use of more advanced NMR pulse technique.

3.5.2 In situ/operando electron paramagnetic resonance (EPR). EPR, a highly specific technique similar to NMR, detects unpaired electrons or free radicals.202 The working principle is unpaired electron spin rather than nuclear spin in the analysed substances with an applied magnetic field.203 Due to the utilization of microwave radiation, EPR features more sensitive and accurate detection capability compared to NMR for Li microstructure evaluation. However, it is impossible to quantitatively analyze Li metal due to its low surface detection depth of ∼1 μm.204 Importantly, the material used to construct in situ/operando electrochemical cell should be compatible with all cell components (such as electrolyte or electrodes), transparent to microwave radiation and inactive to EPR. The inset in Fig. 23 briefly shows the in situ/operando EPR cell structure and this appears similar to the in situ NMR cell.202 The body material of the entire cell is made up of hydrophobic polychlorotrifluoro ethane. Al and Cu discs welded to wires are employed as positive and negative current collectors, respectively, while an EPR transparent Teflon O-ring is used to seal the cell.
image file: d3mh00135k-f23.tif
Fig. 23 (a) In situ EPR spectra of the full cell at OCV and after charging to 3.6 V. (b) In situ EPR spectra of the Li symmetric cell. (c) Operando EPR image of the full cell, which is cycled between 2 and 4.6 V with images taken at voltages defined by letters on the voltage–composition curve by maintaining the cell at rest at these potentials. Reproduced from ref. 202 with permission from Springer.

Although in situ/operando EPR is a powerful analytical tool for cell electrode characterization, it is currently seldom employed in LLZO-based SSLBs. The LLZO surface state after molten Li treatment was investigated using ex situ EPR.205 Results suggested that the LLZO surface exhibits unpaired electrons trapped in oxygen vacancies, resulting in chemical coloration (from initial brownish white to grayish black) and enhanced electronic conductivity. As a result, prolonged molten Li manipulation increased the risk of Li penetration. Ex situ EPR also confirmed the oxidation of the Mn element in LLZO.206 The signal intensity of EPR peaks increased with increasing Mn content, and the calculated g-factor was consistent with the Mn2+ valence state. Moreover, in situ EPR was utilized to visualize the growth of Li microstructures in a liquid cell.202 As shown in Fig. 23a, the EPR spectrum is featureless at OCV. By charging the cell to 3.6 V, a remarkably sharp and slightly distorted signal (g = 2.0023) of 1.5 G line width appeared. Such sharp signals are attributed to small metallic Li aggregates deposited on Li foil during charging. EPR spectra of a Li symmetric cell also revealed the presence of Li aggregates as a bias voltage was applied (Fig. 23b). Subsequently, operando EPR imaging tests were carried out. As shown in Fig. 23c, each pixel in the image is an EPR spectrum. Red, green, and blue colors indicate zones of high, medium, and no-spin electron concentration, respectively. Red regions correspond to Li aggregates, suggesting inhomogeneous Li microstructural deposition. Li dendrite formation was attributed to excessively inhomogeneous Li deposition. Incorporation of fluoroethylene carbonate (FEC) additives into the electrolyte dramatically reduced Li microstructural growth.204

Owing to limitations in terms of cell design and compatibility, the application of in situ EPR to SSLBs is at an early stage. In situ EPR technique is expected to investigate surface/interface issues in LLZO-based SSLBs with continuous development and improvement in the future. In particular, in situ EPR can provide semi-quantitative information on time-resolved Li metal deposition and can be an important supplement to in situ OM, EM, and NMR. This technique shows potential for real-time monitoring of Li plating/stripping. In addition, EPR imaging is still limited to micrometer resolution, and requires long times of >5 days for image collection. With further improvement in resolution, interface evolution can be imaged in real time to explore unknown information on Li dendrite nucleation sites.

3.6 In situ/operando neutron based techniques

Neutron technique, a powerful tool for investigating material and cell systems, enables exploration of structures and dynamics of substances over wide time and length scales. As uncharged particles, neutrons interact directly with the nucleus and generally not with paired electrons. As a result, neutrons exhibit non-destructive and superior transmittance (up to several cm) to cell materials, which is favorable for operando or in situ research. Neutron scattering length or the cross-section of a nucleus is decided by both isotopes and atomic numbers. Consequently, several light elements (such as H and Li) feature high neutron scattering cross-sections, even higher than those of some heavy metals. This indicates that positions of light elements in crystal structures can be directly determined, or images of light elements are available in the background along with heavy elements. Neutron scattering length is also quite varied in different isotopes of the same element or between neighboring elements, resulting in differentiation of neighboring transition metal elements, investigation of doped elements and recognition of isotopes. In addition, since wavelength, energy, and momentum are comparable to interatomic distance, vibrational excitation energy and diffusion excitation energy, respectively, neutrons are ideal for investigating acoustic and transport mechanisms. Until now, neutron-based analytical techniques include neutron powder diffraction (NPD), neutron total scattering (NTS), small-angle neutron scattering (SANS), neutron reflectometry (NR), neutron depth profiling (NDP), neutron imaging (NI), inelastic neutron scattering (INS) and quasielastic neutron scattering (QENS).207 NPD, NTS, and SANS have been utilised for analysing crystal structures (including space groups, lattice constants, and atomic coordinates), local structures (including pair distribution functions, nearest neighbor distance, and coordination) and nanostructures (including size, morphology, and distribution) of materials, respectively. NR can analyse surface and interface structures. NDP determines the distribution of an element as a function of depth in a solid while NI involves imaging of materials by utilizing contrast in neutron transmission/attenuation. INS and QENS are employed to investigate vibrational dynamics (such as phonon dispersion and density of states) and diffusive dynamics (such as diffusion coefficients and jump length) of materials, respectively. Therefore, in situ neutron techniques are critical to reveal surface/interface changes, Li dendrite growth, and Li+ distribution and transport pathways. This section mainly discusses the application of NDP and NI in LLZO-based SSLB research.
3.6.1 In situ/operando neutron depth profiling (NDP). NDP is a non-destructive surface/interface probe technique for measuring elemental distributions at depths using low energy neutrons (4 meV). This utilizes cross-sectional capture reactions between thermal neutrons and specific isotopes. Significantly, NDP technique exhibits high sensitivity to light elements (low atomic number elements such as H, Li and B). Such elements, especially Li,208 are not easily characterized using conventional techniques such as EDS, XPS and XRD.209 As shown in Fig. 24a, as neutrons travel through a Li-rich sample in a vacuum chamber, neutrons react with 6Li isotopes to form 4He (α) and 3H (tritium) particles according to the following equation: 6Li + n4He (2055.51 keV) + 3H (2727.88 keV).2104He and 3H particles lose energy while traveling through the measured material at a specified rate and are collected by a surface barrier detector. Particle counts with various energies are recorded using a multi-channel analyzer to identify positions of reactions and depth information of Li atoms. Naturally, NDP can quantify the Li concentration and establish a functional relationship with depths perpendicular to the electrode surface.211 The depth analysis of Li using NDP has been achieved for distances of several tens μm with depth resolutions of several tens nm.210 The precise value is dependent on sample atomic composition, setup geometry, and detector mass density. Therefore, due to high specific resolution along depth orientation, non-destructive measurements, high detection sensitivity to light elements and specific quantification, in situ NDP is a powerful technique to investigate Li+ migration, distribution, and dynamic processes in LLZO-based SSLBs.
image file: d3mh00135k-f24.tif
Fig. 24 (a) Schematic diagram for NDP investigations. (b) Schematic of the in situ NDP system. (c) Configuration and digital image of an in situ cell. Reproduced from ref. 209 with permission from the American Chemical Society. (d) Setup and sample configuration for in situ NDP measurements with top and bottom dual Si detectors. Reproduced from ref. 212 with permission from Elsevier.

Fig. 24b illustrates the setup for in situ NDP measurement.209 An in situ cell is connected to a temperature-controlled aluminum plate and disc in a vacuum chamber. 4He and 3H particles formed by neutron beam and 6Li reaction are detected using a Si detector. In general, the in situ NDP measurement setup is tested at higher temperatures (such as 90 °C) to ensure large Li+ transfer during short timespans, which permits higher applied current densities. Fig. 24c displays the configuration and digital image of a symmetric/asymmetric in situ cell. The symmetric cell mimicked the structure with a Li metal anode and a sulfur-type Li-free cathode, while the asymmetric cell mimicked the structure with a Li-free anode and a Li-rich cathode. Li metal was melted onto the LLZO surface. During melting, a Ti strip with a punched hole was attached to the Li electrode, functioning as both a mask and a current collector. CNT films were coated on the LLZO surface facing the detector by a solution-based process. Approximately 50 nm thick nickel was deposited by electron beam evaporation on the electrode to ensure favorable electrical contact. Meanwhile, an approximately 7.6 μm thick Kapton film blocks 4He particles and improves depth resolution. Without the blocking effect of Kapton films, 4He and 3H signals would overlap in the low channel range, and significantly affect depth resolution of NDP measurements. Fig. 24d illustrates the setup and sample configuration for in situ NDP measurements with top and bottom dual Si detectors.212 Thin Cu/Au composite bilayers of current collection were deposited on both sides of SSEs by magnetron sputtering with thicknesses of Cu and Au being 30 and 10 nm, respectively. In a vacuum chamber, the in situ cell was fixed to a sample holder made of polyethylene terephthalate glycol (PETG). The inclination angle of the sample holder was ∼5° with respect to the neutron beam plane. Dual detectors enable simultaneous measurement of NDP spectra on both sides of the cell.

The in situ NDP technique is mainly used to investigate the electrolyte/electrode interface behavior. It is expected to address the issues of high interface resistance and low interface stability in LLZO-based SSEs. At first, in situ NDP quantitatively analyzed the spatial heterogeneity during Li plating/stripping.213 Results indicated that initial plating current density significantly affects subsequent cycle processes. Li deposition density is high at higher initial plating current density, while low density and thick surface Li deposition cause massive “dead Li” generation at lower density. The plating/stripping behavior of Li in the Li |LLZO| Li symmetric cell and Li |LLZO| 3D CNT asymmetric cell was monitored in real time using in situ NDP technique.209 CNT was adopted as a Li storage structure for mimicking Li-free metal anodes. As shown in Fig. 25a and b, the counts of Li significantly increased only near the end of cycling in a symmetric cell, while the counts increased and gradually accumulated at the beginning of cycling in an asymmetric cell. Such a behavior is attributed to the formation of a reversible layer around the LLZO/CNT interface, where reversible Li plating/stripping occurs. However, beyond this layer, the reversible behavior is lost, and Li becomes “dead Li” that constantly accumulates and grows in CNT. Since CNT surface energy ranges between ∼2150–2500 keV, the counts around 2375 keV increased significantly. Moreover, the short-circuit behavior of the symmetric cell was predicted from changes in in situ NDP spectra. The integrated NDP counts close to zero indicated that Li is reversibly plated/stripped in the cell. In contrast, accumulation of Li was induced by an unstable electrochemical behavior, resulting in increased integrated NDP counts. An abnormal increase in NDP counts occurred before the voltage profile changed significantly, which led to a short circuit in the cell. Interestingly, the cell experienced unilateral short-circuiting if NDP counts decreased slightly at each stripping stage but increased significantly at each plating stage. The Li deposition behavior in Li |LLZO| 3D Ti cells was also monitored in real-time using in situ NDP technique.214Fig. 25c displays the data that show a rapid increase in counts with time at 2560–2630 keV, indicating the growth of Li in voids. The results suggest that Li is preferentially deposited in the voids of the 3D Ti electrode, which lowers the interfacial stress induced by electrode volume variation and mitigates Li dendrite formation. Hence, an efficiently designed 3D mixed electronic/ionic conductive skeleton can buffer volume expansion, adjust the Li deposition behavior, enhance the contact area, reduce diffusion distance, and suppress Li dendrite formation. In addition to the work on LLZO-based SSLBs, the lithiation/delithiation behavior in LCO |Li3PO4| Si thin film cells was revealed by operando NDP.215 A Li immobilized layer was formed at the electrolyte/anode interface during cycling as Si migrates from the anode into the electrolyte. As the cycle proceeded, the Li immobilized layer grew constantly at a lower speed, which resulted in a reduction in the amount of movable Li+. Accordingly, the cell was not capable of extracting a sufficient amount of Li during normal operation and the storage capacity thus significantly decayed. Interface modification strategies are beneficial to realize reversible and homogeneous Li plating/stripping. Depositing an Al2O3 layer on the LLZO surface by atomic layer deposition (ALD) technique resulted in the formation of an interface region containing high Li concentration compared to bare LLZO.216 High Li concentration ensures consistent charge transfer at anode/electrolyte interfaces, enabling excellent cell-rate performance. Similarly, a ZnO layer has been deposited on the current collector surface using the ALD technique.217


image file: d3mh00135k-f25.tif
Fig. 25 In situ NDP measurement of (a) the Li |LLZO| CNT asymmetric cell and (b) the Li|LLZO|Li symmetric cell while cycling. Reproduced from ref. 209 with permission from the American Chemical Society. (c) Contour map of in situ NDP spectra of the Li |LLZO| 3D Ti cell collected during cycling. Reproduced from ref. 214 with permission from Elsevier. (d) Comparison of the Triton edges in various states corresponding to in situ NDP spectra. Reproduced from ref. 212 with permission from Elsevier.

The electrochemical deposition behavior of Li in Cu (with/without the ZnO layer) |CE| Li cell was investigated using operando NDP technique. Li concentration and plating thickness on a bare Cu current collector were remarkably increased during the cycle, indicating that inactive Li species were accumulated on the CE side. In contrast, a ZnO layer-coated Cu current collector improved the lithiophobic behvaiour and mitigated the accumulation of inactive Li species. The plating/stripping efficiency calculated from Li concentration distribution was 45% for the Cu current collector and 80% for the ZnO layer coated Cu current collector.

Researchers tend to concentrate more on ionic conductivity of SSEs in SSLBs while neglecting the effect of electronic conductivity. Dynamic evolution of Li concentration distribution of SSEs in LCO |LiPON| Cu, Li |LLZO| Cu and Li |LPS| Cu cells during Li plating was monitored in real time with in situ NDP technique.218 Massive Li deposits nucleated and grew into Li dendrites inside LLZO and LPS. In contrast, Li concentration in LiPON did not significantly vary, suggesting the absence of internal Li dendrite formation. The high electronic conductivity of LLZO and LPS is a major cause of internal Li dendrite formation. Therefore, identifying a source of high electronic conductivity and lowering the value is imperative. Furthermore, in situ NDP technique was utilized to track Li concentration in LCO |LiPON| Cu and LMO |LiPON| LNO thin film cells to understand Li migration.219 In LCO |LiPON| Cu cells, LCO was synthesized from 100% 6Li, indicating that the intensity of 6Li in the cathode was higher than that in electrolyte. During the charge process, intensity of Li in LCO was dramatically reduced, which is consistent with the depletion of Li. Notably, Li depletion at the cathode/electrolyte interface was more intense, which was attributed to 6Li migration from LCO to electrolyte and charging reactions. The NDP spectra of charged and discharged LMO |LiPON| LiNiO2 cells showed the largest variation in Li concentration at both electrodes and the smallest variation at the electrolyte. The distribution of Li concentration in the cell is non-uniform and concentrated around the sample center. Consequently, in situ/operando NDP technique enables understanding of Li+ migration pathways and spatial distributions inside the cell via determination of the changes in Li concentration. Based on these features, the route of Li+ migration pathways in CE along polymers and inorganic ceramics or both interfaces can be determined by in situ/operando NDP technology. Recently, Li transport in a novel and prospective Li conductive glass-ceramic electrolyte Li2Al0.5Ti0.75Ge0.75Si0.5P2.5O12 (LICGC) was detected using in situ NDP technique.212 As shown in Fig. 25d, Li concentration at the interface between LICGC and Cu/Au bi-layer collector was dramatically reduced when +2.8 V was applied. Generation of a Li-depleted region was attributed to the formation of space charge layers at the interface and this was the first time that a practical distribution of Li atoms during space charge layer formation was observed. Interestingly, Li was gradually returned to a Li-depleted region when the bias voltage was eliminated. In addition, the changes in the Li migration and diffusion behavior with the application of reverse voltages required further investigation.

In summary, in situ/operando NDP technique achieves real-time monitoring of Li plating/stripping at electrode/electrolyte interfaces, contributing to an in-depth comprehension of the interfacial behavior in LLZO-based SSLBs. Moreover, thanks to high temporal and spatial resolution and high sensitivity, the Li deposition density can be quantified and Li migration pathways and spatial distributions can be analysed. However, NDP technique generally requires testing under vacuum and thus the reliability of the results in comparison to real-cell environments may not be accurate.

3.6.2 In situ/operando neutron imaging (NI). On atomic or nanometer length scales, the dynamic behavior of Li in SSLBs can be tracked using in situ NI technique. Similar to the principle of X-ray imaging, neutron imaging is based on selective attenuation of neutron intensity, mainly by absorption and scattering. Since neutrons without charge easily penetrate through substances and the total cross-section (scattering and absorption) of Li is quite large, Li can be easily contrasted against other elements in imaging. This property compensates for the insensitivity of XRI to Li. Crucially, the NI technique is a valuable supplement to NDP and other neutron techniques, achieving real-time imaging for Li microstructures. NI techniques include neutron radiography imaging (NRI) for 2D imaging and neutron tomography imaging (NTI) for 3D imaging. An achievable spatial resolution is ∼100 μm. Hence, in situ NI technique can provide information related to the evolution of crystal structures, variation of surfaces/interfaces, deposition behaviors and the spatial distribution of Li, all of which are beneficial to interpret issues correlated with cell design and electrochemical performance. Fig. 26a shows the schematic diagram of an in situ NI electrochemical cell.220 The sandwich structure of the cathode, separator, and Li metal anode is utilized to ensure a similar electrochemical behavior to that of a coin cell. The core part of in situ cell design is a thick cathode (load ≈ 200 mg, diameter = 10.6 mm, and thickness ≈ 8 mm) to ensure massive Li transportation and good neutron visualization. A glass fiber wall is attached inside the cell body to prevent a short circuit caused by contact with the conductive Ti–Zr alloy. Fig. 26b displays an experimental setup and a spatial relationship between the cell and the neutron beam. Similar to the XRI technique, the two electrodes are connected to an electrochemical workstation for charging and discharging tests. Neutron attenuation contrast for mapping the Li distribution was achieved using a CCD detector. The 3D tomography was derived by measuring the 2D radiography of the sample at different rotation angles (Fig. 26c).
image file: d3mh00135k-f26.tif
Fig. 26 (a) Schematic diagram of an in situ NI electrochemical cell. (b) Experimental setup for NI measurements. (c) Schematic diagram of 2D radiography and 3D tomography. (d) 3D evolution of the Li distribution in the cell at different stages of charging and discharging. Reproduced from ref. 220 with permission from the American Chemical Society.

In situ NI technique revealed the growth mechanism of Li dendrites and distribution dynamics during cell short-circuits.220Fig. 26d shows the schematic diagram of an in situ cell setup and 3D evolution images of Li distribution during cycling. In situ NI images coupled with electrochemical profiles confirmed that internal short-circuiting of the cells was initiated from Li dendrite growth between the LMO cathode and Li anode. A competition mechanism between self-discharge and charging after a cell short-circuit is proposed. During the charging process, contact between Li and Li1−xMn2O4 leads to a short-circuit. As a result, further charging is not possible and the voltage drops. Due to voltage differences, Li1−xMn2O4 obtains electrons to form the LMO cathode, resulting in contact loss with the Li dendrite. Cell self-discharge is thus terminated and then cell undergoes recharging as indicated by the increase in voltage. Such a mechanism can reasonably explain the abnormal voltage drop/rise after the cell short-circuit. Thus the use of this technique provides an in-depth insight into short-circuiting induced by Li dendrite formation along with the associated effects which need to be monitored to ensure a safe design and operation of the cell. The effect of Li morphology evolution on cell electrochemical performance was clarified by in situ NI and XRT techniques.221 Due to the electrochemical side reactions, irreversible structure transition of Li leads to performance degradation. A greater comprehension of cell potential degradation and failure mechanisms can provide relevant information for improving the future cell design. Unfortunately, in situ NI technique is at a very early stage for application in solid-state batteries. With further development of this in situ technique, it is expected that the growth mechanism of Li dendrites as well as migration pathways and spatial distributions of Li in LLZO-based SSLBs can be determined to a higher precision and detail in the future.

Besides NDP and NI techniques, NR technique has been utilized to explore surfaces/interfaces with the aid of total internal reflection phenomenon. Due to high penetration, non-perturbation, and high sensitivity of neutrons, in situ NR technique exhibits remarkable advantages for identifying thin film layers (such as the SEI or modified artificial layer) on surfaces/interfaces. For example, formation of SEI films at the solid–liquid interface was investigated by in situ NR.222 SEI film generation was spontaneous and featured a bi-layer structure. Low scattering length density of the inner layer suggested that Li2O was the major component. In contrast, scattering length density of the outer layer was proportional to that of solution indicating that solution infiltrates pores or layers containing solvated H. Therefore, in situ NR technique is expected to be employed for investigating the evolution of thin film layers on the surface/interface in LLZO-based SSLBs during electrochemical response. Notably, the requirements of intense signals, long analysis times (namely, low temporal resolution), and constrained characterization facilities limit its extensive application. SANS technique provides information on the size and morphology of nanostructures based on the scattering length density fluctuation being on the order of 1 nm–0.5 μm. The growth of Li filaments in LLZO was monitored in real-time on a nanoscale with operando SANS technique.223 Growth of Li filaments was not seen to occur just by accumulation but relies on competition between the growth rate and self-healing rate. Increasing the temperature can strengthen self-healing ability, which positively suppresses Li filament growth. Accordingly, heat treatment assists in enhancing the cell electrochemical performance. Overall, neutron based in situ techniques are critical in LLZO-based SSLB investigations (such as evolution of surface/interface structures, growth of Li dendrites, and spatial distributions of Li species) due to the high sensitivity to Li and high penetration depth. However, neutron flux is too small to analyze rapidly and the neutron source is too expensive for large-scale laboratory applications. Moreover, the low abundance of 6Li increases the experimental complexity. More importantly, neutron technique-analyzed samples exhibit radioactivity, posing threats to humans and the environment and these issues need to addressed in the future research.

3.7 Other techniques

Besides the abovementioned in situ/operando methods, several other techniques for characterizing LLZO-based SSLBs have been developed, including time-of-flight secondary ion mass spectroscopy (ToF-SIMS), electrochemical impedance spectroscopy (EIS), acoustic characterization, chemo-mechanical measurement, and atom probe tomography (APT). These techniques are briefly discussed in this section.
3.7.1 Time-of-flight secondary ion mass spectroscopy (ToF-SIMS). ToF-SIMS technique utilizes a focused ion beam to steadily bombard atoms or atomic clusters on the sample surface. Separation of ionized secondary ions (such as sputtered atoms, molecules, or atomic clusters) is achieved according to mass-to-charge ratios. The elemental composition and distribution of sample surface are determined by collecting mass spectroscopy separation of secondary ions.224 ToF-SIMS technology offers several advantages. First, it provides multi-element analysis data and secondary ion imaging for a specific elemental distribution both on the surface and to a certain depth when ion etching techniques are used. Second, since full spectra is achieved with one pulse, high ion availability and virtually non-destructive static analysis on sample are possible. Lastly, as ion time of flight is exclusively dependent on mass, lowering the pulse repetition frequency can broaden the mass range. Notably, similar to in situ EM and XPS techniques, ToF-SIMS needs a vacuum environment and the analysis has to be setup to detect specific elements beforehand. Currently, ToF-SIMS technique is extensively used to assess nano-scale chemical information of LLZO, anode/LLZO interfaces, and cathode/LLZO interfaces in LLZO-based SSLBs.

Coupled ion etching and ToF-SIMS techniques for depth orientation measurement are employed most frequently. The application scenario and derived outcomes are comparable to those of XPS depth profiling tests. The preparation of LLZO thin films using CO2-laser assisted chemical vapor deposition was studied using ToF-SIMS technique.149 The real elemental distribution in thin films can be detected by monitoring the variation of positive/negative ionic fragments related to LLZO and substrate in-depth orientation. The homogenous elemental distribution and intensity that matches the LLZO stoichiometry ratio suggested that LLZO thin films were successfully prepared. The film thickness was quantitatively analyzed via multiplication of the sputter time and rate. In addition, the distribution of doped elements in LLZO, as well as composition and thickness of the surface contamination layer, were also investigated using ToF-SIMS technique. At the anode/LLZO interface, introduction of different artificially modified interlayers was observed using ToF-SIMS technique.138,165,225–227 As shown in Fig. 27a, a LiPAA layer was successfully formed on the LLZO ceramic pellet surface.165 The good flexibility of the LiPAA layer relieves interfacial stresses and maintains interface contact. Meanwhile, the electronic insulation of the LiPAA layer shields LLZO against electronic degradation, resulting in suppression of Li dendrite formation in the interior of LLZO. ToF-SIMS tests revealed that the lithiophobic behaviour of LLZO is intimately correlated to surface contaminants of Li and LLZO.146,228 Li surface contaminants mostly consisting of Li2O, Li2CO3, LiOH or LiF are the main causes of poor interface contact. Although LLZO surface impurities increase the interfacial resistance, their effect is less significant compared to Li surface impurities. An effect of introducing an artificially modified interlayer is to determine the impact of Li and LLZO surface impurities. At the cathode/LLZO interface, elemental inter-diffusion and decomposition phenomena during high-temperature melting or annealing have been revealed by ToF-SIMS technique.100 Near the interface, LLZO was converted from a cubic to tetragonal phase while interlayers such as LBO or lithiated Nb2O5 minimized elemental cross-contamination and stabilized the cubic phase.229,230 In addition to depth orientation measurement, cross-sectional ToF-SIMS mapping images of CE-based cells before and after cycling indicate that Li+ preferentially migrates along the amorphous phase in the polymer matrix.231 Furthermore, ToF-SIMS mapping images also reveal the presence of Li dendrites in LLZO ceramic pellets after cycling.28


image file: d3mh00135k-f27.tif
Fig. 27 (a) ToF-SIMS depth profiles and 3D views of the sputtered volume for LLZTO@PAA pellets. Reproduced from ref. 165 with permission from Springer. (b) Pressure-dependent Nyquist plots show external force's large impact on the interface resistance and capacity values. Reproduced from ref. 234 with permission from the American Chemical Society. (c) Configuration for operando acoustic transmission. Reproduced from ref. 244 with permission from Elsevier. (d) Fixed pressure configuration for operando acoustic transmission experiments. (e) Normalized acoustic amplitude (intensities relative to first waveform intensity) at 2, 7.4 and 13 MPa stacking pressures. Reproduced from ref. 179 with permission from Springer.

When combined with FIB technique to investigate interior compositions, a xenon ion source should be used as it provides low ionic contamination, high sputtering yield, low scattering and greater precision for Li mapping compared to a Ga ion source.232 However, milling of the surface obtained using a xenon ion source is not as smooth as seen when using a Ga ion source. Moreover, in situ ToF-SIMS technique that can measure under an electrochemical response is not yet available. Therefore, with continuous development, in situ ToF-SIMS technology is expected to be employed to monitor the real-time evolution of surfaces/interfaces during electrochemical cycling.

3.7.2 In situ/operando electrochemical impedance spectroscopy (EIS). EIS, also known as alternating current (AC) impedance spectroscopy, adopts a small amplitude sinusoidal current or sinusoidal voltage as the disturbance signal and the electrode system responds in an approximately linear manner to it. The ratio of AC potential to current signal, i.e., impedance, is measured over a wide frequency range to investigate the internal dynamic process of the electrode system. EIS is a classical, practical and convenient method for studying the electrode system, which provides several advantages as follows: (1) simultaneous characterization of individual electrochemical processes according to differing relaxation time constants over a wide frequency range; (2) as a linear research method, data handling is comparatively easy; (3) in situ and on-line testing can be conducted; (4) this can be widely applied in industrial manufacturing due to simplicity of the test method; (5) special experimental skills and methods are not needed; (6) the instrument is inexpensive and does not need additional auxiliary equipment; and (7) the detection procedure is non-destructive and chemically inactive. The capacity fade mechanism, Li insertion mechanism, electrode process dynamic parameters, state of charge (SOC)/state of health (SOH) of cells, internal resistance and correlated factors affecting cell performance have been investigated by EIS fitting analysis with an equivalent circuit.233 Importantly, distribution of relaxation time (DRT) technique has progressively attracted attention in order to address the issue that prior hypothesis is required for EIS data handling using an equivalent circuit. In general, various electrochemical processes feature distinct characteristic time constants. The distribution of typical time constants in EIS spectra is identified using the DRT technique. Hence, as an option for resolving EIS data or assisting equivalent circuit fitted analysis, the DRT technique enables more precise elucidation of relatively complicated electrochemical processes. More importantly, in situ cells for EIS measurements are not necessarily specially designed, and are available as coin cells, three-electrode cells or pouch cells. Additionally, in situ EIS technology can be used to test in air instead of vacuum. Therefore, it can be utilized on a large scale in laboratories compared to other in situ/operando techniques.

Recent in situ EIS research on LLZO-based SSLBs mainly concentrated on the interfacial properties between LLZO and Li anodes or cathodes. On the anode side, in situ EIS examination revealed that vacancy diffusion in Li metal constrains the rate capacity of the Li anode.234 Contact loss is induced by void accumulation and pore formation at the LLZO/Li interface, resulting in high interface resistance. As current density increases, both the increased rate of interface resistance and the growth rate of Li dendrites are enhanced.235 In order to lower the interface resistance, several strategies, including application of high stacking pressure, construction of Li alloy anodes, and introduction of artificially modified interlayers have been investigated using the in situ EIS technique. First, as shown in Fig. 27b, interfacial resistance decreased to nearly 0 Ω cm2 at a high enough stacking pressure of ∼400 MPa.234 When the pressure was removed, the interface resistance was preserved. In addition, under high stacking pressure, a decrease/increase in interface resistance is reversible during plating/stripping. However, this cannot prevent Li dendrite formation intrinsically. Notably, high temperatures promote diffusion and plastic deformation of Li. Therefore, increasing temperature modestly will reduce the required stacking pressure, which will lower the cell manufacturing costs. Moreover, use of Li alloys such as Li–Mg alloy eliminates interface pores to lower the interface resistance.236 However, the Li–Mg alloy system suffers from kinetic limitations, namely, in terms of diffusion-controlled decrease of interface Li metal concentration. Analogously, in combination with in situ EIS and DRT techniques, a study of Li–In alloy system suggested that interface charge transfer and Li diffusion are the critical kinetic steps that assist in stabilization of the interface.237 Such kinetic characteristics are seen in almost all Li alloys. Increasing the Li content in Li alloys can induce anode potential ≥0 V and a low charge transfer barrier for Li rapid diffusion, which is vital to stabilize the interface. Lastly, introducing alloy elements to form an alloy interlayer helps to construct a fast charge transfer pathway at the interface by changing the critical kinetic step at the interface from Li diffusion to Li alloying.62 However the kinetics of Li plating/striping are not changed fundamentally. The interface resistance will still increase after long-term cycling, which results in cell failure.

Formation of inorganic substances such as SnF2 at the interface in situ in the form of a conformal interlayer was seen in the interactions of Li–Sn alloy and LiF.238 Since this layer is able to reinforce the adhesion to Li metal and has the function of an alloy layer, interface resistance is minimized. Electronically-insulating and Li+-conducting LiF layers suppressed internal Li dendrite growth initiated by electrons from Li metal into LLZO. On the cathode side, trace amounts of Li2CO3 on LLZO surfaces are decomposed at high potential (>3.8 V).239 The interface resistance was substantially elevated due to the release of CO2 and O2 from the LLZO/cathode interface. Removal of Li2CO3 by soaking LLZO in a LiBF4-based liquid electrolyte in a glovebox before cell assembly was demonstrated using EIS technique.240 The reaction products are CO2 gas and soluble B2O3 and LiF, and the interface resistance is dramatically decreased. Therefore, assembling an effective and contaminant-free LLZO-based SSLB is necessary. In addition, at high temperatures, an apparent cycle-induced reaction between LMO cathode and LLZO was revealed by utilizing in situ EIS technique.134 Due to migration of the Mn element from LMO to LLZO, the interface resistance enhanced and the capacity decayed. The introduction of an artificially modified interlayer is thus an effective approach to mitigate element diffusion. Since EIS technique cannot be used to image the interface in real-time, in situ imaging techniques are required as complementary parts for analysing interfacial characteristics.

3.7.3 In situ/operando acoustic characterization. In situ/operando acoustic characterization has been employed to detect variations in cell modules, inflation, and density in conventional liquid cells.241 Sound waves with frequencies of 1–10 MHz are transmitted in the cell and the transmission speed and amplitude vary depending on the mechanical properties of the medium.242 When encountering a solid/gas interface, most transmitted sound waves are reflected due to high acoustic impedance mismatch. The acoustic signal transmission at such interfaces is poor. Accordingly, minor contact loss at the interface induces sound wave amplitude attenuation.243 In contrast to conventional liquid cells, the Li plating/stripping behavior enhances the roughness of Li metal anode/SSE interfaces. The intensity of sound wave transmission is reduced due to void formation at the interface, causing an amplitude attenuation. Furthermore, low-modulus defects such as cracks or filaments at interfaces also cause significant reduction of sound wave transmission intensity. Therefore, in situ/operando acoustic characterization is a potential technique to non-destructively reveal microstructural variations in LLZO-based SSLBs.

Fig. 27c illustrates a configuration utilized for in situ acoustic transmission.244 The LLZO sample is placed between two Li metal electrodes and a transducer is positioned on one side of the LLZO sample to ensure that only LLZO is affected. Variations in elastic properties and microstructures of LLZO are investigated using sound waves generated from a transducer. The results show that wave speed (LLZO stiffness) is reduced within a few minutes before failure, indicating occurrence of fracture in LLZO. Moreover, the lowering in the rate of LLZO stiffness is proportional to the applied current density. Consequently, in situ/operando acoustic characterization exhibits an ability to detect impending failure in LLZO-based SSLBs. Based on such characteristics, it can be adapted for on-line battery management systems to locate and isolate failed or failing cells. As shown in Fig. 27d, cell testing under different stacking pressures is realized by further modifying the configuration of operando acoustic transmission.179 The stacking pressure is precisely controlled utilizing a double-piston pneumatic cylinder. At 2, 7.4 or 13 MPa stacking pressure, remarkable attenuation of acoustic amplitude occurred and this was attributed to microstructural changes (Fig. 27e). To be specific, amplitude attenuation at low stacking pressures is predominantly attributed to void formation, while that at high stacking pressure is attributed to fracture crack development. In contrast, a stable increase in acoustic amplitude prior to attenuation was attributed to improved contact and reduced cell thickness as Li foil plastically deformed over time. Given the intrinsic degradation mode at high stacking pressures and long duration of cycling, the cells operating at low stacking pressure are more favoured. Notably, in situ/operando acoustic characterization requires a non-vacuum and low-noise environment. Moreover, this approach is currently unavailable for 3D imaging of the interface structure.

3.7.4 In situ/operando chemo-mechanical measurement. LLZO-based SSLBs, composed of dense laminated solid stacks, exhibit intricate chemo-mechanical characteristics at the electrode, electrolyte, and interface surfaces. Due to the lack of buffering effects from liquid electrolyte, even minor volume changes can induce local contact loss or fracture. Therefore, developing a local and non-invasive in situ technique that can detect chemo-mechanical evolution of the cell during cycling is imperative. Fiber Bragg grating (FBG) sensors possess a function of correlating optical signals to variations in temperature, hydraulic pressure and strain.245 FBG shows a periodic modulation of the refractive index along the core of fibers and thus the refractive index of the fiber changes with temperature, hydraulic pressure and strain variations, resulting in a Bragg wavelength shift or/and signal splitting due to induced birefringence.246 Owing to the small size, low costs and multiplexing capability, FBG has been integrated into internal or external coin cells, cylindrical cells, pouch cells, or Swagelok-type cells for real-time monitoring.247,248Fig. 28a illustrates a schematic diagram of the experimental setup for surface strain monitoring in Li-free anode liquid pouch cells according to the working principle of FBG.247 The optical interrogator acts as a light source and data acquisition system. The strain is monitored using FBG1 fixed on the cell surface with adhesive tape. The ambient temperature is monitored using FBG2, which is freely placed on the side of cell. The strain evolution during cycling is in situ monitored utilizing this testing system. As shown in Fig. 28b, regular increases and decreases in strain are consistent with the volume changes caused by lithiation/delithiation. Increased strain during charging is attributed to the volume expansion in the anode over the volume contraction in the cathode while the situation is reversed during discharge. The initial strain is progressively enhanced before each cycle due to dead Li and SEI film formation. Importantly, the decrease in strain fluctuation magnitude induced by loss of active Li is a major indicator of cell failure.
image file: d3mh00135k-f28.tif
Fig. 28 (a) The schematic diagram of an experimental setup for strain monitoring. (b) The surface strain evolution of the anode-free pouch cell during cycling. Reproduced from ref. 247 with permission from Wiley. (c) Schematic illustration of the half-cell used for in situ curvature measurements. (d) Schematic illustration showing how the curvature evolution is measured during plating. Reproduced from ref. 251 with permission from Wiley. (e) Li, La, Zr, O and Ta atomic distribution curves derived from APT tests. (f) Site-specific sample preparation of the APT needle-shaped specimens. Reproduced from ref. 253 with permission from Elsevier.

Accordingly, this technique is available for use in battery management systems to understand the cell's internal health and electrochemical changes. In addition, an FBG integrated inside a Swagelok-type cell was employed to investigate Li–In |LPS| LTO or Li–In |LPS| Li–In cells.248 The viability of accessing orientation anisotropy in Li-driven stress field was proven by birefringence phenomenon. Interestingly, FBG failed to observe responses to stress changes at the electrode level, while it could successfully track stress changes at the electrode/electrolyte interface level. Such phenomenon indicates that internal stress monitoring is preferable for solid-state battery systems rather than external stress monitoring which is generally adopted for liquid batteries. Subsequently, the strain changes in CE-based Li symmetric cells have been monitored in real-time by using an internal stress monitoring system.249 The strain changes are attributed to Li dendrite generation at the interface. The evolution of side-view strain distributions during charging in LLZO-based SSLBs was measured in situ using a digital image correlation technique (DIC).250 The overall strain remarkably increased with SOC or current density increases. Based on principal strain analysis, the Li+ migration pathway was seen to be random and affected by mechanical or current density inhomogeneities.

Recently, a novel experimental configuration (Fig. 28c and d) was designed to in situ monitor the displacement induced by mechanical stresses through the measurement of wafer curvature.251 This technique is capable of investigating the chemo-mechanical evolution during electrochemical cycling. A thin Al2O3 layer is plated between LLZO and melted Li utilizing atomic layer deposition to improve interface contact. A CCD detector is employed to detect a set of parallel laser beams reflected off the current collector. The in situ curvature value is derived from spacing variation of the parallel laser beam. During plating, the stress (displacement) gradually increased and reached a steady state over several hours. Moreover, cell short circuit was directly correlated with well-defined pressure variations. Results indicate that failure originated from inherent defects in LLZO and thus minimizing the inherent defects during manufacturing is essential. However, there are very few in situ analyses of the stress evolution mechanisms related to the field of solid-state batteries and thus research into the chemo-mechanical evolution of cells is expected to attract more attention in the future.

3.7.5 Atom probe tomography (APT). APT provides 3D imaging ability, near-atomic-level resolution, and features equal probability to detect light and heavy elements.252 Since experimental manipulation is generally conducted at low temperatures (20–50 K), APT enables the characterization of Li distribution. Importantly, by combining APT and other microscopy techniques, local features of the region of interest can be investigated. APT and SEM have been used to examine the distribution and concentration of Li at LLZO grain boundaries at sub-nanometer levels.253Fig. 28f shows site-specific sample preparation of APT needle-shaped specimens. It includes four main steps: selecting the research region, preparing a chunk-shaped lamella containing the research region, removing the lamella using a micromanipulator, and mounting the lamella on top of a Si support tip. As shown in Fig. 28e, Li was observed to be intensely accumulated at the grain boundaries forming Li-based phases denoted as complexions with Li compositions exhibiting considerable variation (34.6–82.5 at%). The presence of Li-rich complexions results in low ionic conductivity and embrittlement at the grain boundaries. Adequate amounts of dopants or improved processing routes (such as introduction of sintering aids) can assist in suppressing Li-rich complexion formation at grain boundaries.

In summary, each in situ/operando technique is beneficial for researching the scientific and technical challenges faced by LLZO-based SSLBs at various temporal and spatial resolutions. However, each technique exhibits strengths and limitations and thus combining different testing methods together will be critical to obtain a comprehensive picture of the phenomena occurring at different sections of the battery system.

4. Summary and outlook

LLZO-based SSLBs with both high energy density and safety are considered as promising candidates for next-generation energy storage devices. However, their application and development are impacted by several challenges, including poor ionic conductivity, electrode/electrolyte interface incompatibility, electrode or electrolyte material degradation, and ambiguous reaction mechanisms in cells. Further progress to overcome these challenges is contingent on greater comprehension and insight into ion/electron transport, interface reaction mechanism, and lithiation/delithiation kinetic processes. Owing to the limitations of conventional characterization techniques in requiring cell disassembly and testing only pre- or post-cycling, in situ/operando characterization techniques are critically required to investigate the underlying mechanisms responsible for performance and issues causing failure. This review provides a comprehensive summary of important in situ/operando characterization techniques that have been used for the study of LLZO-based SSLBs. These techniques operate on analysing regions on various spatial scales ranging from Å to mm, temporal scales from ns to h, and environmental scales from ex situ to in situ/operando. The characteristics of each technique have been discussed in terms of working principles, experimental setup, operational conditions, cell structure, advantages, limitations, and applications.

Table 2 summarises the main purpose, advantages, limitations, and issues related to each technique. To summarise, the in situ/operando characterization techniques can help to gain insight into the principles and mechanisms of the phenomena occurring in the LLZO-based SSLBs and provide the essential theoretical foundation and experimental strategies that can be adopted to determine the issues that are impeding the achievement of high energy density, long cycle life, and high safety and commercialization of these units.

Table 2 Summary of in situ/operando characterization techniques and applications
Techniques Main purpose Advantages Limitations Addressed major issues
Optical based techniques
OM Morphology evolution Simple operation; minor damage; non-vacuum condition Low resolution; invisible to opaque materials Li dendrite growth mechanism; interlayer modification mechanism; interface reaction mechanism
Raman spectroscopy Chemical structure analysis Non-destructive; sensitive to non-polar bonds Unable to detect metal signal Thin film preparation; interlayer modification mechanism; electrochemical reaction mechanism
FTIR spectroscopy Chemical structure analysis Non-destructive; monitoring organic species Difficulty in measuring inorganic compounds Not yet applied
Electron based techniques
SEM Morphology evolution; structure analysis Larger depth of field; continuously adjustable magnification High vacuum conditions Li dendrite growth mechanism; interlayer modification mechanism
TEM Morphology evolution; structure analysis Ultra-high spatial resolution High vacuum conditions; extremely thin samples Li dendrite growth mechanism; interface reaction mechanism; material structure evolution
X-Ray based techniques
XRD Phase and crystal structure evolution Fast; easy to operate; non-destructive; non-contamination Not applicable for amorphous or poorly crystalline samples Material phase conversion mechanism
X-Ray imaging 2D/3D structure imaging Non-destructive visualization; high spatial/temporal resolution Affected by contrast, noise artifacts Porous structure preparation; component distribution; Li dendrite growth mechanism; thermal behavior
XPS Elemental analysis Penetration depth is a few nanometers; precise in monitoring the chemical species High vacuum conditions; unable to provide overall composition Interface/surface composition and reaction mechanism; interlayer modification mechanism
XAS Elemental analysis Penetration depth is a few nanometers; precise in monitoring the chemical species Unable to provide overall composition Interface/surface composition and reaction mechanism
Scanning probe based techniques
AFM Morphology evolution; mechanical property; electrical property Atomic resolution; accurate height information; high-resolution topographical images Slow imaging speed; limited magnification range; limited vertical range Interface reaction mechanism; Li dendrite growth mechanism; interlayer modification mechanism; Li+ migration pathway
SECM Morphology evolution; compositional analysis Electrical or electrochemical information with nanoscale resolution Not possible to maintain a constant tip-to-substrate separation Not yet applied
Magnetism based techniques
NMR Elemental distribution; 3D structure imaging Non-destructive; quantitative research; high sensitivity Skin effect; low spatial/temporal resolution Ionic transport characteristics; Li+ migration pathway; Li dendrite growth mechanism
EPR Microstructural formation High sensitivity; high detection ability Complex technique; semi-quantitative research Surface states
Neutron based techniques
NDP Li concentration distribution; active site changes High detection sensitivity to light elements; high spatial resolution; quantitative analysis Under vacuum or pressure condition; high price Li dendrite growth mechanism; interlayer modification mechanism; interface reaction mechanism
NI 2D/3D Li dynamic distribution High penetration; light element sensitivity Slow imaging speed; high price Li dendrite growth mechanism
Other techniques
ToF-SIMS Structure and composition analysis Multi-element analysis; secondary ion image; high ion availability Secondary product contamination; in situ mode under electrochemical response is not available Interlayer modification mechanism; interface reaction mechanism
EIS Electrochemical properties Non-destructive; simple testing setup; low cost Empirical interpretation of EIS Interlayer modification mechanism; interface reaction mechanism
Acoustic characterization Structure analysis Non-destructive; simple testing setup; non-vacuum condition Use of low-noise condition Li dendrite growth mechanism
Chemo-mechanical measurement Chemo-mechanical evolution Non-invasive; low costs; strain detection Non-quantitative research Cell failure prediction; mechanical degradation mechanism


However, some fundamental challenges remain unsolved, including understanding of Li metal-related processes in terms of reaction kinetics, mass transfer, and mechanics. Stable Li plating/stripping in practical solid-state batteries (<100 μm thickness) requires an in-depth understanding of the underlying mechanisms. The interplay of mechanics and transport impacts the degradation mechanism. Similarly, the nature of the electric field distribution in LLZO-based SSEs is an issue that needs to be addressed, which significantly affects the Li filament formation mechanism. Another critical challenge is to optimize composite cathodes in terms of cathode interior stability, reaction kinetics, and three-phase boundary transport. Furthermore, the end-to-end battery production steps such as synthesis, processing and integration need to be studied in a scalable fashion. Therefore, more in situ/operando characterization techniques are still needed to satisfy the requirements for further practical applications of LLZO-based SSLBs with high energy density and high safety. Moreover, each characterization technique exhibits a specific range of precision, depth, detail, and issues. This entails further development of in situ/operando characterization techniques and thus the future perspectives for these in situ technologies are as follows.

(i) More accurate, faster and smarter. Large-scale in situ imaging techniques such as OM fail to provide detailed information under high magnifications, while small-scale imaging techniques such as TEM can provide this missing information. However, small-scale techniques generally examine only very narrow regions, leading to highly subjective analysis results. In situ quantitative spectroscopic techniques can be used to perform initial measurements on the sample to improve the accuracy of results. Although spectroscopic techniques such as Raman, FTIR, NMR, EPR, EIS cannot visualize electrochemical reactions, they can be used to conduct precise analysis of large-scale samples. Increasing the amount of data collected is a perfect approach to avoid subjectivity and single events. However, the massive amount of data will require more analysis time and thus a smarter and simpler data analysis method is needed.

Recently, artificial intelligence technology has quietly emerged to help data collection and analysis, especially in accessing more precise results and saving analysis time, and has been employed in the medical field.254 However, for use of artificial intelligence in battery in situ/operando characterization techniques, two conditions need to be met—high spatial/temporal resolution techniques and large databases. The data collection speed of most characterization techniques is on the time scale of seconds to minutes, and for a few techniques, such as NI, it can reach several hours, while the speed of electrochemical reactions is generally on the time scale of milliseconds to microseconds. Therefore, it is impossible for the speed of data acquisition to match the speed of electrochemical reactions. This requires increasing the strength of the signal source such as the SR source to increase data collection speed to achieve high time resolution, which is also a solution to improve spatial resolution. Furthermore, several constraints such as the need for vacuum and a sample chamber make the cell test very different from the actual situation.

In situ cell design should further consider the real electrode component environment and pressure similar to pouch cells or coin cells. Several in situ/operando characterization techniques such as TEM suffer from low spatial resolution due to the way the experiment is setup. The accuracy of the experimental setup can be enhanced by improving vacuum, cell windows, and use of new materials such as electrostrictive strain materials.255 Therefore, further enhancements in spatial, temporal, and environmental resolution are crucial for data collection and analysis and a comprehensive understanding of the problem.

(ii) Closer to the practical test conditions. Most of the advanced in situ characterization techniques are based on scientific models. Although it is effective for monitoring in situ dynamic evolution, the in situ cell system is relatively simplified and idealized, and thus it cannot completely reveal the internal processes of a real working battery. Even when using the operando characterization techniques, several phenomena have been observed which are different from those seen in a practical cell. In situ SEM/TEM technique has been typically used with a micro-protocell at extremely low current densities and under high vacuum conditions. This is clearly different from the cell's real operating conditions. High electron doses could also damage the cell system (such as Li metal) causing biased test results. In situ OM/Raman/AFM technology enables utilization of an in situ cell with a similar high surrounding voltage as a normal cell. However, the current density applied to this cell is still relatively low. In situ magnetic- and neutron-based techniques generally suffer from long test times and thus the in situ cell must be kept in one state for longer times to ensure proper running of the test. Moreover, such in situ cells would also need a high stack pressure. Notably, operando SR X-ray- and EIS-based technology, and chemo-mechanical measurements are available for testing practical operating cells (e.g., coin cells, three-electrode cells or pouch cells, etc.). In particular, operando EIS technique can be applied for a larger-scale work in the laboratory. Therefore, based on the disadvantages of the test conditions of each in situ/operando technique, it is important to further develop advanced in situ/operando characterization setups to realize multi-scale real-time investigations of electrochemical reactions in SSLBs under practical operating conditions.

(iii) Sample friendly. Sample friendly in situ/operando characterization techniques can provide improved characterization of materials. In terms of sample preparation, complex pre-treatment may be involved, such as FIB milling technique for in situ EM characterization of tiny or cross-sectional samples. This pre-treatment can damage or contaminate the sample. Hence, it is desirable to directly use original samples without pre-treatment. In terms of sample transfer, air-sensitive materials such as LLZO and Li metal require a special inert gas-protected transfer device. Thus the development of a centrally integrated ultra-high vacuum (UHV) cluster tool facility would pave the way for conducting analysis without external contamination since it simultaneously accommodates and interconnects multiple in situ/operando characterization techniques such as XPS, Raman, SEM, TEM, and ToF-SIMS. In terms of characterization, Li or polymers not been encapsulated in the cell become unstable when subjected to high-energy incident sources such as high-energy electron beam or X-rays and this can destroy the interface-modified layer similar to SEI films and influence the electrochemical reactions. To overcome this, use of cryogenic temperature can prevent strong incident signal sources from damaging samples, but further development in this aspect is required.

(iv) New technology development. At present, there are numerous reports on the thermal runaway, thermal expansion, and safety of conventional liquid batteries, but few studies on the thermal behavior and safety of LLZO-based SSLBs. The main causes of low battery safety are mechanical, electrical and thermal abuse. The work carried out by coupling of a calorimeter and operando SR X-ray fast radiography mentioned in this review suggested that LLZO-based SSLBs may be less safe than liquid batteries. Consequently, in situ/operando characterization techniques and simulation software are essential to investigate the thermal behavior of batteries in future research. For example, the thermal runaway behavior of batteries can be predicted by in situ characterization techniques that monitor changes in micro-morphology (sudden expansion) and temperature (sudden rise) or voltage (sudden drop) in real-time. Mastering the thermal runaway path and critical parameter will help improve battery safety. In situ characterization techniques for real-time and accurate monitoring of volume expansion-induced stress evolution during the cell cycle should be further explored. In addition, it would be worthwhile to develop in situ/operando modes for several ex situ techniques such as NMR, MRI, EPR. Overall, the development of a new characterization technique will help gain insights into cell kinetics/thermodynamic mechanisms and uncover complex reaction mechanisms.

(v) Multi-modal and computational simulations. Most in situ/operando characterization techniques are employed independently, and each technique has a best-fit situation and its own limitations. Importantly, the chemical, mechanical and electrochemical transformations of materials in battery systems are strongly interdependent. Accordingly, no technique can adequately characterize and probe the fundamental principle/mechanism of the issues involved. Combining various characterization techniques, namely, multi-modal characterization, is essential to acquire more comprehensive information on interaction, morphology, and chemical composition. For example, a combination of AFM and TEM enables in situ imaging and stress measurements. Recently, first-principles calculation methods based on density functional theory (DFT) and molecular dynamics methods (MD) have attracted much attention. Computational simulations can predict correlated properties of material systems, including phase stability, Li diffusion behavior, interface formation energy and electrode/electrolyte interface reactions. It can reduce research time and cost. Coupling in situ/operando characterization techniques with computational simulations can provide insight into reaction mechanisms. The acquired characterization data can also be used to verify the accuracy of computational simulations, which can guide establishment of new models and development of new methods. In addition, the high-throughput computation of material genome enables simultaneous prediction and screening of massive data related to samples, considerably reducing the experimental cost and improving efficiency. Therefore, computational simulation and in situ/operando characterization techniques complement each other, laying the foundation for the practical application of LLZO-based SSLBs.

List of acronyms and abbreviations

2DTwo-dimensional
3DThree-dimensional
ABFAnnular bright field
ACAlternating current
AFMAtomic force microscopy
ALDAtomic layer deposition
Al-LLZOAl-doped Li7La3Zr2O12
APTAtom probe tomography
BTOBaTiO3
CcChromatic aberration
CCDCharge-coupled device
CEsComposite electrolytes
CNTsCarbon nanotubes
CSIChemical shift imaging
CsSpherical aberration
CTComputed tomography
CVDChemical vapor deposition
c-AFMConductive atomic force microscopy
c-LLZOCubic phase Li7La3Zr2O12
DCDirect current
DFTDensity functional theory
DICDigital image correlation
DMDouble modulation
DPCDifferential phase contrast
DRIFTSDiffuse reflectance infrared transform spectroscopy
DRTDistribution of relaxation time
EBSDElectron backscatter diffraction
EC-AFMElectrochemical atomic force microscopy
EDSEnergy dispersive spectroscopy
EDXEnergy dispersive X-ray spectroscopy
EELSElectron energy loss spectroscopy
EHElectron holography
EISElectrochemical impedance spectroscopy
EMElectron microscopy
EPRElectron paramagnetic resonance
ESEMEnvironmental scanning electron microscopy
ETEMEnvironmental transmission electron microscopy
EXAFSExtended X-ray absorption fine-structure
EXSYExchange spectroscopy
FBGFiber Bragg grating
FECFluoroethylene carbonate
FF-HEDMFar-field high-energy diffraction microscopy
FIBFocused ion beam
FTIRFlourier transform infrared
FWHMFull width at half maximum
G4Tetraethylene glycol dimethyl ether
GFGlass fiber
GIGrazing incidence
GISXRDGrazing incidence synchrotron X-ray diffraction
Ga-LLZOGa-doped Li7La3Zr2O12
g-C3N4Graphitic carbon nitride
HAADFHigh angle annular dark field
h-BNHexagonal boron nitride
INSInelastic neutron scattering
iSCATInterferometric scattering microscopy
ITOIndium tin oxide
KPFMKelvin probe force microscopy
LAGPLi1.5Al0.5Ge1.5(PO4)3
LBOLi3BO3
LCOLiCoO2
LFPLiFePO4
LGPSLi10GeP2S12
LIBsLi-ion batteries
LiBFSIELi-conductive polymer framework
LICGCLi2Al0.5Ti0.75Ge0.75Si0.5P2.5O12
LiFSILi bis(fluorosulfonyl) imide
LiPAALi-inserted polyacrylic acid
LiTFSILi bis(trifluoromethylsulfonyl) imide
LLCZWOCa/W dual-doped Li7La3Zr2O12
LLZOLi7La3Zr2O12
LLZTOTa-doped Li7La3Zr2O12
LMOLiMn2O4
LNMOLiNi0.5Mn1.5O4
LPSLi3PS4
LPSClLi6PS5Cl
LSCMLaser scanning confocal microscopy
LZOLa2Zr2O7
MASMagic angle spinning
MDMolecular dynamics
MEMSMicro-electro-mechanical system
MRIMagnetic resonance imaging
NCMLiNixCoyMn1−xyO2
NCM111LiNi1/3Co1/3Mn1/3O2
NCM523LiNi0.5Co0.2Mn0.3O2
NCM622LiNi0.6Co0.2Mn0.2O2
NCM811LiNi0.8Co0.1Mn0.1O2
NDPNeutron depth profiling
NINeutron imaging
NMRNuclear magnetic resonance
NPDNeutron powder diffraction
NRNeutron reflectometry
NRINeutron radiography imaging
NTINeutron tomography imaging
NTSNeutron total scattering
OCVOpen circuit voltage
ODIOxygen-deficient interfacial layer
OMOptical microscopy
OPSOrganic plastic salt
PANPolyacrylonitrile
PCPropylene carbonate
PCL-PTMCPoly(ε-caprolactone-co-trimethylene carbonate)
PEEKPolyether ether ketone
PEOPoly(ethylene oxide)
PETGPolyethylene terephthalate glycol
PFGPulsed field gradient
PRPressure recording
PSsPolysulfides
PTFEPolytetrafluoroethylene
PVDFPolyvinylidene fluoride
PVDF-HFPPolyvinylidene fluoride-hexafluoropropylene copolymer
QENSQuasielastic neutron scattering
RFRadio frequency
RFSReactive flash sintering
SAMsSelf-assembled monolayers
SANSSmall-angle neutron scattering
SASESelf-adaptive structure electrolyte
SBFSEMSerial block-face scanning electron microscopy
SCLSpace charge layer
SECMScanning electrochemical microscopy
SEISolid electrolyte interphase
SEMScanning electron microscopy
SNSuccinonitrile
SNISubtractive normalized interfacial
SOCState of charge
SOHState of health
SPMScanning probe microscopy
SRSynchrotron radiation
SRSStimulated Raman scattering
SSEsSolid-state electrolytes
SSLBsSolid-state Li batteries
SSNMRSolid state nuclear magnetic resonance
STEMScanning transmission electron microscopy
STMScanning tunneling microscopy
T 1 Spin-lattice relaxation
T 1ρ Spin-lattice relaxation in the rotating frame
T 2 Spin-spin relaxation
TBA tert-Butyl alcohol
TCRThermo couples recording
TEGDMETetraethylene glycol dimethyl ether
TEMTransmission electron microscopy
TEYTotal electron yield
TFYTotal fluorescence yield
ToF-SIMSTime-of-flight secondary ion mass spectroscopy
TXMTransmission X-ray microscopy
t-LLZOTetragonal phase Li7La3Zr2O12
Ta-LLZOTa-doped Li7La3Zr2O12
UHVUltra-high vacuum
XANESX-Ray absorption near-edge structure
XASX-Ray absorption spectroscopy
XPSX-Ray photoelectron spectroscopy
XRDX-Ray diffraction
XRIX-Ray imaging
XRTX-Ray tomography

Author contributions

Lei Zhang: conceptualization, methodology, nvestigation, formal analysis, writing – original draft, and writing – review and editing. Huilin Fan: writing – original draft, and writing – review and editing. Yuzhen Dang: formal analysis, writing – original draft, and writing – review and editing. Quanchao Zhuang: methodology, supervision, writing – original draft, and writing – review and editing. Hamidreza Arandiyan: conceptualization, investigation, methodology, supervision, writing – original draft, and writing – review and editing. Yuan Wang: conceptualization, methodology, writing – original draft, and writing – review and editing. Ningyan Cheng: writing – original draft and writing – review and editing. Hongyu Sun: conceptualization, investigation, methodology, supervision, writing – original draft, and writing – review and editing. H Hugo Pérez Garza: writing – original draft and writing – review and editing. Runguo Zheng: writing – original draft and writing – review and editing. Zhiyuan Wang: formal analysis, writing – original draft, and writing – review and editing. Sajjad S. Mofarah: writing – review and editing. Pramod Koshy: writing – review and editing. Suresh K. Bhargava: formal analysis, writing – original draft, and writing – review and editing. Yanhua Cui: formal analysis, writing – original draft, and writing – review and editing. Zongping Shao: conceptualization, methodology, writing – original draft, and writing – review and editing. Yanguo Liu: conceptualization, investigation, methodology, supervision, writing – original draft, and writing – review and editing.

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

This study was financially supported by National Natural Science Foundation of China (52177208, 52171202, 51971055, and 51871046) and National Safety Academic Fund (U1730136). H. Arandiyan and S. Bhargava acknowledge financial support from Queensland Pacific Metals (QPM).

References

  1. B. Scrosati and J. Garche, J. Power Sources, 2010, 195, 2419–2430 CrossRef CAS .
  2. H. C. Hesse, M. Schimpe, D. Kucevic and A. Jossen, Energies, 2017, 10, 2107 CrossRef .
  3. K. Xu, Chem. Rev., 2004, 104, 4303–4417 CrossRef CAS PubMed .
  4. X. B. Cheng, R. Zhang, C. Z. Zhao and Q. Zhang, Chem. Rev., 2017, 117, 10403–10473 CrossRef CAS PubMed .
  5. M. B. Dixit, J. S. Park, P. Kenesei, J. Almer and K. B. Hatzell, Energy Environ. Sci., 2021, 14, 4672–4711 RSC .
  6. V. Thangadurai, S. Narayanan and D. Pinzaru, Chem. Soc. Rev., 2014, 43, 4714–4727 RSC .
  7. R. S. Chen, A. M. Nolan, J. Z. Lu, J. Y. Wang, X. Q. Yu, Y. F. Mo, L. Q. Chen, X. J. Huang and H. Li, Joule, 2020, 4, 812–821 CrossRef CAS .
  8. Y. Liu, H. Zhang, N. Jiang, W. Zhang, H. Arandiyan, Z. Wang, S. Luo, F. Fang and H. Sun, J. Alloys Compd., 2020, 834, 155030 CrossRef CAS .
  9. L. Zhang, Q. C. Zhuang, R. G. Zheng, Z. Y. Wang, H. Y. Sun, H. Arandiyan, Y. Wang, Y. G. Liu and Z. P. Shao, Energy Storage Mater., 2022, 49, 299–338 CrossRef .
  10. H. C. Yang, P. Tang, N. Piao, J. Li, X. Y. Shan, K. P. Tai, J. Tan, H. M. Cheng and F. Li, Mater. Today, 2022, 57, 279–294 CrossRef CAS .
  11. S. Narayanan, J. S. Gibson, J. Aspinall, R. S. Weatherup and M. Pasta, Curr. Opin. Solid State Mater. Sci., 2022, 26, 100978 CrossRef CAS .
  12. N. Zhao, W. Khokhar, Z. J. Bi, C. Shi, X. X. Guo, L. Z. Fan and C. W. Nan, Joule, 2019, 3, 1190–1199 CrossRef CAS .
  13. R. Pfenninger, M. Struzik, I. Garbayo, E. Stilp and J. L. M. Rupp, Nat. Energy, 2019, 4, 475–483 CrossRef CAS .
  14. H. C. Wang, H. W. Gao, X. X. Chen, J. P. Zhu, W. Q. Li, Z. L. Gong, Y. X. Li, M. S. Wang and Y. Yang, Adv. Energy Mater., 2021, 11, 2102148 CrossRef CAS .
  15. A. Sharafi, S. H. Yu, M. Naguib, M. Lee, C. Ma, H. M. Meyer, J. Nanda, M. F. Chi, D. J. Siegel and J. Sakamoto, J. Mater. Chem. A, 2017, 5, 13475–13487 RSC .
  16. Y. Zhou, F. Chen, H. Arandiyan, P. Guan, Y. Liu, Y. Wang, C. Zhao, D. Wang and D. Chu, J. Energy Chem., 2021, 57, 516–542 CrossRef CAS .
  17. P. Barai, T. Rojas, B. Narayanan, A. T. Ngo, L. A. Curtiss and V. Srinivasan, Chem. Mater., 2021, 33, 5527–5541 CrossRef CAS .
  18. Y. Gong, Y. Y. Chen, Q. H. Zhang, F. Q. Meng, J. A. Shi, X. Y. Liu, X. Z. Liu, J. N. Zhang, H. Wang, J. Y. Wang, Q. Yu, Z. Zhang, Q. Xu, R. J. Xiao, Y. S. Hu, L. Gu, H. Li, X. J. Huang and L. Q. Chen, Nat. Commun., 2018, 9, 3341 CrossRef PubMed .
  19. Y. Nomura, K. Yamamoto, M. Fujii, T. Hirayama, E. Igaki and K. Saitoh, Nat. Commun., 2020, 11, 2824 CrossRef CAS PubMed .
  20. W. B. Zhang, T. Leichtweiss, S. P. Culver, R. Koerver, D. Das, D. A. Weber, W. G. Zeier and J. Janek, ACS Appl. Mater. Interfaces, 2017, 9, 35888–35896 CrossRef CAS PubMed .
  21. H. Okamoto, J. Phase Equilib. Diffus., 2009, 30, 118–119 CrossRef CAS .
  22. E. Kazyak, R. Garcia-Mendez, W. S. LePage, A. Sharafi, A. L. Davis, A. J. Sanchez, K. H. Chen, C. Haslam, J. Sakamoto and N. P. Dasgupta, Matter, 2020, 2, 1025–1048 CrossRef .
  23. J. Kasemchainan, S. Zekoll, D. S. Jolly, Z. Y. Ning, G. O. Hartley, J. Marrow and P. G. Bruce, Nat. Mater., 2019, 18, 1105–1111 CrossRef CAS PubMed .
  24. M. Fingerle, R. Buchheit, S. Sicolo, K. Albe and R. Hausbrand, Chem. Mater., 2017, 29, 7675–7685 CrossRef CAS .
  25. S. Randau, D. A. Weber, O. Kotz, R. Koerver, P. Braun, A. Weber, E. Ivers-Tiffee, T. Adermann, J. Kulisch, W. G. Zeier, F. H. Richter and J. Janek, Nat. Energy, 2020, 5, 259–270 CrossRef CAS .
  26. C. Brissot, M. Rosso, J. N. Chazalviel, P. Baudry and S. Lascaud, Electrochim. Acta, 1998, 43, 1569–1574 CrossRef CAS .
  27. K. N. Wood, E. Kazyak, A. F. Chadwick, K. H. Chen, J. G. Zhang, K. Thornton and N. P. Dasgupta, ACS Cent. Sci., 2016, 2, 790–801 CrossRef CAS PubMed .
  28. W. C. Guo, F. Shen, J. W. Liu, Q. Q. Zhang, H. Guo, Y. T. Yin, J. Gao, Z. T. Sun, X. G. Han and Y. S. Hu, Energy Storage Mater., 2021, 41, 791–797 CrossRef .
  29. A. A. Wang, J. Li, M. Y. Yi, Y. Y. Xie, S. L. Chang, H. B. Shi, L. Y. Zhang, M. H. Bai, Y. E. Zhou, Y. Q. Lai and Z. Zhang, Energy Storage Mater., 2022, 49, 246–254 CrossRef .
  30. A. Sharafi, H. M. Meyer, J. Nanda, J. Wolfenstine and J. Sakamoto, J. Power Sources, 2016, 302, 135–139 CrossRef CAS .
  31. B. Kinzer, A. L. Davis, T. Krauskopf, H. Hartmann, W. S. LePage, E. Kazyak, J. Janek, N. P. Dasgupta and J. Sakamoto, Matter, 2021, 4, 1947–1961 CrossRef CAS .
  32. W. Manalastas, J. Rikarte, R. J. Chater, R. Brugge, A. Aguadero, L. Buannic, A. Llordes, F. Aguesse and J. Kilner, J. Power Sources, 2019, 412, 287–293 CrossRef CAS .
  33. C. Monroe and J. Newman, J. Electrochem. Soc., 2005, 152, A396–A404 CrossRef CAS .
  34. L. Porz, T. Swamy, B. W. Sheldon, D. Rettenwander, T. Fromling, H. L. Thaman, S. Berendts, R. Uecker, W. C. Carter and Y. M. Chiang, Adv. Energy Mater., 2017, 7, 1701003 CrossRef .
  35. S. Kim, C. Jung, H. Kim, K. E. Thomas-Alyea, G. Yoon, B. Kim, M. E. Badding, Z. Song, J. Chang, J. Kim, D. Im and K. Kang, Adv. Energy Mater., 2020, 10, 1903993 CrossRef CAS .
  36. J. M. Xu, C. Ma, C. Y. Chang, X. F. Lei, Y. W. Fu, J. Wang, X. Z. Liu and Y. Ding, ACS Appl. Mater. Interfaces, 2021, 13, 38179–38187 CrossRef CAS PubMed .
  37. M. X. Zhang, Z. P. Qi, S. Cai, G. J. Li, Q. B. Deng, S. Liu, L. Wang, L. Dai and N. Hu, ACS Appl. Energy Mater., 2022, 5, 862–869 CrossRef CAS .
  38. Y. X. Song, Y. Shi, J. Wan, S. Y. Lang, X. C. Hu, H. J. Yan, B. Liu, Y. G. Guo, R. Wen and L. J. Wan, Energy Environ. Sci., 2019, 12, 2496–2506 RSC .
  39. Y. X. Song, Y. Shi, J. Wan, B. Liu, L. J. Wan and R. Wen, Adv. Energy Mater., 2020, 10, 2000465 CrossRef CAS .
  40. Y. X. Song, J. Wan, H. J. Guo, Y. Shi, X. C. Hu, B. Liu, H. J. Yan, R. Wen and L. J. Wan, Energy Storage Mater., 2021, 41, 642–649 CrossRef .
  41. A. J. Merryweather, C. Schnedermann, Q. Jacquet, C. P. Grey and A. K. Rao, Nature, 2021, 594, 522–528 CrossRef CAS PubMed .
  42. J. Ortega-Arroyo and P. Kukura, Phys. Chem. Chem. Phys., 2012, 14, 15625–15636 RSC .
  43. H. K. Tian, A. Chakraborty, A. A. Talin, P. Eisenlohr and Y. Qi, J. Electrochem. Soc., 2020, 167, 090541 CrossRef CAS .
  44. H. Li, D. L. Chao, B. Chen, X. Chen, C. Chuah, Y. H. Tang, Y. Jiao, M. Jaroniec and S. Z. Qiao, J. Am. Chem. Soc., 2020, 142, 2012–2022 CrossRef CAS PubMed .
  45. Z. Cheng, H. Pan, F. Li, C. Duan, H. Liu, H. Y. Zhong, C. C. Sheng, G. J. Hou, P. He and H. S. Zhou, Nat. Commun., 2022, 13, 125 CrossRef CAS PubMed .
  46. M. Otoyama, H. Kowada, A. Sakuda, M. Tatsumisago and A. Hayashi, J. Phys. Chem. Lett., 2020, 11, 900–904 CrossRef CAS PubMed .
  47. Y. Matsuda, N. Kuwata, T. Okawa, A. Dorai, O. Kamishima and J. Kawamura, Solid State Ionics, 2019, 335, 7–14 CrossRef CAS .
  48. S. Rajendran, N. K. Thangavel, K. Mahankali and L. M. R. Arava, ACS Appl. Energy Mater., 2020, 3, 6775–6784 CrossRef CAS .
  49. M. Otoyama, Y. Ito, A. Sakuda, M. Tatsumisago and A. Hayashi, Phys. Chem. Chem. Phys., 2020, 22, 13271–13276 RSC .
  50. J. H. Hu, Z. T. Sun, Y. R. Gao, P. Li, Y. F. Wu, S. W. Chen, R. B. Wang, N. A. Li, W. E. Yang, Y. X. Shen and S. H. Bo, Cell Rep. Phys. Sci., 2022, 3, 100938 CrossRef CAS .
  51. A. Sharafi, S. Yu, M. Naguib, M. Lee, C. Ma, H. M. Meyer, J. Nanda, M. Chi, D. J. Siegel and J. Sakamoto, J. Mater. Chem. A, 2017, 5, 13475–13487 RSC .
  52. Z. H. Ren, J. X. Li, Y. Y. Gong, C. Shi, J. N. Liang, Y. L. Li, C. X. He, Q. L. Zhang and X. Z. Ren, Energy Storage Mater., 2022, 51, 130–138 CrossRef .
  53. M. Otoyama, Y. Ito, A. Hayashi and M. Tatsumisago, J. Power Sources, 2016, 302, 419–425 CrossRef CAS .
  54. X. N. Li, J. W. Liang, X. Li, C. H. Wang, J. Luo, R. Y. Li and X. L. Sun, Energy Environ. Sci., 2018, 11, 2828–2832 RSC .
  55. J. Zhang, C. Zheng, L. J. Li, Y. Xia, H. Huang, Y. P. Gan, C. Liang, X. P. He, X. Y. Tao and W. K. Zhang, Adv. Energy Mater., 2020, 10, 1903311 CrossRef CAS .
  56. J. Pu, C. Zhong, J. Liu, Z. Wang and D. Chao, Energy Environ. Sci., 2021, 14, 3872–3911 RSC .
  57. C. W. Freudiger, W. Min, B. G. Saar, S. Lu, G. R. Holtom, C. W. He, J. C. Tsai, J. X. Kang and X. S. Xie, Science, 2008, 322, 1857–1861 CrossRef CAS PubMed .
  58. Q. Cheng, L. Wei, Z. Liu, N. Ni, Z. Sang, B. Zhu, W. H. Xu, M. J. Chen, Y. P. Miao, L. Q. Chen, W. Min and Y. Yang, Nat. Commun., 2018, 9, 2942 CrossRef PubMed .
  59. A. M. Tripathi, W. N. Su and B. J. Hwang, Chem. Soc. Rev., 2018, 47, 736–851 RSC .
  60. M. L. McKelvy, T. R. Britt, B. L. Davis, J. K. Gillie, L. A. Lentz, A. Leugers, R. A. Nyquist and C. L. Putzig, Anal. Chem., 1996, 68, R93–R160 CrossRef .
  61. J. X. Wu, M. Ihsan-Ul-Haq, Y. M. Chen and J. K. Kim, Nano Energy, 2021, 89, 106489 CrossRef CAS .
  62. T. Krauskopf, R. Dippel, H. Hartmann, K. Peppler, B. Mogwitz, F. H. Richter, W. G. Zeier and J. Janek, Joule, 2019, 3, 2030–2049 CrossRef CAS .
  63. M. Golozar, A. Paolella, H. Demers, S. Savoie, G. Girard, N. Delaporte, R. Gauvin, A. Guerfi, H. Lorrmann and K. Zaghib, Sci. Rep., 2020, 10, 18410 CrossRef CAS PubMed .
  64. J. Zhao, Y. F. Tang, Q. S. Dai, C. C. Du, Y. Zhang, D. C. Xue, T. W. Chen, J. Z. Chen, B. Wang, J. M. Yao, N. Zhao, Y. S. Li, S. M. Xia, X. X. Guo, S. J. Harris, L. Q. Zhang, S. L. Zhang, T. Zhu and J. Y. Huang, Energy Environ. Mater., 2022, 5, 524–532 CrossRef CAS .
  65. T. Krauskopf, B. Mogwitz, H. Hartmann, D. K. Singh, W. G. Zeier and J. Janek, Adv. Energy Mater., 2020, 10, 2000945 CrossRef CAS .
  66. S. Heo, D. Lee, K. Kim, Y. Kim, D. J. Yun, S. Park, J. Lee, S. Kim, J. S. Kim and S. Park, J. Power Sources, 2021, 510, 230389 CrossRef CAS .
  67. C. Chen, Q. Li, Y. Q. Li, Z. H. Cui, X. X. Guo and H. Li, ACS Appl. Mater. Interfaces, 2018, 10, 2185–2190 CrossRef CAS PubMed .
  68. H. Arandiyan, S. S. Mofarah, C. C. Sorrell, E. Doustkhah, B. Sajjadi, D. Hao, Y. Wang, H. Sun, B.-J. Ni, M. Rezaei, Z. Shao and T. Maschmeyer, Chem. Soc. Rev., 2021, 50, 10116–10211 RSC .
  69. S. J. Tang, G. W. Chen, F. C. Ren, H. C. Wang, W. Yang, C. X. Zheng, Z. L. Gong and Y. Yang, J. Mater. Chem. A, 2021, 9, 3576–3583 RSC .
  70. Y. Huang, B. Chen, J. Duan, F. Yang, T. R. Wang, Z. F. Wang, W. J. Yang, C. C. Hu, W. Luo and Y. H. Huang, Angew. Chem., Int. Ed., 2020, 59, 3699–3704 CrossRef CAS PubMed .
  71. C. Cui, C. Zeng, G. X. Huang, X. Feng, Y. Zhang, T. Y. Zhai and H. Q. Li, Adv. Energy Mater., 2022, 12, 2202250 CrossRef CAS .
  72. E. J. Cheng, K. Kanamura, Y. Kushida and T. Abe, ACS Appl. Mater. Interfaces, 2022, 14, 40881–40889 CrossRef CAS PubMed .
  73. E. A. Torres and A. J. Ramirez, Sci. Technol. Weld. Join., 2011, 16, 68–78 CrossRef .
  74. M. J. Zachman, Z. Y. Tu, L. A. Archer and L. F. Kourkoutis, ACS Energy Lett., 2020, 5, 1224–1232 CrossRef CAS .
  75. B. W. He, Y. X. Zhang, X. Liu and L. W. Chen, ChemCatChem, 2020, 12, 1853–1872 CrossRef CAS .
  76. X. Y. Wu, S. M. Li, B. Yang and C. M. Wang, Electrochem. Energy Rev., 2019, 2, 467–491 CrossRef CAS .
  77. H. W. Gao, X. Ai, H. C. Wang, W. Q. Li, P. Wei, Y. Cheng, S. W. Gui, H. Yang, Y. Yang and M. S. Wang, Nat. Commun., 2022, 13, 5050 CrossRef CAS PubMed .
  78. M. Gu, L. R. Parent, B. L. Mehdi, R. R. Unocic, M. T. McDowell, R. L. Sacci, W. Xu, J. G. Connell, P. H. Xu, P. Abellan, X. L. Chen, Y. H. Zhang, D. E. Perea, J. E. Evans, L. J. Lauhon, J. G. Zhang, J. Liu, N. D. Browning, Y. Cui, I. Arslan and C. M. Wang, Nano Lett., 2013, 13, 6106–6112 CrossRef CAS PubMed .
  79. Y. Gong, J. N. Zhang, L. W. Jiang, J. A. Shi, Q. H. Zhang, Z. Z. Yang, D. L. Zou, J. Y. Wang, X. Q. Yu, R. J. Xiao, Y. S. Hu, L. Gu, H. Li and L. Q. Chen, J. Am. Chem. Soc., 2017, 139, 4274–4277 CrossRef CAS PubMed .
  80. C. Ma, Y. Q. Cheng, K. B. Yin, J. Luo, A. Sharafi, J. Sakamoto, J. C. Li, K. L. More, N. J. Dudney and M. F. Chi, Nano Lett., 2016, 16, 7030–7036 CrossRef CAS PubMed .
  81. X. Y. Tao, Y. Y. Liu, W. Liu, G. M. Zhou, J. Zhao, D. C. Lin, C. X. Zu, O. Sheng, W. K. Zhang, H. W. Lee and Y. Cui, Nano Lett., 2017, 17, 2967–2972 CrossRef CAS PubMed .
  82. Y. Yang, J. Cui, H. J. Guo, X. Shen, Y. Yao, R. C. Yu and R. Wen, J. Mater. Chem. A, 2021, 9, 15038–15044 RSC .
  83. X. M. Liu, R. Garcia-Mendez, A. R. Lupini, Y. Q. Cheng, Z. D. Hood, F. D. Han, A. Sharafi, J. C. Idrobo, N. J. Dudney, C. S. Wang, C. Ma, J. Sakamoto and M. F. Chi, Nat. Mater., 2021, 20, 1485–1490 CrossRef CAS PubMed .
  84. L. L. Wang, R. C. Xie, B. B. Chen, X. R. Yu, J. Ma, C. Li, Z. W. Hu, X. W. Sun, C. J. Xu, S. M. Dong, T. S. Chan, J. Luo, G. L. Cui and L. Q. Chen, Nat. Commun., 2020, 11, 5889 CrossRef CAS PubMed .
  85. C. Y. Huang, Y. T. Tseng, H. Y. Lo, J. K. Chang and W. W. Wu, Nano Energy, 2020, 71, 104625 CrossRef CAS .
  86. Y. Z. Li, Y. B. Li, A. L. Pei, K. Yan, Y. M. Sun, C. L. Wu, L. M. Joubert, R. Chin, A. L. Koh, Y. Yu, J. Perrino, B. Butz, S. Chu and Y. Cui, Science, 2017, 358, 506–510 CrossRef CAS PubMed .
  87. T. W. Hansen and J. B. Wagner, ACS Catal., 2014, 4, 1673–1685 CrossRef CAS .
  88. L. Mele, S. Konings, P. Dona, F. Evertz, C. Mitterbauer, P. Faber, R. Schampers and J. R. Jinschek, Microsc. Res. Tech., 2016, 79, 239–250 CrossRef PubMed .
  89. H. Saka, T. Kamino, S. Arai and K. Sasaki, MRS Bull., 2008, 33, 93–100 CrossRef CAS .
  90. S. Mehraeen, J. T. McKeown, P. V. Deshmukh, J. E. Evans, P. Abellan, P. H. Xu, B. W. Reed, M. L. Taheri, P. E. Fischione and N. D. Browning, Microsc. Microanal., 2013, 19, 470–478 CrossRef CAS PubMed .
  91. M. Ahmadi, F. Tichelaar and H. Zandbergen, Microsc. Microanal., 2021, 27, 242–245 CrossRef .
  92. Q. S. Dai, J. M. Yao, C. C. Du, H. J. Ye, Z. Y. Gao, J. Zhao, J. Z. Chen, Y. Su, H. Li, X. J. Fu, J. T. Yan, D. D. Zhu, X. D. Zhang, M. Y. Li, Z. Y. Luo, H. L. Qiu, Q. Huang, L. Q. Zhang, Y. F. Tang and J. Y. Huang, Adv. Funct. Mater., 2022, 2208682,  DOI:10.1002/adfm.202208682 .
  93. W. H. Li, M. S. Li, Y. F. Hu, J. Lu, A. Lushington, R. Y. Li, T. P. Wu, T. K. Sham and X. L. Sun, Small Methods, 2018, 2, 1700341 CrossRef .
  94. M. T. Xia, T. T. Liu, N. Peng, R. T. Zheng, X. Cheng, H. J. Zhu, H. X. Yu, M. Shui and J. Shu, Small Methods, 2019, 3, 1900119 CrossRef .
  95. P. Bleith, H. Kaiser, P. Novak and C. Villevieille, Electrochim. Acta, 2015, 176, 18–21 CrossRef CAS .
  96. V. Avila, B. Yoon, S. Ghose, R. Raj and L. M. Jesus, J. Eur. Ceram. Soc., 2021, 41, 4552–4557 CrossRef CAS .
  97. G. Ferraresi, S. Uhlenbruck, C. L. Tsai, P. Novak and C. Villevieille, Batteries Supercaps, 2020, 3, 557–565 CrossRef CAS .
  98. Y. Ma, S. W. Li and B. Q. Wei, Nanoscale, 2019, 11, 20429–20436 RSC .
  99. M. Rawlence, A. N. Filippin, A. Wackerlin, T. Y. Lin, E. Cuervo-Reyes, A. Remhof, C. Battaglia, J. L. M. Rupp and S. Buecheler, ACS Appl. Mater. Interfaces, 2018, 10, 13720–13728 CrossRef CAS PubMed .
  100. J. Sastre, T. Y. Lin, A. N. Filippin, A. Priebe, E. Ayancini, J. Michler, A. N. Tiwari, Y. E. Rornanyuk and S. Buecheler, ACS Appl. Energy Mater., 2019, 2, 8511–8524 CrossRef CAS .
  101. F. R. Sun, Y. B. Yang, S. Zhao, Y. T. Wang, M. X. Tang, Q. Z. Huang, Y. Ren, H. Su, B. Y. Wang, N. Zhao, X. X. Guo and H. J. Yu, ACS Energy Lett., 2022, 7, 2835–2844 CrossRef CAS .
  102. P. Barai, T. Fister, Y. J. Liang, J. Libera, M. Wolfman, X. P. Wang, J. Garcia, H. Iddir and V. Srinivasan, Chem. Mater., 2021, 33, 4337–4352 CrossRef CAS .
  103. J. Cai, B. Polzin, L. Fan, L. Yin, Y. Liang, X. Li, Q. Liu, S. E. Trask, Y. Liu, Y. Ren, X. Meng and Z. Chen, Mater. Today Energy, 2021, 20, 100669 CrossRef CAS .
  104. L. Cheng, M. Liu, A. Mehta, H. L. Xin, F. Lin, K. Persson, G. Y. Chen, E. J. Crumlin and M. Doeff, ACS Appl. Energy Mater., 2018, 1, 7244–7252 CrossRef CAS .
  105. S. Hong, S. H. Song, M. Cho, S. Kim, S. H. Yu, D. Lee and H. Kim, Small, 2021, 17, 2103306 CrossRef CAS PubMed .
  106. G. O. Park, J. Yoon, E. Park, S. B. Park, H. Kim, K. H. Kim, X. Jin, T. J. Shin, H. Kim, W. S. Thon and J. M. Kim, ACS Nano, 2015, 9, 5470–5477 CrossRef CAS PubMed .
  107. H. Jung, B. Lee, M. Lengyel, R. Axelbaum, J. Yoo, Y. S. Kim and Y. S. Jun, J. Mater. Chem. A, 2018, 6, 4629–4635 RSC .
  108. T. Foroozan, S. Sharifi-Asl and R. Shahbazian-Yassar, J. Power Sources, 2020, 461, 228135 CrossRef CAS .
  109. M. Holler, M. Guizar-Sicairos, E. H. R. Tsai, R. Dinapoli, E. Muller, O. Bunk, J. Raabe and G. Aeppli, Nature, 2017, 543, 402–406 CrossRef CAS PubMed .
  110. Y. S. Cohen, Y. Cohen and D. Aurbach, J. Phys. Chem. B, 2000, 104, 12282–12291 CrossRef CAS .
  111. M. B. Dixit, B. S. Vishugopi, W. Zaman, P. Kenesei, J. S. Park, J. Almer, P. P. Mukherjee and K. B. Hatzell, Nat. Mater., 2022, 21, 1298–1305 CrossRef CAS PubMed .
  112. J. Charbonnel, N. Darmet, C. Deilhes, L. Broche, M. Reytier, P. X. Thivel and R. Vincent, ACS Appl. Energy Mater., 2022, 5, 10862–10871 CrossRef CAS .
  113. K. Dong, M. Osenberg, F. Sun, H. Markotter, C. J. Jafta, A. Hilger, T. Arlt, J. Banhart and I. Manke, Nano Energy, 2019, 62, 11–19 CrossRef CAS .
  114. V. Wood, Nat. Rev. Mater., 2018, 3, 293–295 CrossRef .
  115. K. Suzuki, B. Barbiellini, Y. Orikasa, S. Kaprzyk, M. Itou, K. Yamamoto, Y. J. Wang, H. Hafiz, Y. Uchimoto, A. Bansil, Y. Sakurai and H. Sakurai, J. Appl. Phys., 2016, 119, 025103 CrossRef .
  116. H. Shen, E. Y. Yi, M. Amores, L. Cheng, N. Tamura, D. Y. Parkinson, G. Y. Chen, K. Chen and M. Doeff, J. Mater. Chem. A, 2019, 7, 20861–20870 RSC .
  117. H. Shen, E. Y. Yi, S. Heywood, D. Y. Parkinson, G. Y. Chen, N. Tamura, S. Sofie, K. Chen and M. M. Doeff, ACS Appl. Mater. Interfaces, 2020, 12, 3494–3501 CrossRef CAS PubMed .
  118. E. Yi, H. Shen, S. Heywood, J. Alvarado, D. Y. Parkinson, G. Y. Chen, S. W. Sofie and M. M. Doeff, ACS Appl. Energy Mater., 2020, 3, 170–175 CrossRef CAS .
  119. W. Zaman, N. Hortance, M. B. Dixit, V. De Andrade and K. B. Hatzell, J. Mater. Chem. A, 2019, 7, 23914–23921 RSC .
  120. Y. G. Huang, M. C. Ma and Y. L. Guo, J. Polym. Sci., 2020, 58, 466–477 CrossRef CAS .
  121. J. Li, Y. J. Cai, Y. Y. Cui, H. Wu, H. R. Da, Y. J. Yang, H. T. Zhang and S. J. Zhang, Nano Energy, 2022, 95, 107027 CrossRef CAS .
  122. F. Y. Shen, M. B. Dixit, X. H. Xiao and K. B. Hatzell, ACS Energy Lett., 2018, 3, 1056–1061 CrossRef CAS .
  123. S. Hao, J. J. Bailey, F. Iacoviello, J. F. Bu, P. S. Grant, D. J. L. Brett and P. R. Shearing, Adv. Funct. Mater., 2021, 31, 2007564 CrossRef CAS .
  124. B. S. Vishnugopi, M. B. Dixit, F. Hao, B. Shyam, J. B. Cook, K. B. Hatzell and P. P. Mukherjee, Adv. Energy Mater., 2022, 12, 2102825 CrossRef CAS .
  125. M. B. Dixit, B. S. Vishugopi, W. Zaman, P. Kenesei, J. S. Park, J. Almer, P. P. Mukherjee and K. B. Hatzell, Nat. Mater., 2022, 21, 1298–1305 CrossRef CAS PubMed .
  126. M. B. Dixit, A. Verma, W. Zaman, X. L. Zhong, P. Kenesei, J. S. Park, J. Almer, P. P. Mukherjee and K. B. Hatzell, ACS Appl. Energy Mater., 2020, 3, 9534–9542 CrossRef CAS .
  127. M. Muller, J. Schmieg, S. Dierickx, J. Joos, A. Weber, D. Gerthsen and E. Ivers-Tiffee, ACS Appl. Mater. Interfaces, 2022, 14, 14739–14752 CrossRef PubMed .
  128. Y. Zhang, Y. Shi, X. C. Hu, W. P. Wang, R. Wen, S. Xin and Y. G. Guo, Adv. Energy Mater., 2020, 10, 1903325 CrossRef CAS .
  129. P. S. Bagus, E. S. Ilton and C. J. Nelin, Surf. Sci. Rep., 2013, 68, 273–304 CrossRef CAS .
  130. R. Endo, T. Ohnishi, K. Takada and T. Masuda, J. Phys. Chem. Lett., 2020, 11, 6649–6654 CrossRef CAS PubMed .
  131. J. G. Connell, T. Fuchs, H. Hartmann, T. Krauskopf, Y. S. Zhu, J. Sann, R. Garcia-Mendez, J. Sakamoto, S. Tepavcevic and J. Janek, Chem. Mater., 2020, 32, 10207–10215 CrossRef CAS .
  132. Z. H. Gao, Y. Bai, H. Y. Fu, J. Y. Yang, T. Ferber, J. R. Feng, W. Jaegermann and Y. H. Huang, Adv. Funct. Mater., 2022, 32, 2112113 CrossRef CAS .
  133. G. Vardar, W. J. Bowman, Q. Y. Lu, J. Y. Wang, R. J. Chater, A. Aguadero, R. Seibert, J. Terry, A. Hunt, I. Waluyo, D. D. Fong, A. Jarry, E. J. Crumlin, S. L. Hellstrom, Y. M. Chiang and B. Yildiz, Chem. Mater., 2018, 30, 6259–6276 CrossRef CAS .
  134. A. A. Delluva, J. Dudoff, G. Teeter and A. Holewinski, ACS Appl. Mater. Interfaces, 2020, 12, 24992–24999 CrossRef CAS PubMed .
  135. M. Naguib, A. Sharafi, E. C. Self, H. M. Meyer, J. Sakamoto and J. Nanda, ACS Appl. Mater. Interfaces, 2019, 11, 42042–42048 CrossRef CAS PubMed .
  136. D. Hao, Y. Liu, S. Gao, H. Arandiyan, X. Bai, Q. Kong, W. Wei, P. K. Shen and B.-J. Ni, Mater. Today, 2021, 46, 212–233 CrossRef CAS .
  137. S. A. Pervez, B. P. Vinayan, M. A. Cambaz, G. Melinte, T. Diemant, T. Braun, G. Karkera, R. J. Behm and M. Fichtner, J. Mater. Chem. A, 2020, 8, 16451–16462 RSC .
  138. R. Dubey, J. Sastre, C. Cancellieri, F. Okur, A. Forster, L. Pompizii, A. Priebe, Y. E. Romanyuk, L. P. H. Jeurgens, M. V. Kovalenko and K. V. Kravchyk, Adv. Energy Mater., 2021, 11, 2102086 CrossRef CAS .
  139. J. Y. Wen, Y. Huang, J. Duan, Y. M. Wu, W. Luo, L. H. Zhou, C. C. Hu, L. Q. Huang, X. Y. Zheng, W. J. Yang, Z. Y. Wen and Y. H. Huang, ACS Nano, 2019, 13, 14549–14556 CrossRef CAS PubMed .
  140. J. Y. Wen, L. Q. Huang, Y. Huang, W. Luo, H. Y. Huo, Z. F. Wang, X. Y. Zheng, Z. Y. Wen and Y. H. Huang, Energy Storage Mater., 2022, 45, 934–940 CrossRef .
  141. T. R. Wang, J. Duan, B. Zhang, W. Luo, X. Ji, H. H. Xu, Y. Huang, L. Q. Huang, Z. Y. Song, J. Y. Wen, C. S. Wang, Y. H. Huang and J. B. Goodenough, Energy Environ. Sci., 2022, 15, 1325–1333 RSC .
  142. J. Leng, H. Y. Wang, H. M. Liang, Z. Q. Xiao, S. T. Wang, Z. T. Zhang and Z. L. Tang, ACS Appl. Energy Mater., 2022, 5, 5108–5116 CrossRef CAS .
  143. M. Cheng, Y. Z. Jiang, W. T. Yao, Y. F. Yuan, R. Deivanayagam, T. Foroozan, Z. N. Huang, B. Song, R. Rojaee, T. Shokuhfar, Y. Y. Pan, J. Lu and R. Shahbazian-Yassar, Adv. Mater., 2018, 30, 1800615 CrossRef PubMed .
  144. J. Li, H. T. Zhang, Y. Y. Cui, H. R. Da, Y. J. Cai and S. J. Zhang, Chem. Eng. J., 2022, 450, 138457 CrossRef CAS .
  145. S. Kim, J. S. Kim, L. Miara, Y. Wang, S. K. Jung, S. Y. Park, Z. Song, H. Kim, M. Badding, J. Chang, V. Roev, G. Yoon, R. Kim, J. H. Kim, K. Yoon, D. Im and K. Kang, Nat. Commun., 2022, 13, 12 CrossRef PubMed .
  146. H. P. Zheng, G. Y. Li, R. X. Ouyang, Y. Han, H. Zhu, Y. M. Wu, X. Huang, H. Z. Liu and H. A. Duan, Adv. Funct. Mater., 2022, 32, 2205778 CrossRef CAS .
  147. Y. Gao, S. Y. Sun, X. Zhang, Y. F. Liu, J. J. Hu, Z. G. Huang, M. X. Gao and H. G. Pan, Adv. Funct. Mater., 2021, 31, 2009692 CrossRef CAS .
  148. Y. B. Hu, T. S. Feng, L. Xu, L. F. Zhang and L. L. Luo, ACS Appl. Mater. Interfaces, 2022, 14, 41978–41987 CrossRef CAS PubMed .
  149. C. Loho, R. Djenadic, M. Bruns, O. Clemens and H. Hahn, J. Electrochem. Soc., 2017, 164, A6131–A6139 CrossRef CAS .
  150. J. M. Tao, D. Y. Wang, Y. M. Yang, J. X. Li, Z. G. Huang, S. Mathur, Z. S. Hong and Y. B. Lin, Adv. Sci., 2022, 9, 2103786 CrossRef CAS PubMed .
  151. J. Q. Sun, X. M. Yao, C. H. He, Y. G. Li, Q. H. Zhang, C. Y. Hou, Y. Qiu and H. Z. Wang, J. Power Sources, 2022, 545, 231928 CrossRef CAS .
  152. J. Zhong, H. Zhang, X. H. Sun and S. T. Lee, Adv. Mater., 2014, 26, 7786–7806 CrossRef CAS PubMed .
  153. F. Lin, Y. J. Liu, X. Q. Yu, L. Cheng, A. Singer, O. G. Shpyrko, H. L. L. Xing, N. Tamura, C. X. Tian, T. C. Weng, X. Q. Yang, Y. S. Meng, D. Nordlund, W. L. Yang and M. M. Doeff, Chem. Rev., 2017, 117, 13123–13186 CrossRef CAS PubMed .
  154. N. Zhang, X. H. Long, Z. Wang, P. F. Yu, F. D. Han, J. M. Fu, G. X. Ren, Y. R. Wu, S. Zheng, W. C. Huang, C. S. Wang, H. Li and X. S. Liu, ACS Appl. Energy Mater., 2018, 1, 5968–5976 CrossRef CAS .
  155. X. Li, Z. H. Ren, M. N. Banis, S. X. Deng, Y. Zhao, Q. Sun, C. H. Wang, X. F. Yang, W. H. Li, J. W. Liang, X. N. Li, Y. P. Sun, K. Adair, R. Y. Li, Y. F. Hu, T. K. Sham, H. Huang, L. Zhang, S. G. Lu, J. Luo and X. L. Sun, ACS Energy Lett., 2019, 4, 2480–2488 CrossRef CAS .
  156. G. Binnig, C. F. Quate and C. Gerber, Phys. Rev. Lett., 1986, 56, 930–933 CrossRef PubMed .
  157. F. J. Giessibl, Rev. Mod. Phys., 2003, 75, 949–983 CrossRef CAS .
  158. A. Alessandrini and P. Facci, Meas. Sci. Technol., 2005, 16, R65–R92 CrossRef CAS .
  159. Z. Y. Zhang, S. Said, K. Smith, R. Jervis, C. A. Howard, P. R. Shearing, D. J. L. Brett and T. S. Miller, Adv. Energy Mater., 2021, 11, 2101518 CrossRef CAS .
  160. X. R. Liu, D. Wang and L. J. Wan, Sci. Bull., 2015, 60, 839–849 CrossRef CAS .
  161. C. Shen, Y. B. Huang, J. R. Yang, M. J. Chen and Z. P. Liu, Energy Storage Mater., 2021, 39, 271–277 CrossRef .
  162. X. G. Hao, Q. Zhao, S. M. Su, S. Q. Zhang, J. B. Ma, L. Shen, Q. P. Yu, L. Zhao, Y. Liu, F. Y. Kang and Y. B. He, Adv. Energy Mater., 2019, 9, 1901604 CrossRef .
  163. Z. K. Zhao, Z. Y. Wen, X. H. Liu, H. Yang, S. Chen, C. L. Li, H. J. Lv, F. Wu, B. R. Wu and D. B. Mu, Chem. Eng. J., 2021, 405, 127031 CrossRef CAS .
  164. L. Q. Zhang, T. T. Yang, C. C. Du, Q. N. Liu, Y. S. Tang, J. Zhao, B. L. Wang, T. W. Chen, Y. Sun, P. Jia, H. Li, L. Geng, J. Z. Chen, H. J. Ye, Z. F. Wang, Y. S. Li, H. M. Sun, X. M. Li, Q. S. Dai, Y. F. Tang, Q. M. Peng, T. D. Shen, S. L. Zhang, T. Zhu and J. Y. Huang, Nat. Nanotechnol., 2020, 15, 94–98 CrossRef CAS PubMed .
  165. H. Y. Huo, J. Gao, N. Zhao, D. X. Zhang, N. G. Holmes, X. N. Li, Y. P. Sun, J. M. Fu, R. Y. Li, X. X. Guo and X. L. Sun, Nat. Commun., 2021, 12, 176 CrossRef CAS PubMed .
  166. M. B. Dixit, W. Zaman, N. Hortance, S. Vujic, B. Harkey, F. Y. Shen, W. Y. Tsai, V. De Andrade, X. C. Chen, N. Balke and K. B. Hatzell, Joule, 2020, 4, 207–221 CrossRef CAS .
  167. X. N. Shi, W. H. Qing, T. Marhaba and W. Zhang, Electrochim. Acta, 2020, 332, 135472 CrossRef CAS .
  168. B. Krueger, L. Balboa, J. F. Dohmann, M. Winter, P. Bieker, G. Wittstock and A. Bar, ChemElectroChem, 2020, 7, 3543–3544 CrossRef .
  169. H. J. Chang, A. J. Ilott, N. M. Trease, M. Mohammadi, A. Jerschow and C. P. Grey, J. Am. Chem. Soc., 2015, 137, 15209–15216 CrossRef CAS PubMed .
  170. O. Pecher, J. Carretero-Gonzalez, K. J. Griffith and C. P. Grey, Chem. Mater., 2017, 29, 213–242 CrossRef CAS .
  171. M. Liu, Z. Cheng, S. Ganapathy, C. Wang, L. A. Haverkate, M. Tulodziecki, S. Unnikrishnan and M. Wagemaker, ACS Energy Lett., 2019, 4, 2336–2342 CrossRef CAS .
  172. B. Stanje, D. Rettenwander, S. Breuer, M. Uitz, S. Berendts, M. Lerch, R. Uecker, G. Redhammer, I. Hanzu and M. Wilkening, Ann. Phys. - Berlin, 2017, 529, 1700140 CrossRef .
  173. M. Wilkening and P. Heitjans, Chem. Phys. Chem., 2012, 13, 53–65 CrossRef CAS PubMed .
  174. C. Yu, S. Ganapathy, E. R. H. Van Eck, H. Wang, S. Basak, Z. L. Li and M. Wagemaker, Nat. Commun., 2017, 8, 1086 CrossRef PubMed .
  175. C. Xu, Z. Ahmad, A. Aryanfar, V. Viswanathan and J. R. Greer, Proc. Natl. Acad. Sci. U. S. A., 2017, 114, 57–61 CrossRef CAS PubMed .
  176. N. M. Trease, L. N. Zhou, H. J. Chang, B. Y. X. Zhu and C. P. Grey, Solid State Nucl. Magn. Reson., 2012, 42, 62–70 CrossRef CAS PubMed .
  177. J. A. Tang, S. Dugar, G. M. Zhong, N. S. Dalal, J. P. Zheng, Y. Yang and R. Q. Fu, Sci. Rep., 2013, 3, 1–6 Search PubMed .
  178. P. H. Clnen, X. Y. Feng, M. X. Tang, J. T. Rosenberg, S. O'Neill, J. Zheng, S. C. Grant and Y. Y. Hu, J. Phys. Chem. Lett., 2018, 9, 1990–1998 CrossRef PubMed .
  179. W. Chang, R. May, M. Wang, G. Thorsteinsson, J. Sakamoto, L. Marbella and D. Steingart, Nat. Commun., 2021, 12, 6369 CrossRef CAS PubMed .
  180. B. Karasulu, S. P. Emge, M. F. Groh, C. P. Grey and A. J. Morris, J. Am. Chem. Soc., 2020, 142, 3132–3148 CrossRef CAS PubMed .
  181. D. W. Wang, G. M. Zhong, W. K. Pang, Z. P. Guo, Y. X. Li, M. J. McDonald, R. Q. Fu, J. X. Mi and Y. Yang, Chem. Mater., 2015, 27, 6650–6659 CrossRef CAS .
  182. Y. Meesala, Y. K. Liao, A. Jena, N. H. Yang, W. K. Pang, S. F. Hu, H. Chang, C. E. Liu, S. C. Liao, J. M. Chen, X. X. Guo and R. S. Liu, J. Mater. Chem. A, 2019, 7, 8589–8601 RSC .
  183. G. Larraz, A. Orera, J. Sanz, I. Sobrados, V. Diez-Gomez and M. L. Sanjuan, J. Mater. Chem. A, 2015, 3, 5683–5691 RSC .
  184. B. Dong, A. R. Haworth, S. R. Yeandel, M. P. Stockham, M. S. James, J. W. Xiu, D. W. Wang, P. Goddard, K. E. Johnston and P. R. Slater, J. Mater. Chem. A, 2022, 10, 11172–11185 RSC .
  185. X. Xiang, F. Chen, Q. Shen, L. M. Zhang and C. L. Chen, Mater. Res. Express, 2019, 6, 085546 CrossRef CAS .
  186. P. Bottke, D. Rettenwander, W. Schmidt, G. Amthauer and M. Wilkening, Chem. Mater., 2015, 27, 6571–6582 CrossRef CAS .
  187. P. Posch, S. Lunghammer, S. Berendts, S. Ganschow, G. J. Redhammer, A. Wilkening, M. Lerch, B. Gadermaier, D. Rettenwander and H. M. R. Wilkening, Energy Storage Mater., 2020, 24, 220–228 CrossRef .
  188. E. J. Cheng, T. Kimura, M. Shoji, H. Ueda, H. Munakata and K. Kanamura, ACS Appl. Mater. Interfaces, 2020, 12, 10382–10388 CrossRef CAS PubMed .
  189. G. Foran, A. Mery, M. Bertrand, S. Rousselot, D. Lepage, D. Ayme-Perrot and M. Dolle, ACS Appl. Mater. Interfaces, 2022, 14, 43226–43236 CrossRef CAS PubMed .
  190. J. Zheng and Y. Y. Hu, ACS Appl. Mater. Interfaces, 2018, 10, 4113–4120 CrossRef CAS PubMed .
  191. J. Zagorski, J. M. L. del Amo, M. J. Cordill, F. Aguesse, L. Buannic and A. Llordes, ACS Appl. Energy Mater., 2019, 2, 1734–1746 CrossRef CAS .
  192. P. Ranque, J. Zagorski, S. Devaraj, F. Aguesse and J. M. L. del Amo, J. Mater. Chem. A, 2021, 9, 17812–17820 RSC .
  193. W. P. Chen, H. Duan, J. L. Shi, Y. M. Qian, J. Wan, X. D. Zhang, H. Sheng, B. Guan, R. Wen, Y. X. Yin, S. Xin, Y. G. Guo and L. J. Wan, J. Am. Chem. Soc., 2021, 143, 5717–5726 CrossRef CAS PubMed .
  194. M. J. Lee, D. O. Shin, J. Y. Kim, J. Oh, S. H. Kang, J. Kim, K. M. Kim, Y. M. Lee, S. O. Kim and Y. G. Lee, Energy Storage Mater., 2021, 37, 306–314 CrossRef .
  195. T. Yang, J. Zheng, Q. Cheng, Y. Y. Hu and C. K. Chan, ACS Appl. Mater. Interfaces, 2017, 9, 21773–21780 CrossRef CAS PubMed .
  196. K. W. Liu, X. Li, J. Y. Cai, Z. Z. Yang, Z. H. Chen, B. Key, Z. C. Zhang, T. L. Dzwiniel and C. Liao, ACS Energy Lett., 2021, 6, 1315–1323 CrossRef CAS .
  197. H. J. Chang, N. M. Trease, A. J. Ilott, D. L. Zeng, L. S. Du, A. Jerschow and C. P. Grey, J. Phys. Chem. C, 2015, 119, 16443–16451 CrossRef CAS .
  198. L. E. Marbella, S. Zekoll, J. Kasemchainan, S. P. Emge, P. G. Bruce and C. P. Grey, Chem. Mater., 2019, 31, 2762–2769 CrossRef CAS PubMed .
  199. C. Y. Yan, Y. Zhou, H. Cheng, R. Orenstein, P. Zhu, O. Yildiz, P. Bradford, J. Jur, N. Q. Wu, M. Dirican and X. W. Zhang, Energy Storage Mater., 2022, 44, 136–144 CrossRef .
  200. M. F. Groh, M. J. Sullivan, M. W. Gaultois, O. Pecher, K. J. Griffith and C. P. Grey, Chem. Mater., 2018, 30, 5886–5895 CrossRef CAS .
  201. 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, Adv. Energy Mater., 2021, 11, 2000806 CrossRef CAS .
  202. M. Sathiya, J. B. Leriche, E. Salager, D. Gourier, J. M. Tarascon and H. Vezin, Nat. Commun., 2015, 6, 6276 CrossRef CAS PubMed .
  203. A. Niemoller, P. Jakes, R. A. Eichel and J. Granwehr, Sci. Rep., 2018, 8, 14331 CrossRef PubMed .
  204. J. Wandt, C. Marino, H. A. Gasteiger, P. Jakes, R. A. Eichel and J. Granwehr, Energy Environ. Sci., 2015, 8, 1358–1367 RSC .
  205. J. Wolfenstine, J. L. Allen, J. Read and J. Sakamoto, J. Mater. Sci., 2013, 48, 5846–5851 CrossRef CAS .
  206. B. P. Dubey, A. Sahoo, V. Thangadurai and Y. Sharma, Solid State Ionics, 2020, 351, 115339 CrossRef CAS .
  207. Y. X. Xiang, X. Li, Y. Q. Cheng, X. L. Sun and Y. Yang, Mater. Today, 2020, 36, 139–157 CrossRef CAS .
  208. J. Mujtaba, H. Sun, G. Huang, Y. Zhao, H. Arandiyan, G. Sun, S. Xu and J. Zhu, RSC Adv., 2016, 6, 31775–31781 RSC .
  209. C. W. Wang, Y. H. Gong, J. Q. Dai, L. Zhang, H. Xie, G. Pastel, B. Y. Liu, E. Wachsman, H. Wang and L. B. Hu, J. Am. Chem. Soc., 2017, 139, 14257–14264 CrossRef CAS PubMed .
  210. R. G. Downing, G. P. Lamaze, J. K. Langland and S. T. Hwang, J. Res. Natl. Inst. Stand. Technol., 1993, 98, 109–126 CrossRef CAS PubMed .
  211. T. W. Verhallen, S. S. Lv and M. Wagemaker, Front. Energy Res., 2018, 6, 1–11 CrossRef .
  212. I. Tomandl, T. Kobayashi, A. Cannavo, J. Vacik, G. Ceccio, T. Sassa and V. Hnatowicz, J. Power Sources, 2022, 542, 231719 CrossRef CAS .
  213. S. S. Lv, T. Verhallen, A. Vasileiadis, F. Ooms, Y. L. Xu, Z. L. Li, Z. C. Li and M. Wagemaker, Nat. Commun., 2018, 9, 2152 CrossRef PubMed .
  214. Q. Li, T. C. Yi, X. L. Wang, H. Y. Pan, B. G. Quan, T. J. Liang, X. X. Guo, X. Q. Yu, O. W. R. Wang, X. J. Huang, L. Q. Chen and H. Li, Nano Energy, 2019, 63, 103895 CrossRef CAS .
  215. C. G. Chen, J. F. M. Oudenhoven, D. L. Danilov, E. Vezhlev, L. Gao, N. Li, F. M. Mulder, R. A. Eichel and P. H. L. Notten, Adv. Energy Mater., 2018, 8, 1801430 CrossRef .
  216. X. G. Han, Y. H. Gong, K. Fu, X. F. He, G. T. Hitz, J. Q. Dai, A. Pearse, B. Y. Liu, H. Wang, G. Rublo, Y. F. Mo, V. Thangadurai, E. D. Wachsman and L. B. Hu, Nat. Mater., 2017, 16, 572–579 CrossRef CAS PubMed .
  217. M. Liu, C. Wang, Z. Cheng, S. Ganapathy, L. A. Haverkate, S. Unnikrishnan and M. Wagemaker, ACS Mater. Lett., 2020, 2, 665–670 CrossRef CAS .
  218. F. D. Han, A. S. Westover, J. Yue, X. L. Fan, F. Wang, M. F. Chi, D. N. Leonard, N. Dudney, H. Wang and C. S. Wang, Nat. Energy, 2019, 4, 187–196 CrossRef CAS .
  219. J. F. M. Oudenhoven, F. Labohm, M. Mulder, R. A. H. Niessen, F. M. Mulder and P. H. L. Notten, Adv. Mater., 2011, 23, 4103–4106 CrossRef CAS PubMed .
  220. B. H. Song, I. Dhiman, J. C. Carothers, G. M. Veith, J. Liu, H. Z. Bilheux and A. Huq, ACS Energy Lett., 2019, 4, 2402–2408 CrossRef CAS .
  221. F. Sun, R. Gao, D. Zhou, M. Osenberg, K. Dong, N. Kardjilov, A. Hilger, H. Markotter, P. M. Bieker, X. F. Liu and I. Manke, ACS Energy Lett., 2019, 4, 306–316 CrossRef CAS .
  222. E. D. Rus and J. A. Dura, ACS Appl. Mater. Interfaces, 2019, 11, 47553–47563 CrossRef CAS PubMed .
  223. J. H. Yang, F. J. Mo, J. M. Hu, S. Y. Li, L. Z. Huang, F. Fang, D. L. Sun, G. A. Sun, F. Wang and Y. Song, Appl. Phys. Lett., 2022, 121, 163901 CrossRef CAS .
  224. A. Benninghoven, Surf. Sci., 1973, 35, 427–457 CrossRef CAS .
  225. S. J. Guo, Y. T. Li, B. Li, N. S. Grundish, A. M. Cao, Y. G. Sun, Y. S. Xu, Y. L. M. Ji, Y. Qiao, Q. H. Zhang, F. Q. Meng, Z. H. Zhao, D. Wang, X. Zhang, L. Gu, X. Q. Yu and L. J. Wan, J. Am. Chem. Soc., 2022, 144, 2179–2188 CrossRef CAS PubMed .
  226. H. Y. Huo, Y. Chen, R. Y. Li, N. Zhao, J. Luo, J. G. P. da Silva, R. Mucke, P. Kaghazchi, X. X. Guo and X. L. Sun, Energy Environ. Sci., 2020, 13, 127–134 RSC .
  227. H. H. Xu, Y. T. Li, A. J. Zhou, N. Wu, S. Xin, Z. Y. Li and J. B. Goodenough, Nano Lett., 2018, 18, 7414–7418 CrossRef CAS PubMed .
  228. H. P. Zheng, S. P. Wu, R. Tian, Z. M. Xu, H. Zhu, H. N. Duan and H. Z. Liu, Adv. Funct. Mater., 2020, 30, 1906189 CrossRef CAS .
  229. P. Shen, B. Zhang, Y. Wang, X. Liu, C. Yu, T. Xu, S. S. Mofarah, Y. Yu, Y. Liu, H. Sun and H. Arandiyan, J. Nanostruct. Chem., 2021, 11, 33–68 CrossRef .
  230. W. Zhang, P. Shen, L. Qian, Y. Wang, H. Arandiyan, H. Mahmoud, X. Liu, J. Wang, X. Wang, Y. Liu, H. Sun and Y. Yu, ACS Appl. Energy Mater., 2021, 4, 4551–4560 CrossRef CAS .
  231. G. Y. Tian, H. Li, B. Khalid and Z. J. Zhao, Chem. Eng. J., 2022, 430, 132803 CrossRef CAS .
  232. K. B. Dermenci, H. Tesarova, T. Samoril and S. Turan, J. Microsc., 2020, 277, 42–48 CrossRef CAS PubMed .
  233. P. Vadhva, J. Hu, M. J. Johnson, R. Stocker, M. Braglia, D. J. L. Brett and A. J. E. Rettie, ChemElectroChem, 2021, 8, 1930–1947 CrossRef CAS .
  234. T. Krauskopf, H. Hartmann, W. G. Zeier and J. Janek, ACS Appl. Mater. Interfaces, 2019, 11, 14463–14477 CrossRef CAS PubMed .
  235. D. W. Wang, K. Y. Peng, Y. P. Fu, C. B. Zhu and Y. Yang, J. Power Sources, 2021, 487, 229421 CrossRef CAS .
  236. T. Krauskopf, B. Mogwitz, C. Rosenbach, W. C. Zeier and J. Janek, Adv. Energy Mater., 2019, 9, 1902568 CrossRef CAS .
  237. Y. Lu, C. Z. Zhao, R. Zhang, H. Yuan, L. P. Hou, Z. H. Fu, X. Chen, J. Q. Huang and Q. Zhang, Sci. Adv., 2021, 7, 1–10 CAS .
  238. K. Lee, S. Han, J. Lee, S. Lee, J. Kim, Y. M. Ko, S. Kim, K. Yoon, J. H. Song, J. H. Noh and K. Kang, ACS Energy Lett., 2022, 7, 381–389 CrossRef CAS .
  239. A. A. Delluva, J. Kulberg-Savercool and A. Holewinski, Adv. Funct. Mater., 2021, 31, 2103716 CrossRef CAS .
  240. M. M. Besli, C. Usubelli, M. Metzger, V. Pande, K. Harry, D. Nordlund, S. Sainio, J. Christensen, M. M. Doeff and S. Kuppan, ACS Appl. Mater. Interfaces, 2020, 12, 20605–20612 CrossRef CAS PubMed .
  241. W. Chang, R. Mohr, A. Kim, A. Raj, G. Davies, K. Denner, J. H. Park and D. Steingart, J. Mater. Chem. A, 2020, 8, 16624–16635 RSC .
  242. J. O. Majasan, J. B. Robinson, R. E. Owen, M. Maier, A. N. P. Radhakrishnan, M. Pham, T. G. Tranter, Y. S. Zhang, P. R. Shearing and D. J. L. Brett, J. Phys-Energy, 2021, 3, 032011 CrossRef CAS .
  243. W. Chang, C. Bommier, T. Fair, J. Yeung, S. Patil and D. Steingart, J. Electrochem. Soc., 2020, 167, 090503 CrossRef CAS .
  244. R. D. Schmidt and J. Sakamoto, J. Power Sources, 2016, 324, 126–133 CrossRef CAS .
  245. K. S. C. Kuang, R. Kenny, M. P. Whelan, W. J. Cantwell and P. R. Chalker, Compos. Sci. Technol., 2001, 61, 1379–1387 CrossRef .
  246. J. A. Guemes and J. M. Menendez, Compos. Sci. Technol., 2002, 62, 959–966 CrossRef .
  247. Y. P. Li, Y. Zhang, Z. Li, Z. J. Yan, X. P. Xiao, X. T. Liu, J. Chen, Y. Shen, Q. Z. Sun and Y. H. Huang, Adv. Sci., 2022, 9, 2203247 CrossRef CAS PubMed .
  248. L. A. Blanquer, F. Marchini, J. R. Seitz, N. Daher, F. Betermier, J. Q. Huang, C. Gervillie and J. M. Tarascon, Nat. Commun., 2022, 13, 1153 CrossRef PubMed .
  249. J. Xi, J. Li, H. Sun, T. Ma, L. Deng, N. Liu, X. Huang and J. Zhang, Sens. Actuators, A, 2022, 347, 113888 CrossRef CAS .
  250. C. A. W. Li, S. Y. Yang, L. P. Xin, Z. Y. Wang, Q. Xu, L. A. Li and S. B. Wang, J. Electrochem. Soc., 2021, 168, 070551 CrossRef CAS .
  251. J. H. Cho, K. Kim, S. Chakravarthy, X. C. Xiao, J. L. M. Rupp and B. W. Sheldon, Adv. Energy Mater., 2022, 12, 2200369 CrossRef CAS .
  252. B. Gault, A. Chiaramonti, O. Cojocaru-Mirédin, P. Stender, R. Dubosq, C. Freysoldt, S. K. Makineni, T. Li, M. Moody and J. M. Cairney, Nat. Rev. Methods Primers, 2021, 1, 52 CrossRef .
  253. O. Cojocaru-Miredin, J. Schmieg, M. Mueller, A. Weber and E. Ivers-Tiffee, J. Power Sources, 2022, 539, 231417 CrossRef CAS .
  254. S. Yang, X. Min, H. Fan, J. Xiao, Y. Liu, R. Mi, X. Wu, Z. Huang, K. Xi and M. Fang, J. Mater. Chem. A, 2022, 10, 17917–17947 RSC .
  255. H. Yang, P. Tang, N. Piao, J. Li, X. Shan, K. Tai, J. Tan, H.-M. Cheng and F. Li, Mater. Today, 2022, 57, 279–294 CrossRef CAS .

This journal is © The Royal Society of Chemistry 2023
Click here to see how this site uses Cookies. View our privacy policy here.