Abstract

•Mechanical properties of extrinsic interfaces mapped for model system using AFM•Sub-surface ceramic distribution correlated to variation in surface mechanics•Electrochemical performance linked to improved adhesion at extrinsic interface•Tailoring interfacial properties will enable high-performance ASSBs ASSBs are promising high-energy density energy storage devices. Hybrid electrolytes are material systems containing an organic and inorganic ion conductor that can achieve manufacturing, transport, and mechanical requirements of ASSBs. Hybrid electrolytes contain material junctions between the organic and inorganic phases (intrinsic) and between the electrode and electrolyte (extrinsic). Both extrinsic and intrinsic interfaces in these systems witness dynamic physical and chemical changes that can adversely impact the battery performance. Limited experimental knowledge is available for these interfaces due to the relative difficulty of accessing these interfaces. Rational design of interfaces is crucial to achieve stable, high-performance ASSBs. This work evaluates extrinsic interfaces for a model PEO-LLZO hybrid electrolyte system and shows the impact of interfacial properties on electrochemical performance. Inorganic-organic hybrid solid electrolytes are promising material systems for all solid-state batteries (ASSBs). These electrolytes contain numerous solid|solid interfaces that govern transport pathways, electrode|electrolyte compatibility, and durability. This paper evaluates the role that electrode|electrolyte interfaces and electrolyte structure have on electrochemical performance. Atomic force microscopy techniques reveal how mechanical, adhesion, and morphological properties transform in a series of model hybrid solid electrolytes. These measurements are mapped to sub-surface microstructural features using synchrotron nano-tomography. Hybrid solid electrolytes with shorter polymer chains demonstrate a higher adhesion (>100 nN), Young’s Modulus (6.4 GPa), capacity (114.6 mAh/g), and capacity retention (92.9%) than hybrid electrolytes with longer polymer chains (i.e., higher molecular weight). Extrinsic interfacial properties largely dictate electrochemical performance in ASSBs. Microstructural control over inorganic constituents may provide a means for tailoring interfacial properties in hybrid solid electrolytes. Inorganic-organic hybrid solid electrolytes are promising material systems for all solid-state batteries (ASSBs). These electrolytes contain numerous solid|solid interfaces that govern transport pathways, electrode|electrolyte compatibility, and durability. This paper evaluates the role that electrode|electrolyte interfaces and electrolyte structure have on electrochemical performance. Atomic force microscopy techniques reveal how mechanical, adhesion, and morphological properties transform in a series of model hybrid solid electrolytes. These measurements are mapped to sub-surface microstructural features using synchrotron nano-tomography. Hybrid solid electrolytes with shorter polymer chains demonstrate a higher adhesion (>100 nN), Young’s Modulus (6.4 GPa), capacity (114.6 mAh/g), and capacity retention (92.9%) than hybrid electrolytes with longer polymer chains (i.e., higher molecular weight). Extrinsic interfacial properties largely dictate electrochemical performance in ASSBs. Microstructural control over inorganic constituents may provide a means for tailoring interfacial properties in hybrid solid electrolytes. Hybrid solid electrolytes are an emerging family of solid electrolytes that are promising alternatives to liquid electrolytes for energy and power dense solid-state batteries.1Dirican M. Yan C. Zhu P. Zhang X. Composite solid electrolytes for all-solid-state lithium batteries.Mater. Sci. Eng. R Rep. 2019; 136: 27-46Crossref Scopus (190) Google Scholar,2Liu X. Li X. Li H. Wu H.B. Recent progress of hybrid solid-state electrolytes for lithium batteries.Chem. Eur. J. 2018; 24: 18293-18306Crossref PubMed Scopus (97) Google Scholar These electrolytes contain an inorganic (ceramic and/or glass) ion conductor encapsulated in a polymer (organic). The inorganic ion conductor can decrease the polymer crystallinity and is also an ion conductor.3Croce F. Persi L.L. Scrosati B. Serraino-Fiory F. Plichta E. Hendrickson M.A. Role of the ceramic fillers in enhancing the transport properties of composite polymer electrolytes.Electrochim. Acta. 2001; 46: 2457-2461Crossref Scopus (605) Google Scholar, 4Croce F. Appetecchi G.B. Persi L. Scrosati B. Nanocomposite polymer electrolytes for lithium batteries.Nature. 1998; 394: 456-458Crossref Scopus (2639) Google Scholar, 5Kumar B. Scanlon L.G. Polymer-ceramic composite electrolytes.J. Power Sources. 1994; 52: 261-268Crossref Scopus (252) Google Scholar Inorganic ion conductors are typically thick (>500 μm) and hard to manufacture at scale but have excellent transport properties (conductivity and transference number)6Chen L. Li Y. Li S.P. Fan L.Z. Nan C. Goodenough J.B. PEO/garnet composite electrolytes for solid-state lithium batteries: from “ceramic-in-polymer” to “polymer-in-ceramic”.Nano Energy. 2018; 46: 176-184Crossref Scopus (810) Google Scholar, 7Schnell J. Günther T. Knoche T. Vieider C. Köhler L. Just A. Keller M. Passerini S. Reinhart G. All-solid-state lithium-ion and lithium metal batteries – paving the way to large-scale production.J. Power Sources. 2018; 382: 160-175Crossref Scopus (301) Google Scholar, 8Zhang X. Liu T. Zhang S. Huang X. Xu B. Lin Y. Xu B. Li L. Nan C.W. Shen Y. Synergistic coupling between Li6.75La3Zr1.75Ta0.25O12 and poly(vinylidene fluoride) induces high ionic conductivity, mechanical strength, and thermal stability of solid composite electrolytes.J. Am. Chem. Soc. 2017; 139: 13779-13785Crossref PubMed Scopus (537) Google Scholar (Figure 1A). In contrast, polymer ion conductors are easy to manufacture (<50 μm) and are more chemically stable in air than most inorganic materials. However, polymer electrolytes have lower transference numbers and ionic conductivities. Hybrid solid electrolytes aim to combine the advantages of both types of solid ion conductors for scalable applications of solid-state batteries (Figure 1A). All solid electrolytes consist of both intrinsic interfaces formed by grains, vacancies, and material junctions and extrinsic interfaces formed at solid|solid interfaces. Intrinsic interfaces in hybrid electrolytes are largely responsible for the ion conduction pathways9Zheng J. Hu Y.-Y. New insights into the compositional dependence of Li-Ion transport in polymer-ceramic composite electrolytes.ACS Appl. Mater. Interfaces. 2018; 10: 4113-4120Crossref PubMed Scopus (240) Google Scholar (Figure 1B). The contact region between the polymer and inorganic media is an example of an intrinsic interface. The confined interfacial polymer region demonstrates different transport properties than the bulk polymer and may augment transport properties because of Lewis acid interactions between the mobile ion and surface functional groups10Keller M. Varzi A. Passerini S. Hybrid electrolytes for lithium metal batteries.J. Power Sources. 2018; 392: 206-225Crossref Scopus (143) Google Scholar The ability to tune ion conduction pathways within the electrolyte may provide a means to ameliorate dendrite formation and control plating mechanisms. There are two percolation thresholds that exist in hybrid electrolytes: (1) long-range connectivity of inorganic particles (contact mode) (Figure 1B) and (2) long-range connectivity of the confined polymer region. The former is achieved at high loadings of 33 vol % (ceramic:polymer) and the latter is achieved at ≈4 vol % (ceramic:polymer). The literature is filled with contradicting observations supporting optimum transport through high-ceramic loading hybrid electrolytes11Zheng J. Tang M. Hu Y.-Y. Lithium ion pathway within Li7 La3 Zr2 O12 polyethylene oxide composite electrolytes.Angew. Chem. Int. Ed. Engl. 2016; 55: 12538-12542Crossref PubMed Scopus (370) Google Scholar, 12Pandian A.S. Chen X.C. Chen J. Lokitz B.S. Ruther R.E. Yang G. Lou K. Nanda J. Delnick F.M. Dudney N.J. Facile and scalable fabrication of polymer-ceramic composite electrolyte with high ceramic loadings.J. Power Sources. 2018; 390: 153-164Crossref Scopus (55) Google Scholar, 13Wang C. Yang Y. Liu X. Zhong H. Xu H. Xu Z. Shao H. Ding F. Suppression of lithium dendrite formation by using LAGP-PEO (LiTFSI) composite solid electrolyte and lithium metal anode modified by PEO (LiTFSI) in all-solid-state lithium batteries.ACS Appl. Mater. Interfaces. 2017; 9: 13694-13702Crossref PubMed Scopus (265) Google Scholar, 14Liu Q.Q. Petibon R. Du C.Y. Dahn J.R. Effects of electrolyte additives and solvents on unwanted lithium plating in lithium-ion cells.J. Electrochem. Soc. 2017; 164: A1173-A1183Crossref Scopus (74) Google Scholar and low-ceramic loading electrolytes (interface percolated).15Wang W. Yi E. Fici A.J. Laine R.M. Kieffer J. Lithium ion conducting poly (ethylene oxide)-based solid electrolytes containing active or passive ceramic nanoparticles.J. Phys. Chem. C. 2017; 121: 2563-2573Crossref Scopus (170) Google Scholar, 16Przyluski J. Siekierski M. Wieczorek W. Effective medium theory in studies of conductivity of composite polymeric electrolytes.Electrochim. Acta. 1995; 40: 2101-2108Crossref Scopus (161) Google Scholar, 17Yang T. Zheng J. Cheng Q. Hu Y.-Y. Chan C.K. Composite polymer electrolytes with Li7La3Zr2O12 garnet-type nanowires as ceramic fillers: mechanism of conductivity enhancement and role of doping and morphology.ACS Appl. Mater. Interfaces. 2017; 9: 21773-21780Crossref PubMed Scopus (257) Google Scholar, 18Zheng J. Dang H. Feng Xuyong Chien P.-H. Hu Y.-Y. Li-ion transport in a representative ceramic–polymer–plasticizer composite electrolyte: Li7La3Zr2O12–polyethylene oxide–tetraethylene glycol dimethyl ether.J. Mater. Chem. A. 2017; 5: 18457-18463Crossref Google Scholar There is currently a knowledge gap regarding transport mechanisms in these hybrid solid electrolyte systems. Extrinsic interfaces that emerge during device integration (i.e., electrode|electrolyte) are responsible for ion transport between an ion conducting phase (electrolyte) and a mixed ion and electron conducting phase (i.e., electrode). The nature of ionic transport at intrinsic and extrinsic interfaces is important for mitigating chemical and structural instabilities. Extrinsic interface instabilities are responsible for high interfacial resistances. This interfacial resistance can grow and ultimately cause a decrease in capacity or catastrophic failure because of dendrite formation,19Ren Y. Shen Y. Lin Y. Nan C. Direct observation of lithium dendrites Inside garnet-type lithium-ion solid electrolyte.Electrochem. Commun. 2015; 57: 27-30Crossref Scopus (386) Google Scholar, 20Rosso M. Brissot C. Teyssot A. Dollé M. Sannier L. Tarascon J.-M. Bouchet R. Lascaud S. Dendrite short-circuit and fuse effect on Li/polymer/Li cells.Electrochim. Acta. 2006; 51: 5334-5340Crossref Scopus (427) Google Scholar, 21Dixit M.B. Regala M. Shen F. Xiao X. Hatzell K.B. Tortuosity effects in garnet-type Li7La3Zr2O12 solid electrolytes.ACS Appl. Mater. Interfaces. 2018; 11: 2022-2030Crossref PubMed Scopus (56) Google Scholar, 22Shen F. Dixit M.B. Xiao X. Hatzell K.B. Effect of pore connectivity on Li dendrite propagation within LLZO electrolytes observed with synchrotron X-ray tomography.ACS Energy Lett. 2018; 3: 1056-1061Crossref Scopus (200) Google Scholar capacity loss via the formation of dead Li,23Zachman M.J. Tu Z. Choudhury S. Archer L.A. Kourkoutis L.F. Cryo-STEM mapping of solid–liquid interfaces and dendrites in lithium-metal batteries.Nature. 2018; 560: 345-349Crossref PubMed Scopus (408) Google Scholar regions of excess and deficient Li content,24Wang J.J. Guo Z. Xiong S.M. Characterization of the morphology of primary silicon particles using synchrotron X-ray tomography.Mater. Charact. 2017; 123: 354-359Crossref Scopus (28) Google Scholar,25Wang Z. Santhanagopalan D. Zhang W. Wang F. Xin H.L. He K. Li J. Dudney N. Meng Y.S. In situ STEM-EELS Observation of nanoscale interfacial phenomena in all-solid-state batteries.Nano Lett. 2016; 16: 3760-3767Crossref PubMed Scopus (213) Google Scholar and interface delamination.9Zheng J. Hu Y.-Y. New insights into the compositional dependence of Li-Ion transport in polymer-ceramic composite electrolytes.ACS Appl. Mater. Interfaces. 2018; 10: 4113-4120Crossref PubMed Scopus (240) Google Scholar,18Zheng J. Dang H. Feng Xuyong Chien P.-H. Hu Y.-Y. Li-ion transport in a representative ceramic–polymer–plasticizer composite electrolyte: Li7La3Zr2O12–polyethylene oxide–tetraethylene glycol dimethyl ether.J. Mater. Chem. A. 2017; 5: 18457-18463Crossref Google Scholar,26Ohta N. Takada K. Sakaguchi I. Zhang L. Ma R. Fukuda K. Osada M. Sasaki T. LiNbO3-coated LiCoO2 as cathode material for all solid-state lithium secondary batteries.Electrochem. Commun. 2007; 9: 1486-1490Crossref Scopus (512) Google Scholar,27Takada K. Nakano S. Inada T. Kajiyama A. Sasaki H. Kondo S. Watanabe M. Compatibility of lithium ion conductive sulfide glass with carbon-lithium electrode.J. Electrochem. Soc. 2003; 150: A274-A277Crossref Scopus (37) Google Scholar Currently, there are very few reports that describe the role extrinsic interfaces play on cathode performance. Engineering these interfaces to promote ion transport between an ion conductor (electrolyte) and a mixed ion and electron conductor (electrode) presents a crucial challenge toward realizing state batteries (ASSBs).28Xu L. Tang S. Cheng Y. Wang K. Liang J. Liu C. Cao Y.C. Wei F. Mai L. Interfaces in solid-state lithium batteries.Joule. 2018; 2: 1-4Abstract Full Text Full Text PDF Scopus (319) Google Scholar The physical properties of extrinsic interfaces (electrolyte|electrode) in hybrid electrolytes are largely unknown (Figure 1B). Lithiation and delithiation during battery cycling leads to volumetric expansion and contraction dynamics at the cathode.29Gu M. He Y. Zheng J. Wang C. Nanoscale silicon as anode for Li-Ion batteries: the fundamentals, promises, and challenges.Nano Energy. 2015; 17: 366-383Crossref Scopus (200) Google Scholar The extrinsic interface’s Young’s modulus is a useful property that describes how well an electrolyte can resist non-equilibrium mechanical strains.30Stokes R.J. Fennell Evans D. Fundamentals of Interfacial Engineering Advances in interfacial engineering series. Wiley, 1997Google Scholar Additionally, the modulus also provides a measure of resilience to dendrite growth.31Monroe C. Newman J. The impact of elastic deformation on deposition kinetics at lithium/polymer interfaces.J. Electrochem. Soc. 2005; 152: A396Crossref Scopus (1024) Google Scholar Adhesion is another property that quantifies the minimum force required to lose contact between the electrode and electrolyte (extrinsic interface). Adhesion affects the quality of electrode contact with the hybrid electrolyte and stability during cycling.32Stone G.M. Mullin S.A. Teran A.A. Hallinan D.T. Minor A.M. Hexemer A. Balsara N.P. Resolution of the modulus versus adhesion dilemma in solid polymer electrolytes for rechargeable lithium metal batteries.J. Electrochem. Soc. 2012; 159: A222-A227Crossref Scopus (328) Google Scholar Increasing stress at extrinsic interfaces with higher assembly pressure results in less resistive interfaces and higher critical current densities.33Gupta A. Kazyak E. Craig N. Christensen J. Dasgupta N.P. Sakamoto J. Evaluating the effects of temperature and pressure on Li/PEO-LiTFSI interfacial stability and kinetics.J. Electrochem. Soc. 2018; 165: A2801-A2806Crossref Scopus (52) Google Scholar Thus, during operation, an energy storage system can experience micromechanical stresses that can govern performance and lifetime. Techniques that can provide micro- and nanoscale spatial resolution property measurements are ideal. Raman spectroscopy is an indirect technique that can map interfaces at micron-scale resolutions and measure stress-strain relationships.34Colomban P. Analysis of strain and stress in ceramic, polymer and metal matrix composites by Raman spectroscopy.Adv. Eng. Mater. 2002; 4: 535-542Crossref Scopus (47) Google Scholar Raman spectroscopy is challenging for a transparent or highly reflecting material system (i.e., polymers) and also requires significant calibration and post-processing to obtain quantitative mechanical properties. Atomic force microscopy (AFM) addresses these challenges and can map mechanical properties with excellent spatial resolutions (≈rtip lateral and ≤1 nm vertical).35Griepentrog M. Krämer G. Cappella B. Comparison of nanoindentation and AFM methods for the determination of mechanical properties of polymers.Polym. Test. 2013; 32: 455-460Crossref Scopus (42) Google Scholar,36Young T.J. Monclus M.A. Burnett T.L. Broughton W.R. Ogin S.L. Smith P.A. The use of the PeakForceTM quantitative nanomechanical mapping AFM-based method for high-resolution Young’s modulus measurement of polymers.Meas. Sci. Technol. 2011; 22: 125703Crossref Scopus (258) Google Scholar AFM is a powerful tool that can probe numerous interfacial properties such as adhesion, Young’s modulus, and deformation in a single scan, allowing for direct correlation with topography data that can provide valuable insight into Li ion battteries.37Huang S. Wang S. Hu G. Cheong L.-Z. Shen C. Modulation of solid electrolyte interphase of lithium-ion batteries by LiDFOB and LiBOB electrolyte additives.Appl. Surf. Sci. 2018; 441: 265-271Crossref Scopus (54) Google Scholar, 38Shen C. Hu G. Cheong L.-Z. Huang S. Zhang J.-G. Wang D. Direct observation of the growth of lithium dendrites on graphite anodes by operando EC-AFM.Small Methods. 2018; 2: 1700298Crossref Scopus (101) Google Scholar, 39Shen C. Wang S. Jin Y. Han W.-Q. In situ AFM imaging of solid electrolyte interfaces on HOPG with ethylene carbonate and fluoroethylene carbonate-based electrolytes.ACS Appl. Mater. Interfaces. 2015; 7: 25441-25447Crossref PubMed Scopus (106) Google Scholar In this study, we investigate how electrode|electrolyte interfacial properties (adhesion and Young’s modulus) impact electrochemical performance in ASSBs with hybrid electrolytes. Extrinsic interfacial properties are systematically altered in a series of hybrid electrolytes via changing the molecular weight of the organic material (i.e., poly(ethylene oxide) [PEO]). By design, all electrolytes demonstrate similar transport properties but display varying interfacial properties. AFM is used to track the mechanical properties at the extrinsic interfaces ex situ. Interfacial ion transport properties are estimated using an effective mean field theory model. Variation in physical properties at the extrinsic interfaces can be correlated to the distribution of inorganic material in the hybrid electrolytes. Synchrotron X-ray nano-tomography is used to evaluate the heterogeneous distribution of inorganic particles in order to quantify the role that microstructure plays on interfacial mechanical properties. Three-dimensional (3D) structural reconstructions of hybrid solid electrolytes obtained from synchrotron experiments are combined with physics-based modeling to elucidate the origin of heterogenous interfacial properties. The findings reveal that mechanical properties at the extrinsic interface (rather than transport properties) dictate solid-state battery performance. Microstructural control over the sub-surface inorganic particles within a hybrid electrolyte may enable a pathway toward controlling extrinsic interfacial properties with spatial control. This control is important for achieving facile and long-lasting Li-metal solid-state batteries. Three model hybrid solid electrolytes were processed in a ratio of 50:50 wt % between (PEO) and Li7La3Zr2O12 (LLZO) (see Supplemental Experimental Procedures). Lithium perchlorate (LiClO4) is used as the salt in each polymer electrolyte. The polymer molecular weight is a control variable in this study that enables a systematic approach toward altering the extrinsic interfacial properties. The molecular weight of the PEO was altered between 300,000 g mol–1 (300 K), 1,000,000 g mol−1 (1 M), and 5,000,000 g mol−1 (5 M). The viscosity increases with increasing molecular weight, which can impact the processing step (Figure S1B). A shear thinning behavior for 1 M and 5 M PEO is observed and suggests entangled polymer chains at a ≈4 wt % concentration in an acetonitrile solvent. The ionic conductivity for all solid electrolytes are similar and increase from 10−6 S/cm at room temperature to 10−3 S/cm at 80°C. The electrolytes’ transport properties are all similar and independent of the molecular weight of the polymer (Figure 1C). Hybrid electrolytes show improvement over the polymer electrolytes, especially at lower temperatures.40Langer F. Bardenhagen I. Glenneberg J. Kun R. Microstructure and temperature dependent lithium ion transport of ceramic-polymer composite electrolyte for solid-state lithium ion batteries based on garnet-type Li7La3Zr2O12.Solid State Ion. 2016; 291: 8-13Crossref Scopus (80) Google Scholar Hybrid electrolytes may have higher transport properties than polymer electrolytes because of a decrease in the crystallinity in the polymer matrix as well as a lower cation-polymer interaction.11Zheng J. Tang M. Hu Y.-Y. Lithium ion pathway within Li7 La3 Zr2 O12 polyethylene oxide composite electrolytes.Angew. Chem. Int. Ed. Engl. 2016; 55: 12538-12542Crossref PubMed Scopus (370) Google Scholar,18Zheng J. Dang H. Feng Xuyong Chien P.-H. Hu Y.-Y. Li-ion transport in a representative ceramic–polymer–plasticizer composite electrolyte: Li7La3Zr2O12–polyethylene oxide–tetraethylene glycol dimethyl ether.J. Mater. Chem. A. 2017; 5: 18457-18463Crossref Google Scholar,40Langer F. Bardenhagen I. Glenneberg J. Kun R. Microstructure and temperature dependent lithium ion transport of ceramic-polymer composite electrolyte for solid-state lithium ion batteries based on garnet-type Li7La3Zr2O12.Solid State Ion. 2016; 291: 8-13Crossref Scopus (80) Google Scholar, 41Chen F. Yang D. Zha W. Zhu B. Zhang Y. Li J. Gu Y. Shen Q. Zhang L. Sadoway D.R. Solid polymer electrolytes incorporating cubic Li7La3Zr2O12 for all-solid-state lithium rechargeable batteries.Electrochim. Acta. 2017; 258: 1106-1114Crossref Scopus (155) Google Scholar, 42Choi J.H. Lee C.H. Yu J.H. Hoon Doh C. Lee S.M. Enhancement of ionic conductivity of composite membranes for all-solid-state lithium rechargeable batteries incorporating tetragonal Li7La3Zr2O12 into a polyethylene oxide matrix.J. Power Sources. 2015; 274: 458-463Crossref Scopus (215) Google Scholar Above >60°C (melting temperature of PEO [Tm]) polymer segmental motion is enhanced. At higher ceramic loading, LLZO particles can increase the tortuosity of ion transport pathways, leading to a reduction in ionic conductivity enhancement in hybrid electrolytes over the polymer electrolytes. This effect is stronger for hybrid electrolytes processed with PEO of 300 K molecular weight owing to the shorter chain lengths of these polymers. Hybrid electrolytes show lower activation energies compared to the polymer electrolytes (Table 1). Changes in the activation energies suggest the presence of an additional ion transport pathway in the system. The intrinsic interface between the organic and inorganic phase is a potential pathway for ion conduction within these hybrid electrolytes.6Chen L. Li Y. Li S.P. Fan L.Z. Nan C. Goodenough J.B. PEO/garnet composite electrolytes for solid-state lithium batteries: from “ceramic-in-polymer” to “polymer-in-ceramic”.Nano Energy. 2018; 46: 176-184Crossref Scopus (810) Google Scholar,9Zheng J. Hu Y.-Y. New insights into the compositional dependence of Li-Ion transport in polymer-ceramic composite electrolytes.ACS Appl. Mater. Interfaces. 2018; 10: 4113-4120Crossref PubMed Scopus (240) Google Scholar,11Zheng J. Tang M. Hu Y.-Y. Lithium ion pathway within Li7 La3 Zr2 O12 polyethylene oxide composite electrolytes.Angew. Chem. Int. Ed. Engl. 2016; 55: 12538-12542Crossref PubMed Scopus (370) Google Scholar,18Zheng J. Dang H. Feng Xuyong Chien P.-H. Hu Y.-Y. Li-ion transport in a representative ceramic–polymer–plasticizer composite electrolyte: Li7La3Zr2O12–polyethylene oxide–tetraethylene glycol dimethyl ether.J. Mater. Chem. A. 2017; 5: 18457-18463Crossref Google Scholar,43Yang F. Xin L. Uzunoglu A. Qiu Y. Stanciu L. Ilavsky J. Li W. Xie J. Investigation of the interaction between nafion ionomer and surface functionalized carbon black using both ultrasmall angle X-ray scattering and Cryo-TEM.ACS Appl. Mater. Interfaces. 2017; 9: 6530-6538Crossref PubMed Scopus (60) Google Scholar Conductivity of these intrinsic interfaces are estimated using an effective mean field theory model. This model considers a three-component system: (1) backbone matrix (PEO), (2) filler material (LLZO), and (3) interfacial region and predicts the effective conductivity of the resultant effective medium. The following implicit equation is solved to estimate the conductivity of the intrinsic interface44Li Yuanji Zhao Y. Cui Y. Zou Zheyi Wang D. Shi S. Screening polyethylene oxide-based composite polymer electrolytes via combining effective medium theory and Halpin-Tsai model.Comput. Mater. Sci. 2018; 144: 338-344Crossref Scopus (15) Google Scholar:(1−fc)σpol−σcompσcomp+Li*(σpol−σcomp)+fcσint−σcompσcomp+Li*(σint−σcomp)=0(Equation 1) where σpol, σcomp, and σint is the conductivity of the polymer matrix, composite, and the interface, respectively; Li* is the effective depolarization factor; and fc is the volume fraction of the inserted grains (details in Supplemental Information).Table 1Activation Energies (Ea) of Hybrid Solid Electrolyte and Polymer Solid ElectrolytesEa (eV)300K1M5M< Tm> Tm< Tm> Tm< Tm> TmPolymer0.660.340.640.360.640.36Hybrid0.480.220.480.180.460.21 Open table in a new tab The model estimates higher ionic conductivity for the intrinsic interfaces compared to the polymer and hybrid solid electrolytes at temperatures below the melting temperature of PEO (Figures 1D and S3). It is hypothesized that the interface could be a preferred ion transport pathway in this temperature range. Above the melting temperature, the interfacial conductivity shows a decrease for the 300 K and 1 M system. The differences in the behavior of the intrinsic interface could arise from the varying chain lengths of the polymer in these systems. Above the melting temperature, the ion transport should be favored within the polymer matrix as it shows higher ion conductivity. The transference numbers for the three hybrid solid electrolytes were calculated to be 0.18 ± 0.02, 0.17 ± 0.02, and 0.18 ± 0.02 (Figure S2). While the transport properties in each electrolyte are similar and independent of the molecular weight of the polymer, the electrochemical properties (stability and capacity) vary significantly. Hybrid solid electrolytes processed with 5 M molecular weight PEO failed significantly faster (≈100 h) than those processed with 300 K (>180 h) and 1 M molecular weight PEO (≈130 h). The charge passed to failure (Qfailure) is >2× times longer for hybrid solid electrolytes processed with 300 K than those processed with 5 M (Figure 2A; Table 2). Furthermore, full-cell experiments using a lithium iron phosphate cathode (LiFePO4 or LFP) || demonstrate that the capacity and capacity retention is also dependent on the molecular weight. The 5 M hybrid solid electrolyte has the lowest capacity (92.0 mAh/g, 30th cycle) and poor capacity retention over cycle life (Figures 2B and S15). The 300 K hybrid solid electrolyte shows greater capacity retention and a significantly higher capacity (114.6 mAh/g, 30th cycle). The trends in capacity and retention for full cells and charge passed to failure (Qfailure) for symmetric cells were verified with multiple tests. Since transport properties in each solid electrolyte are similar, the variability in full-cell performance and symmetric stability testing must arise from either (1) electrode microstructure or (2) extrinsic interfacial properties.Table 2Electrochemical Performance Metrics of Hybrid ElectrolytesPEO MW (g/Mol)Qfailure (C)Retention (%)300 K44.9 ± 1.992.91 M25.8 ± 1.792.95 M17.6 ± 0.788.8Qfailure is the charge passed to failure in a symmetric cell experiment. Retention is capacity retained in a full cell experiment after 30 cycles. Open table in a new tab Qfailure is the charge passed to failure in a symmetric cell experiment. Retention is capacity retained in a full cell experiment after 30 cycles. The electrodes for all tests were processed using identical cathode inks and processing conditions. The inherent microstructure thus should be identical across the three systems. Differences, if any, should arise from the cell assembly procedure (pressure driven). Galvanostatic intermittent titration technique (GITT) and electrode imaging (scanning electron microscopy) were carried out on the three systems to quantify the effective diffusion coefficients in the electrodes. Cross-sectional scanning electron microscopy images demonstrate similar LFP microstructures that are independent of the electrolyte molecular weight (Figures 2D–2F and S5). Experimentally, Li diffusion coefficients can provide an indirect measure of the electrode microstructure. Tortuous ion transport pathways will lead to lower diffusion coefficients and vice-versa. The galvanostatic intermittent titration technique is used to estimate the Li diffusion coefficient for each respective cathode:DGITT=4π[I0VMAFdE/dxdE/dt1/2]2,t≪τ(Equation 2) where I0 (A) is the applied current, A (m2) is the electrode area, F is the Faraday constant (96,485 C mol−1), and VM is the electrode molar v

Full Text
Published version (Free)

Talk to us

Join us for a 30 min session where you can share your feedback and ask us any queries you have

Schedule a call