Carbonyl-coordinating polymers Hongli Xu Jingbing Xie ...€¦ · lithium batteries. Why...

25
MRS Energy & Sustainability: A Review Journal page 1 of 25 © Materials Research Society, 2020 doi:10.1557/mre.2020.3 ABSTRACT Solid polymer electrolytes are a crucial class of compounds in the next-generation solid-state lithium batteries featured by high safety and extraordinary energy density. This review highlights the importance of carbonyl-coordinating polymer-based solid polymer electro- lytes in next-generation safe and high–energy density lithium metal batteries, unraveling their synthesis, sustainability, and electro- chemical performance. With the massive consumption of fossil fuel in vehicles nowadays, the resulted air pollution and greenhouse gases issue have now aroused the global interest on the replacement of the internal combustion engines with engine systems using renewable energy. Thus, the commercial electric vehicle market is growing fast. As the requirement for longer driving distances and higher safety in commercial electric vehicles becomes more demanding, great endeavors have been devoted to developing the next-generation solid-state lithium metal batteries using high-voltage cathode materials, e.g., high nickel (Ni) ternary active materials, LiCoO 2 , and spinel LiNi 0.5 Mn 1.5 O 4 . However, the most extensively investigated solid polymer electrolytes (SPEs) are based on polyether-based polymers, especially the archetypal poly(ethylene oxide), which are still suffering from low ionic conductivity (10 −7 to 10 −6 S/cm at room temperature), limited lithium ion transference number (<0.2), and narrow electrochemical stability window (<3.9 V), restricting this type of SPEs from realizing their full potential for the next-generation lithium-based energy storage technologies. As a promising class of alternative polymer hosts for SPEs, carbonyl-coordinating polymers have been extensively researched, exhibiting unique and promising electrochemical properties. Herein, the synthesis, sustainability, and electrochemical performance of carbonyl-coordinating SPEs for high-voltage solid-state lithium batteries will be reviewed. Keywords: solid polymer electrolyte; carbonyl-coordinating polymers; polycarbonate; polyester Carbonyl-coordinating polymers for high-voltage solid-state lithium batteries: Solid polymer electrolytes Hongli Xu and Jingbing Xie, Department of Materials Science and Engineering, School of Innovation and Entrepreneurship, Academy for Advanced Interdisciplinary Studies, Southern University of Science and Technology, Shenzhen 518055, China Zhongbo Liu, Shenzhen Capchem Technology Co., Ltd., Shenzhen 518118, China Jun Wang and Yonghong Deng, Department of Materials Science and Engineering, School of Innovation and Entrepreneurship, Academy for Advanced Interdisciplinary Studies, Southern University of Science and Technology, Shenzhen 518055, China Address all correspondence to Jun Wang at [email protected] and Yonghong Deng at [email protected] (Received 22 January 2020; accepted 3 March 2020) DISCUSSION POINTS • Carbonyl-coordinating polymer-based SPEs exhibit unique electrochemical properties compared with their polyether-based analogues, making them promising for the next-generation solid-state high–energy density lithium batteries. • What makes the carbonyl-coordinating polymers outperform their polyether-based analogues in terms of electrochemical properties? • The bio-based synthesis approach and biodegradable characteristics endow the carbonyl-coordinating polymers with increasing attention from both industry and academia. REVIEW https://doi.org/10.1557/mre.2020.3 Downloaded from https://www.cambridge.org/core. IP address: 54.39.106.173, on 12 Jan 2021 at 17:29:03, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.

Transcript of Carbonyl-coordinating polymers Hongli Xu Jingbing Xie ...€¦ · lithium batteries. Why...

Page 1: Carbonyl-coordinating polymers Hongli Xu Jingbing Xie ...€¦ · lithium batteries. Why carbonyl-coordinating polymers The dawn of the investigation on SPEs was from the work of

MRS Energy & Sustainability: A Review Journalpage 1 of 25© Materials Research Society, 2020doi:10.1557/mre.2020.3

ABSTRACT

Solid polymer electrolytes are a crucial class of compounds in the next-generation solid-state lithium batteries featured by high safety and extraordinary energy density. This review highlights the importance of carbonyl-coordinating polymer-based solid polymer electro-lytes in next-generation safe and high–energy density lithium metal batteries, unraveling their synthesis, sustainability, and electro-chemical performance.

With the massive consumption of fossil fuel in vehicles nowadays, the resulted air pollution and greenhouse gases issue have now

aroused the global interest on the replacement of the internal combustion engines with engine systems using renewable energy. Thus,

the commercial electric vehicle market is growing fast. As the requirement for longer driving distances and higher safety in commercial

electric vehicles becomes more demanding, great endeavors have been devoted to developing the next-generation solid-state lithium

metal batteries using high-voltage cathode materials, e.g., high nickel (Ni) ternary active materials, LiCoO2, and spinel LiNi0.5Mn1.5O4.

However, the most extensively investigated solid polymer electrolytes (SPEs) are based on polyether-based polymers, especially the

archetypal poly(ethylene oxide), which are still suffering from low ionic conductivity (10−7 to 10−6 S/cm at room temperature), limited

lithium ion transference number (<0.2), and narrow electrochemical stability window (<3.9 V), restricting this type of SPEs from realizing

their full potential for the next-generation lithium-based energy storage technologies. As a promising class of alternative polymer hosts for

SPEs, carbonyl-coordinating polymers have been extensively researched, exhibiting unique and promising electrochemical properties.

Herein, the synthesis, sustainability, and electrochemical performance of carbonyl-coordinating SPEs for high-voltage solid-state lithium

batteries will be reviewed.

Keywords: solid polymer electrolyte; carbonyl-coordinating polymers; polycarbonate; polyester

Carbonyl-coordinating polymers for high-voltage solid-state lithium batteries: Solid polymer electrolytes

Hongli Xu and Jingbing Xie, Department of Materials Science and Engineering, School of Innovation and Entrepreneurship, Academy for Advanced Interdisciplinary Studies, Southern University of Science and Technology, Shenzhen 518055, China

Zhongbo Liu, Shenzhen Capchem Technology Co., Ltd., Shenzhen 518118, China

Jun Wang and Yonghong Deng, Department of Materials Science and Engineering, School of Innovation and Entrepreneurship, Academy for Advanced Interdisciplinary Studies, Southern University of Science and Technology, Shenzhen 518055, China

Address all correspondence to Jun Wang at [email protected] and Yonghong Deng at [email protected]

(Received 22 January 2020; accepted 3 March 2020)

DISCUSSION POINTS• Carbonyl-coordinatingpolymer-basedSPEsexhibitunique

electrochemicalpropertiescomparedwiththeirpolyether-basedanalogues,makingthempromisingforthenext-generationsolid-statehigh–energydensitylithiumbatteries.

• Whatmakesthecarbonyl-coordinatingpolymersoutperformtheirpolyether-basedanaloguesintermsofelectrochemicalproperties?

• Thebio-basedsynthesisapproachandbiodegradablecharacteristicsendowthecarbonyl-coordinatingpolymerswithincreasingattentionfrombothindustryandacademia.

RevIew

https://doi.org/10.1557/mre.2020.3Downloaded from https://www.cambridge.org/core. IP address: 54.39.106.173, on 12 Jan 2021 at 17:29:03, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.

Page 2: Carbonyl-coordinating polymers Hongli Xu Jingbing Xie ...€¦ · lithium batteries. Why carbonyl-coordinating polymers The dawn of the investigation on SPEs was from the work of

2 n MRS eNeRgy & SUSTAINABIlITy // V O L U M E 7 // e 2 // www.mrs.org/energy-sustainability-journal

Introduction

Why solid polymer electrolytes

The world overall energy consumption has increased tre-mendously over the past several decades.1 Take the USA as an example, it consumed 101.3 quadrillion Btu (British thermal unit, equals approximately 1.1 × 1018 Joule) of energy in 2018, of which 80% was based on fossil fuels (petroleum, natural gas, and coal) and 28% was used in the transportation systems.2 Such massive utilization of fossil fuel gives rise to the world severe challenges of air pollution and greenhouse gases issues (mainly carbon dioxide and methane).3 To tackle these issues and to replace the heavy dependence on fossil fuel, renewable energy techniques, i.e., the classical lithium ion batteries, have been advanced dramatically and are considered as one of the most successful energy conservation systems in human his-tory.4 However, the conventional lithium ion batteries based on liquid electrolyte systems are virtually impossible to satisfy the stringent and booming demand for higher energy density and safety from 3 C electronics, auto industry, and grid energy con-servation systems. Developing lithium ion batteries with lower cost, higher energy density, longer calendar life, and more satis-fying safety level is still in desperate need.

Many leading economies, i.e., China, Germany, Japan, and the USA, have set up their own ambitious milestones toward the next-generation high safety and high–energy density lithium battery.5 The state-of-art lithium ion battery industry is built on the liquid electrolyte systems with highly volatile and flamma-ble solvents, which also suffer from inadequate electrochemical and thermal stabilities and low ion selectivity.6,7 To solve these problems, the replacement of liquid electrolytes with solid-state electrolytes, and the employment of high–energy density anode and cathode has been regarded as the best choice.4,8–10 As one of the most promising solid-state electrolytes, solid polymer elec-trolytes (SPEs) show proximity with respect to device integra-tion, processing cost, and mechanical properties over inorganic solid electrolytes.11 The first successful landmark of com-mercial EV applications of SPE was made by the launch of pure electric vehicles (Bluecar) provided by Bolloré in 2011 in Paris.12,13 The battery using polyethylene oxide (PEO)-based SPE and LiFePO4 cathode (working plateau at 3.5 V) coupled with lithium metal anode delivers a competitive energy density of 180–200 Wh/kg over liquid electrolyte-based auto-motive batteries. As an encouraging industrial demonstration of the realistic application of solid-state batteries, SPEs are exhibiting more than just potentials toward the next-generation lithium batteries.

Why carbonyl-coordinating polymers

The dawn of the investigation on SPEs was from the work of P.V. Wright, who studied the ionic conductivity of PEO doped with sodium and potassium salts,14 and was boosted by the seminal work of M. Armand, who used this type of material in electrochemical systems, especially in lithium batteries.15 Since then, tremendous efforts from both academia and

industry have been dedicated to the polyether-related polymers, especially PEO, owing to its comparatively low Tg (≈ −60 °C) and excellent solvation of various types of lithium salts.16 PEO is an excellent complexing agent for Li+, even though the dielec-tric constant is rather low (≈5).17–19 Besides, polyethers are generally recognized as one of the most (electro)chemically stable polymer hosts against lithium metal anode, and the solid electrolyte interface (SEI) layer formed on the anode side is originated from the lithium salt decomposition and water traces.20,21 However, polyether-based SPEs are hardly to real-ize the full potentials of solid-state batteries due to several shortcomings, of which primarily are low ambient-tempera-ture ionic conductivity (<10−6 S/cm), low lithium ion trans-ference number (<0.2), and narrow electrochemical window (<3.8 V), as reviewed in many publications.7,22–28

As a promising class of alternative SPE hosts, carbonyl- coordinating polymers, primarily aliphatic polycarbonates and polyesters, endow the SPEs with superior electrochemical properties owing to their fast chain segmental motions, high dielectric constant, and unique lithium ion complexation and transport mechanism.29 Their aromatic analogues generally show very poor ionic conductivity because of the low local seg-mental motion of the rigid polymer chain and high degree of crystallinity. The electrochemical properties, i.e., ionic con-ductivity, lithium ion transference number, and electrochemi-cal stability, are critically important when one benchmarking an SPE. The unique properties of carbonyl-coordinating poly-mers will be discussed in the following sections. To the best of our knowledge, polyketones (PKs) and polyanhydrides were rarely reported as SPE hosts, despite they are also in the family of carbonyl-coordinating polymers. This is most likely due to their chemical instability, high crystallinity, poor solubility, and the resulted extremely low ionic conductivity.

In the following sections, the three main indexes in measur-ing electrochemical performance, i.e., ionic conductivity, lith-ium ion transference number, and electrochemical stability window (ESW), are compared between polyether-based SPEs and carbonyl-based SPEs.

Ionic conductivity

The ionic conductivity represents the capability of an SPE that shuttles ions. With respect to PEO-based SPE, the conduc-tion of Li+ is realized by solvating and coordinating lithium ions with the ether oxygen on PEO segments and transporting in terms of dynamically breaking/forming lithium–oxygen (Li–O) bonds by intrachain or interchain hooping in the PEO seg-ments. Owing to such dynamical displacement of the ligands for the solvation of lithium ions, the continuous polymer chain seg-mental motion (local chain motion) results in a long-range motion of lithium ions.27,30,31 The dynamical rearrangement of this complexation in ionic transport can be well simulated by the dynamic bond percolation model.32,33 Typically, the ionic conductivity of a polyether-based SPE at room temperature is in the order of 10−7 to 10−6 S/cm, which is consequentially lower than the generally recognized sufficient value of 10−4 S/cm toward usable solid-state batteries for realistic applications.34,35

https://doi.org/10.1557/mre.2020.3Downloaded from https://www.cambridge.org/core. IP address: 54.39.106.173, on 12 Jan 2021 at 17:29:03, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.

Page 3: Carbonyl-coordinating polymers Hongli Xu Jingbing Xie ...€¦ · lithium batteries. Why carbonyl-coordinating polymers The dawn of the investigation on SPEs was from the work of

MRS eNeRgy & SUSTAINABIlITy // V O L U M E 7 // e 2 // www.mrs.org/energy-sustainability-journal n 3

Compared with other ligand candidates such as carbonates and esters, PEO solvates and coordinates lithium ions with a much higher complexation strength.36 A good example is that in the electrolyte system consisting of ≈90 mol% of propylene car-bonate (PC) and ≈10 mol% of tetraethylene glycol dimethyl ether (TEGDME), most lithium ions chelate to ether units of TEGDME.17 However, such good complexation not only facili-tates high salt dissolution but also highly restricts the breaking of Li+/ether oxygen complexation by local segmental motions. Thus, the complexation renders the polymer chain a very stable coordination structure and the ions therefore inferiorly mobile, despite a low Tg (≈ −60 °C). Besides, the ion–dipole interactions between Li+/ether oxygen have significant impacts on the microscopic morphology of SPEs, and in most cases, they act as intermolecular cross-linker in SPEs to increase Tg.37,38 Hence, the favorable complexation of Li+/ether oxygen that makes PEO dissolve lithium salt well indeed also serves to restrict its segmental motion. Furthermore, PEO is a highly crystalline polymer, with a high degree of crystallization of approximately 80% at room temperature.39 Substantial evidences suggested that the dominant lithium ions transportation takes place in the amorphous region,7 due to the higher polymer chain dynamics on the amorphous nature. Although some reports presented the superior ionic conductivity of Li+ in the crystal lattice of PEO,40 prevailingly PEO crystallization is generally considered to be detrimental to lithium ions transportation due to the sac-rificed polymer chain dynamics.41

As a promising analogue, carbonyl-coordinating polymer–based SPEs generally exhibit relatively high ionic conductivity at ambient temperature.42 In carbonyl-coordinating polymer–based SPEs, the lithium ions are coordinated by the carbonyl group oxygen (primarily) and the alkoxy oxygen adjacent to the carbonyl group ether oxygen atoms (partly) on polymer seg-ments in a similar fashion to the polyether-based SPEs.28,43–45 With the processes of dynamically breaking/forming lithium–oxygen bonds, the lithium ions transport is realized by the intrachain or interchain hooping in the carbonyl-coordinating polymer segments. Owing to the continuous displacement of the ligands for the solvation of Li+ by the segmental rear-rangement of glass transitions, the long-range conduction of lithium ions is thus occurred. Compared with polyether- based SPEs, carbonyl-coordinating polymer–based SPEs show significantly faster lithium ion transport, attributed to their fast “decoupling” during ion transport and segmental dynamics.46 Generally, the ionic conductivity of carbonyl- coordinating polymer–based SPEs closely correlates to Tg and reaches a maximum value at high lithium salt loadings (polymer-in-salt systems), where the Tg is significantly depressed owing to the plasticizing effect.46–48 This may ena-ble the carbonyl-coordinating SPEs baring with high lithium salt loadings to achieve high ionic conductivity at ambient tem-perature and realize the practical applications. A comparison of the ionic conductivity as a function of the salt concentration for polyether-based (PEO) SPE and carbonyl-coordinating polymer–based (polyethylene carbonate, PEC) SPE is pre-sented in Fig. 1.

lithium ion transference number

Lithium ion transference number ( )+Lit is another fundamen-

tal electrochemical parameter.49 A high +Lit implies that the

conductivity primarily depends on the lithium ion transport rather than on the anionic mobility. As only the mobility of lith-ium ions is the effective migration, which takes part in the elec-trochemical redox reaction in electrodes, a high +Li

t (ideally, a unity of +Li

t ) is desired in lithium batteries to avoid local deple-tion and high internal resistance caused by gradient polariza-tion from anion migrations. Besides, it was found that an SPE with high +Li

t is effective in the utilization of active electrode materials and the suppression of lithium dendrite growth.50–52 Unfortunately, a low value of ≈0.2 was typically reported on the PEO/LiTFSI systems,53 implying the high anion dynamics. The reason is that the anion mobility is less restricted by the polymer segmental motions, while the lithium cations are dominantly complexed with ether oxygen as mentioned above.54,55

To date, all the carbonyl-coordinating polymer–based SPEs exhibited significantly higher +Li

t than the PEO/LiTFSI sys-tems. As discussed above, the lithium ion shuttle in PEO-based SPE system is realized by the dynamically breaking/forming lithium–oxygen (Li–O) bonds of intrachain or interchain hoop-ing and assisted by the polymer chain local segmental motions. Due to the high donor numbers of ethers, the dynamic coordi-nation cage formed by the polymer chains in polyether-based SPE systems is quite stable and hard to break.56 The lower donor numbers of carbonyl-coordinating polymers translate into an enhanced dissociation ability and favorable dynamic dislocation of lithium ions. The typical DC polarization and AC impedance results of both the PEO-based SPE and the PEC-based SPE at 60 °C are presented in Fig. 2, showing the much higher +Li

t of the PEC/LiFSI SPE system.

Figure 1. Comparison of the ionic conductivity as a function of the salt concentration for (a) polyether-based (PEO) SPE and (b) carbonyl-coordinating polymer–based (polyethylene carbonate, PEC) SPE. Reproduced with permission.46 Copyright 2014, Royal Society of Chemistry.

https://doi.org/10.1557/mre.2020.3Downloaded from https://www.cambridge.org/core. IP address: 54.39.106.173, on 12 Jan 2021 at 17:29:03, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.

Page 4: Carbonyl-coordinating polymers Hongli Xu Jingbing Xie ...€¦ · lithium batteries. Why carbonyl-coordinating polymers The dawn of the investigation on SPEs was from the work of

4 n MRS eNeRgy & SUSTAINABIlITy // V O L U M E 7 // e 2 // www.mrs.org/energy-sustainability-journal

electrochemical stability window

A sufficiently wide ESW is another prerequisite for SPEs toward high-voltage lithium batteries, as it determines the cycle life of lithium batteries.4,8,9,24,36,58,59 It should be noted that the morphological and chemical aspects of the polymer matrix and its complex interactions with lithium salts have obvious impacts on the ESW of SPEs, which should be considered when evaluat-ing an SPE host.57 In general, a lithium battery often operates to 4.3 V versus Li/Li+ and sometimes approaching 5 V in the case of using high-voltage cathode materials, which is a tough challenge for SPEs. Even though the polyether-based SPEs are (electro)chemically stable with lithium metal anode, they are

generally unable to sustain an operating voltage higher than 3.9 V.60,61 This highly restricts the further improvement toward the high–energy density systems, with the employment of high-voltage cathodes, e.g., LiNi1−x−yCoxMnyO2, LiCoO2, LiNi0.5Mn1.5O4, and Li-rich Mn-based oxides.62 Up to date, LiFePO4 seems to be the only successful cathode candidate for the utilization of PEO electrolyte.

Numerous publications on carbonyl-coordinating SPEs claimed the significant enhancement of high-voltage compatibility, with a typical onset of oxidation in the range of 4.5–5 V versus Li/Li+.63–65 However, there are still limited reports on underlying the mecha-nism of high-voltage compatibility of SPEs, even for the most investigated PEO-based SPEs.37 Ramprasad et al. presented the theoretical estimation of ESW for polyethers [PEO and polypropyl-ene oxide (PPO)] and carbonyl-coordinating SPEs [polycaprolac-tone (PCL), polyketone (PK), polymethyl methacrylate (PMMA), and polyethyl acrylate (PEA)] using first-principles density func-tional theory computations,57 as shown in Fig. 3. The carbonyl- coordinating SPEs showed improved high-voltage compatibility because the valence band minimum (VBM) is lower than the elec-trochemical potential of cathode (μC). In a similar fashion but involved the impacts of SEI and cathode electrolyte interface (CEI), Zhou et al. implied that the higher LUMO values of carbon-yl-coordinating SPEs [PMMA, polypropylene carbonate (PPC) and polyvinyl carbonate (PVCA)] compared to their polyether-based SPE (PEO) analogues were responsible for their better high- voltage compatibility.37

Benefitted from the unique structural characteristics of the carbonyl groups, the carbonyl-coordinating polymer–based SPEs present more appealing properties and surpass PEO-based analogues with respect to ionic conductivity, lithium ion transference number, and ESW, as summarized in Table 1. The differences are rooted in the different donor numbers of the

Figure 3. Energy diagram of the electrolyte interface with anode and cathode. ESW of the SPE is determined by its reduction and oxidation potentials, controlled by the conduction band maximum (CBM) and the VBM, respectively. μA and μC are the electrochemical potentials of the anode and the cathode, respectively. Adapted with permission. 57 Copyright 2019, John Wiley and Sons.

Figure 2. Comparison of the DC polarization (left) and AC impedance (right) experiment results of the (a) PEO-based SPE and (b) PEC-based SPE at 60 °C. Reproduced with permission.46 Copyright 2014, Royal Society of Chemistry.

https://doi.org/10.1557/mre.2020.3Downloaded from https://www.cambridge.org/core. IP address: 54.39.106.173, on 12 Jan 2021 at 17:29:03, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.

Page 5: Carbonyl-coordinating polymers Hongli Xu Jingbing Xie ...€¦ · lithium batteries. Why carbonyl-coordinating polymers The dawn of the investigation on SPEs was from the work of

MRS eNeRgy & SUSTAINABIlITy // V O L U M E 7 // e 2 // www.mrs.org/energy-sustainability-journal n 5

functionalities, the disparate dielectric constants, the diverse polymer segmental motion abilities and the distinct Li+/oxygen coupling energy.29

These advantages bring carbonyl-coordinating polymer–based SPEs even closer toward the implementation in commer-cial devices. As summarized in Fig. 4, carbonyl-coordinating polymer–based SPEs exhibit several potential advantages: (i) high ionic conductivity, (ii) excellent lithium ion transfer-ence number, (iii) superior electrochemical stability, (iv) versa-tile chemistry approach, and (v) environmental benignity.

Synthesis and sustainability

Carbonyl-coordinating polymers (primarily aliphatic poly-carbonate and polyester) have long been recognized as biode-gradable and sustainable polymers. This is because the ester- and the carbonate-linked polymers are among one of the

largest families of natural biomacromolecules and are capable to undergo environmentally friendly biodegradation with appropriate structures.66 A vast number of review articles on the synthesis, characterization, properties, application, and degradation of aliphatic polyesters and polycarbonates have been reported in the literature.67–85 Herein, a brief summary of the synthesis strategies and sustainability of aliphatic polyes-ters and polycarbonates, which can potentially be involved in the application of SPEs, will be discussed in the context below.

Synthesis approach toward polycarbonates

Traditionally, there are primarily three synthesis methods toward polycarbonates, i.e., polycondensation, ring-opening polymerization (ROP), and CO2–epoxide coupling reaction, as summarized in Scheme 1. Recently, Tamura et al. presented a synthesis approach for versatile polycarbonates from various diols and CO2 directly, which was promising in terms of the sim-ple and environmentally friendly chemistry.86 Aliphatic poly-carbonates were initially prepared via polycondensation involved aliphatic phosgene or derivatives, which was not a fine chemistry, and the obtained polymers generally were with low molecular weight and high polydispersity.87 By using the melt polycondensation and post transesterification techniques, high–molecular weight aliphatic polycarbonates were obtained via the polycondensation from dialkyl carbonates and aliphatic diols.88–90 In addition, the polycondensation technique enables one to prepare various structured polycarbonates straightfor-ward from different diol-functionalized aliphatic oligomers.

As an alternative, ROP of cyclic carbonates is a more effec-tive strategy to synthesize polycarbonates with high molecular weight and narrow polydispersity without the formation and removal of any by-products as in polycondensation. The most commonly used monomers for ROP were the five- and six-membered rings of cyclic carbonates, e.g., trimethylene carbonate (TMC), a well-known biodegradable material.91 The thermodynamic analysis of ROP showed that the high angular strain (Bayer’s strain) of the small (three- or four-membered) size rings had a favorable enthalpy effect, while the medium (six-membered) and larger size rings were primarily driven by entropy.92–94 Various types of catalysts such as transition metal catalysts, alkyl halides, basic and acidic organocatalysts, as well as enzyme catalysts were reported to be effective in catalyzing

Table 1. A brief comparison of electrochemical properties between carbonyl-coordinating polymer–based SPEs and polyether-based SPEs.

Carbonyl-coordinating polymer–based SPes Polyether-based SPes

σ High (10−4 S/cm at room temperature in the case of polymer-in-salt systems)

Low at room temperature (10−7 to 10−6 S/cm), high at elevated temperature (10−4 S/cm)

+Lit High (generally >0.5) Low (generally <0.2)

ESW Broad (typically >4.5 V) Narrow (generally <3.9 V)

σ: ionic conductivity; Lit + : lithium ion transference number; ESW: electrochemical stability window.

Figure 4. Structural characteristics of carbonyl-coordinating polymer–based SPEs (polycarbonates and polyesters) and their unique advantages.

https://doi.org/10.1557/mre.2020.3Downloaded from https://www.cambridge.org/core. IP address: 54.39.106.173, on 12 Jan 2021 at 17:29:03, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.

Page 6: Carbonyl-coordinating polymers Hongli Xu Jingbing Xie ...€¦ · lithium batteries. Why carbonyl-coordinating polymers The dawn of the investigation on SPEs was from the work of

6 n MRS eNeRgy & SUSTAINABIlITy // V O L U M E 7 // e 2 // www.mrs.org/energy-sustainability-journal

the ROP of cyclic carbonates.95 Depending on the selection of catalyst, the ROP could be conducted in different mechanisms including cationic, anionic, coordination–insertion, monomer activation, monomer and initiator dual activation, and enzymatic activation, in solution or in bulk.95 Notably, enzyme-catalyzed ROP was also powerful for synthesizing polycarbonates, with respect to the unique advantages of enzymes over conventional catalysts due to their milder reaction conditions, higher toler-ance for functional groups, and higher selectivity. However, the catalyzing efficiency and the control on molecular weight and polydispersity were rather restricted.77,84

As one of the most promising and sustainable strategies of green chemistry, synthesizing polycarbonates from carbon dioxide (CO2) has aroused extensive research interest to tackle the greenhouse gas pollution issue and to alleviate the shortage in conventional petroleum fuel supplies.96–99 Ever since the first report of copolymerizing CO2 with epoxide by Inoue et al. in 1969,100 this copolymerization strategy has now turned into one of the most innovative techniques involved in the utiliza-tion of CO2 in industry. The main challenge of this synthesis strategy is searching for a highly efficient and selective catalyst while suppressing the side reactions, e.g., polyether formation from epoxide, backbiting reactions leading to cyclic car-bonates.95 A bunch of organometal catalysts have been devel-oped, e.g., zinc-containing catalysts, aluminum complex, and transition metal complex.99 The mechanism of the alternating copolymerization of CO2 with epoxide catalyzed by metal

coordination complexes has been revealed in depth in several excellent reviews.98,101,102

Synthesis strategies toward polyesters

Despite the very similar chemical structure to polycar-bonates, the synthesis strategies for polyesters are slightly different. Aliphatic polyesters were primarily prepared via polycondensation and ROP, as shown in Scheme 2.66,103

The most traditional way of synthesizing polyesters involves the step-growth mechanism (polycondensation) from dicarbox-ylic acid (or dicarboxylic acid derivatives, such as diacid chlo-ride) and diols, or from an oxyacid or its ester derivatives. The study of polycondensation of aliphatic polyesters was pioneered by Carothers et al. 87 to extract the fundamental knowledge related to the step-growth polymerization and the relationship between molecular weight and extent of reaction, the stoichio-metric imbalance of functionalities. However, polycondensa-tion suffer from some major drawbacks, e.g., high reaction temperature, long reaction time, and removal of reaction by-products. Polyesters with sufficiently high molecular weight could only be achieved in the case of very high conversion and precise stoichiometry.103 Oligomers with molecular weight of several thousand (g/mol) were easily obtained, which could be further improved by chain extension treatment.104–106

To tackle these challenges, the ROP approach was developed to achieve a high molecular weight at mild reaction conditions. The monomers used in the ROP involve cyclic esters, e.g., lac-tones, cyclic diesters (lactides and glycolides), and cyclic ketene acetals. In sharp contrast to polycondensation, ROP can be car-ried out with complete monomer conversion and virtually free from side reactions, providing very good control over molecular structures, e.g., molecular weight, polydispersity, and end-group functionalization. In most cases, both the synthesis

Scheme 2. Synthesis strategies toward aliphatic polyesters. (a) Polycon-densation, including AA + BB type and AB type, and (b) ROP.

Scheme 1. Synthesis strategies for aliphatic polycarbonates. (a) Polyconden-sation, (b) ROP, (c) CO2–epoxide coupling reaction, and (d) CO2–diol coupling reaction.

https://doi.org/10.1557/mre.2020.3Downloaded from https://www.cambridge.org/core. IP address: 54.39.106.173, on 12 Jan 2021 at 17:29:03, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.

Page 7: Carbonyl-coordinating polymers Hongli Xu Jingbing Xie ...€¦ · lithium batteries. Why carbonyl-coordinating polymers The dawn of the investigation on SPEs was from the work of

MRS eNeRgy & SUSTAINABIlITy // V O L U M E 7 // e 2 // www.mrs.org/energy-sustainability-journal n 7

approaches for aliphatic polyesters were carried out in the pres-ence of catalysts, e.g., protonic acid (HCl, RCO2H, etc.), Lewis acids (AlCl3, BF3, etc.), and organometallic compounds (stan-nous octoate, etc.),107 even though some lactones polymerized spontaneously on heating. In principle, most catalysts utilized in the ROP of cyclic carbonates are also effective for the ROP of cyclic esters owing to the chemical structural similarity. The mechanism of ROP is highly dependent on the selection of cata-lysts and monomers, including cationic, anionic, free radical, active hydrogen, zwitterionic, and coordination.107 Like the enzyme-catalyzed synthesis of polycarbonates, enzymatic syn-thesis of polyesters via both the polycondensation and the ROP techniques has been studied extensively. However, enzymatic polymerization generally yielded product with low molecular weight (<10,000 g/mol) and broad molecular weight distribu-tion,66,77,79,84 which is in need for further improvement.

Sustainability

In the past several decades, environmental concerns have spurred tremendous research interest in renewable resources and green chemistry. As a promising class of sustainable alter-natives to commodity plastics,66,85 the synthesis, characteriza-tion, properties, and application of bio-based and biodegradable polycarbonates and polyesters have been studied extensively.67,79,85,103,108–112 The sustainable and eco-friendly synthesis approaches for polycarbonates and polyesters have been sum-marized above, among which enzyme-catalyzed polymeriza-tions are promising with respect to their unique advantages of low requirement of reaction conditions, high tolerance for func-tional groups, and high selectivity. However, enzyme-catalyzed polymerizations also show drawbacks, which must be improved by further study. Notably, polycarbonates can be synthesized effectively in large scale by copolymerization of CO2 with epox-ides. As one of the most abundant and renewable carbon resources, the industrial scale utilization of CO2 and the selec-tive transformation of CO2 into degradable polycarbonates are regarded as the most successful green and sustainable chemistries.95

The carbonate and ester functionalities are naturally formed, thus can go easy naturally breaking under certain con-ditions and appreciate backbone structures. As an example, Li et al. reported the enzyme-catalyzed biodegradation of poly(tri-methylene carbonate) (PTMC) in the presence of lipases from Candida antarctica.113 It was found that the PTMC degraded rapidly by Candida antarctica lipase, with 98% mass loss after 9 days, which was promising in terms of green recycling of battery materials. To date, the well-studied example of bio-degradable material is the bio-based polylactide (PLA), which is currently manufactured on a large scale by the ROP of lactide, a cyclic ester from the fermentation product based on corn or wheat, as shown in Fig. 5.80 PLA has been utilized in a variety of applications, e.g., sutures, stents, dental implants, vascular grafts, bone screws, and pins, because of its biocompatibility. Under certain conditions, such as enzyme, PLA is capable of undergoing enzyme-catalyzed biodegradation into lactic acid and further into CO2 and H2O. This great success of PLA

highlights the commercial and environmental potential for plastics sourced from plants, with renewable life cycles and which are carbon neutral. The mentioned green aspects of enzyme catalysis are expected to contribute to solving the envi-ronmental problems.

The synthesis, properties, application, and recycling of poly-esters and polycarbonates have been explored extensively over the years. As a promising class of high-voltage SPEs, aliphatic polyester and polycarbonate-based SPEs for lithium batteries have also been studied. In this review, we will summarize the recent key advances in polycarbonate- and polyester-based SPEs. The primary focus will be on the electrochemical prop-erty, lithium ion transport mechanism, and battery perfor-mance (if available). Based on our understanding and to the best of our knowledge, we tentatively suggest some challenges and perspective for aliphatic polycarbonate- and polyester-based SPEs in the last part of this review. We hope this review may give some helpful hints on understanding the correlation between structure and electrochemical performance and advancing the carbonyl-coordinating polymer-based SPEs.

Polycarbonate-based SPes

Side chain polycarbonate-based SPEs

To date, conventional liquid electrolytes are still the domi-nant player in the field of lithium batteries, of which the major-ity are based on the carbonate solvent families, such as ethylene carbonate (EC), vinylene carbonate (VC), PC, and dimethyl carbonate (DMC).36 From a chemical perspective, the poly-carbonates are essentially the linearly polymerized carbonates with high molecular weights either in the main chain or in the

Figure 5. The cycle life of PLA. Adapted with permission.80 Copyright 2007, Royal Society of Chemistry.

https://doi.org/10.1557/mre.2020.3Downloaded from https://www.cambridge.org/core. IP address: 54.39.106.173, on 12 Jan 2021 at 17:29:03, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.

Page 8: Carbonyl-coordinating polymers Hongli Xu Jingbing Xie ...€¦ · lithium batteries. Why carbonyl-coordinating polymers The dawn of the investigation on SPEs was from the work of

8 n MRS eNeRgy & SUSTAINABIlITy // V O L U M E 7 // e 2 // www.mrs.org/energy-sustainability-journal

side chain, thus can solvate, coordinate, and conduct lithium ions in a similar fashion. The strategy of polymerizing car-bonates endows polycarbonate-based SPEs with unique proper-ties in terms of lithium ion conductivity, lithium ion transference number, ESW, and mechanical properties, etc. The molecular structures of polymer hosts and electrochemical properties of some typical polycarbonate-based SPEs are summarized in Table 2.

The earlier work of applying polycarbonates for SPEs was provided by Holt et al.114 using poly(vinylene carbonate) (PVIC) as the polymer host (Entry 1 in Table 2). By integrating with crown ether (12-crown-4) and LiCF3SO3, such polycarbonate- based SPE showed considerably high ionic conductivity of 2.5 × 10−4 S/cm at ambient temperature. Shriver and Wei also reported the SPEs based on PVIC or poly(1,3-dioxolan- 2-one-4,5-diyloxalate) (PVICOX) with LiCF3SO3,115 delivering ionic conductivities of 1 × 10−6 S/cm (PVIC) and 1 × 10−4 S/cm (PVICOX) at 60 °C, respectively. The enhanced ionic con-ductivity of PVICOX/LiCF3SO3 SPE system was attributed to the larger free volume of PVICOX over the PVIC analogue. Besides, the crystalline behavior did not appear to suppress the ion conduction in such highly rigid chain SPE, which was in contrast with polyether-based SPEs.27 The ion transport was decoupled from the polymer chain segmental motion in these SPEs. By replacing the backbone of poly(vinylethylene carbonate) with highly f lexible polysilanes, the polycar-bosilanes, i.e., polySBMC and polySBDC, delivered impres-sively high ionic conductivities of 6.1 × 10−5 and 1.5 × 10−4 S/cm at 30 °C, respectively, and at high LiTFSI loading (Entry 2 in Table 2).116 As the result, the mechanical properties were severely compromised by such high lithium salt concentration.

Recently, Chai et al. reported the in situ polymerized poly (vinylene carbonate)/lithium difluoro(oxalate)borate (PVC/LiDFOB)-integrated SPE for high-voltage lithium metal bat-tery, as shown in Fig. 6 and Entry 3 of Table 2.117 A high ionic conductivity of 9.8 × 10−5 S/cm, high +Li

t of 0.57 and wide elec-trochemical window of 4.5 V were obtained at 50 °C, ensuring the outstanding electrochemical performance of LiCoO2/Li cell. The in situ technique intrinsically endowed the cell with superior interfacial stability and structural integrity, thus the high-voltage LiCoO2/Li cell delivered satisfactory rate capa-bility (0.5 C, 73 mAh/g) and long-term cycling stability (84% after 150 cycles at 0.1 C). However, the abnormally high ionic conductivity is more likely because of plasticizing effect from the unreacted VC monomers rather than the contribution from the polymer host.

Compared with their homopolymer-based analogues, SPEs from block copolymer are capable of exhibiting combined properties from each block,118 thus to achieve balanced prop-erties with respect to ionic conductivity, electrochemical sta-bility, and mechanical strength.119 Itoh et al. reported the integrated SPE from block copolymers of poly(AcIM/VC) and LiTFSI, exhibiting a significantly higher ionic conductivity value (7 × 10−4 S/cm at 80 °C) compared with poly(AcIM) homopolymer–based SPE (8.5 × 10−5 S/cm), as shown in Entry 4 of Table 2.120 The block copolymers with high VC content

(>75 mol%) were mechanically brittle, failing to perform the electrochemical measurements.

By copolymerizing poly(VC) with a series of poly(methoxy oligo(ethyleneoxy)ethyl vinyl ether) with different chain lengths, the poly(1a-g-alt-VC) alternating copolymer–based SPEs showed a significant increase in ionic conduction, as shown in Entry 5 of Table 2.121 The highest ionic conductivity reached 1.2 × 10−4 S/cm at 30 °C due to the incorporation of large amount of f lexible oligoether side chains. As a compro-mise, the electrochemical stability and lithium ion transfer-ence number were consequentially sacrificed, indicating that the ion conduction is mainly coupled with the ether seg-mental motions indeed. Wang et al. recently reported the block copolymer of poly(vinylene carbonate-acrylonitrile) [P(VC-AN)], aiming to achieve high-voltage cathode mate-rial (LiNi0.5Mn1.5O4) compatibility by taking complementary advantages of EC and nitrile functionality, as shown in Entry 6 of Table 2.122 However, the SPE most likely shows low ionic conductivity (not presented in the paper) due to the highly rigid polymer chain and possible crystallinity. Gel polymer electrolyte was prepared to perform all the electrochemical experiments and battery cycling tests. To avoid using such a highly rigid backbone structure, preparing f lexible block copolymers with pendant carbonate moieties seems to be a reasonable choice. To this end, Wegner et al. synthesized block copolymers of polyacrylate and polymethacrylate with f lexible carbonate functionalities at the side chain, as shown in Entry 7 of Table 2.123 The ionic conductivity was still insuf-ficient for the electrochemical application at room tempera-ture. Preparing gel electrolyte using PC was able to reach an order of 10−4 in ionic conductivity.

Main chain polycarbonate-based SPEs

Typically, aromatic polycarbonates exhibit high Tg and poor segmental motions due to the rigid backbone of the pol-ymer chain, which are detrimental for lithium ion transpor-tation. The exploration of applying aromatic polycarbonates as SPE hosts has been reported in the early stage. Spiegel et al. reported the solvation behavior of lithium salts using poly(dimethyl siloxane)/poly(bisphenol-A carbonate) block copolymers with LiCF3SO3 (Entry 1 in Table 3).125 Even though the ionic conductivity could be improved signifi-cantly from 7 × 10−9 S/cm to 8.1 × 10−7 S/cm (both at room temperature) by the induction of soft poly(dimethyl silox-ane) block, such ionic conductivity was still far below the essential value (10−4 S/cm). Attempts to improve the ionic conductivity of aromatic polycarbonate-based SPEs using oligo(ethylene oxide) side chains were effective (Entry 2 in Table 3); however, the electrochemical stability and lithium ion transference number were compromised.126 This is easy to be understood that the ion conduction is dominated by the ether segmental motions.121 All these results further confirm the failure of using aromatic polycarbonates as SPE host. Nevertheless, aromatic polycarbonates are potentially useful in strengthening the mechanical properties of SPEs via pre-paring block copolymer strategy.

https://doi.org/10.1557/mre.2020.3Downloaded from https://www.cambridge.org/core. IP address: 54.39.106.173, on 12 Jan 2021 at 17:29:03, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.

Page 9: Carbonyl-coordinating polymers Hongli Xu Jingbing Xie ...€¦ · lithium batteries. Why carbonyl-coordinating polymers The dawn of the investigation on SPEs was from the work of

MRS eNeRgy & SUSTAINABIlITy // V O L U M E 7 // e 2 // www.mrs.org/energy-sustainability-journal n 9

Table 2. Molecular structure and electrochemical characteristics of side chain polycarbonate-based SPEs.

entry Polymer host lithium salt σ (S/cm) eSw +Lit References

1 LiCF3SO3 PVIC: 2.5 × 10−4 (25 °C)114 N/A N/A 114 and 115

PVIC: 1.0 × 10−6 (60 °C), PVICOX: 1.0 × 10−4

(60 °C)115

2 LiCF3SO3 polySBMC: 6.1 × 10−5 (30 °C)

N/A N/A 116

LiTFSI polySBDC: 1.5 × 10−4 (30 °C)

3 LiDFOB 9.8 × 10−5 (50 °C) 4.5 V (50 °C) 0.57 117

4 LiTFSI Poly(AcIM): 8.5 × 10−5 (80 °C), 1.7 × 10−6 (30 °C)

N/A N/A 120

Poly(AcIM/VC): 7.0 × 10−4 (80 °C), 6.7 × 10−5 (30 °C)

5 LiTFSI 1.2 × 10−4 (30 °C) 4.1–4.6 V (80 °C) 0.1–0.3 (80 °C) 121

6 LiPF6 2.6 × 10−4 (25 °C) 5.2 V 0.52 122

7 LiTFSI PDOA: 3.7 × 10−6 (40 °C) N/A N/A 123

σ: ionic conductivity; ESW: electrochemical stability window; LiCF3SO3: lithium trifluoromethanesulfonate; LiDFOB: lithium difluoro(oxalate) borate; LiTFSI: lithium bis(trifluoromethanesulfonyl)imide; LiPF6: lithium hexafluorophosphate.

https://doi.org/10.1557/mre.2020.3Downloaded from https://www.cambridge.org/core. IP address: 54.39.106.173, on 12 Jan 2021 at 17:29:03, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.

Page 10: Carbonyl-coordinating polymers Hongli Xu Jingbing Xie ...€¦ · lithium batteries. Why carbonyl-coordinating polymers The dawn of the investigation on SPEs was from the work of

10 n MRS eNeRgy & SUSTAINABIlITy // V O L U M E 7 // e 2 // www.mrs.org/energy-sustainability-journal

As a contrast, aliphatic polycarbonates undergo polymer chain segmental motions much easier due to the higher molec-ular flexibility of aliphatic backbones. The simplest aliphatic polycarbonate used as SPE host was the principal ring-opening products from cyclic carbonates (EC) with the characteristic backbone structure of –O–(C=O)–O–, as shown in Entry 3 of Table 3. Practically, this polymer was more often synthesized

via the green chemistry of the copolymerization of carbon diox-ide with epoxides,85 rather than the ring-opening reaction from EC owing to its thermodynamic stability of the five-membered ring structure of EC.127 Besides, the ROP on EC is generally associated with the emission of carbon dioxide, resulting in polymer backbones with ether junctions alongside with car-bonate linkages.128 Novakov and colleagues investigated the

Figure 6. (a) In situ polymerization of PVC-based SPE and (b) self-standing SPE film. (c–f) Battery cycling performance of LiCoO2/Li cells with PVC/LiDFOB-integrated electrolyte. Reproduced with permission under CC BY-NC-ND 4.0,117 Copyright 2016, the Authors.

https://doi.org/10.1557/mre.2020.3Downloaded from https://www.cambridge.org/core. IP address: 54.39.106.173, on 12 Jan 2021 at 17:29:03, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.

Page 11: Carbonyl-coordinating polymers Hongli Xu Jingbing Xie ...€¦ · lithium batteries. Why carbonyl-coordinating polymers The dawn of the investigation on SPEs was from the work of

MRS eNeRgy & SUSTAINABIlITy // V O L U M E 7 // e 2 // www.mrs.org/energy-sustainability-journal n 11

Table 3. Molecular structures and electrochemical characteristics of main chain polyester-based SPEs.

entry Polymer host lithium salt σ (S/cm) eSw +Lit References

1 LiCF3SO3 4.4 × 10−6 (25 °C) N/A N/A 125

2 LiTFSI 1 × 10−6 to 1 × 10−4 (80 °C)

4.3–5.0 V 0.35–0.67 (80 °C)

126

3 LiClO4 9.6 × 10−4 (80 °C) N/A N/A 129

LiFSI 4.1 × 10−4 (60 °C) N/A >0.5 46, 47, 135, and 136

LiTFSI LiTFSI: 10−4 (80 °C) >4.3 V LiTFSI: 0.63 130–134

LIClO4 LIClO4: 10−5 (80 °C) LIClO4: 0.26

LiBF4 LiBF4: 10−5 (80 °C) LiBF4: 0.22

LiPF6 LiPF6: 10−5 (80 °C) LiPF6: 0.4

4 LiTFSI 10−5 (25 °C) N/A N/A 124, 137, 142, and

143

5 LiCF3SO3, LiClO4, LiBF4

3 × 10−4 (95 °C) 4.5 V N/A 64 and 144

LiSbF6 6 × 10−5 (85 °C) 5.0 V N/A 145

LiPF6 4.8 × 10−6 (98 °C) 4.5 V N/A 65

1.8 × 10−8 (25 °C)

LiTFSI 10−7 (60 °C) 5.0 V N/A 63

LiTFSI N/A N/A 0.8 (60 °C) 148

Continued

https://doi.org/10.1557/mre.2020.3Downloaded from https://www.cambridge.org/core. IP address: 54.39.106.173, on 12 Jan 2021 at 17:29:03, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.

Page 12: Carbonyl-coordinating polymers Hongli Xu Jingbing Xie ...€¦ · lithium batteries. Why carbonyl-coordinating polymers The dawn of the investigation on SPEs was from the work of

12 n MRS eNeRgy & SUSTAINABIlITy // V O L U M E 7 // e 2 // www.mrs.org/energy-sustainability-journal

entry Polymer host lithium salt σ (S/cm) eSw +Lit References

6 LiTFSI 1 × 10−4 (25 °C) 4.0 V (60 °C) N/A 149

7 LiTFSI 1.1 × 10−5 (25 °C) 4 V 0.39 150

2.0 × 10−4 (80 °C)

8 LiTFSI 1.7 × 10−4 (25 °C) 4.5 V 0.47 151

9 LiTFSI 1 × 10−3 (80 °C) 4.9 V 0.4 152

4 × 10−4 (60 °C)

4 × 10−5 (25 °C)

10 LiCF3SO3 10−4 (60 °C) N/A N/A 138

11 LiCF3SO3 10−4 (60 °C) N/A N/A 153

12 LiTFSI 1 × 10−4 (60 °C) 6 × 10−6 (20 °C)

N/A N/A 154

13 LiClO4 8.4 × 10−3 (25 °C) N/A 0.42 155

1.4 × 10−2 (50 °C)

Table 3. Continued

Continued

https://doi.org/10.1557/mre.2020.3Downloaded from https://www.cambridge.org/core. IP address: 54.39.106.173, on 12 Jan 2021 at 17:29:03, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.

Page 13: Carbonyl-coordinating polymers Hongli Xu Jingbing Xie ...€¦ · lithium batteries. Why carbonyl-coordinating polymers The dawn of the investigation on SPEs was from the work of

MRS eNeRgy & SUSTAINABIlITy // V O L U M E 7 // e 2 // www.mrs.org/energy-sustainability-journal n 13

entry Polymer host lithium salt σ (S/cm) eSw +Lit References

14 LiClO4 1 × 10−4 (30 °C) 4.7 V (60 °C) N/A 156

5 V (25 °C)

15 N/A 1.61 × 10−4 (80 °C) 4.3 V 0.86 (25 °C) 157

16 N/A 1.2 × 10−4 (70 °C) N/A 0.89 158

17 Blend of PPC, polybutadiene and PEG LiPF6 1.3 × 10−3 (25 °C) 4.5 V N/A 159

3.5 × 10−3 (80 °C)

18 PVDF/PPC blend LiPF6 4.1 × 10−3 (30 °C) 5.2 V N/A 160

19 LiODFB 1.1 × 10−3 (25 °C) 5 V 0.68 161

20 LiClO4 1.5 × 10−3 (25 °C) N/A N/A 165

Table 3. Continued

Continued

https://doi.org/10.1557/mre.2020.3Downloaded from https://www.cambridge.org/core. IP address: 54.39.106.173, on 12 Jan 2021 at 17:29:03, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.

Page 14: Carbonyl-coordinating polymers Hongli Xu Jingbing Xie ...€¦ · lithium batteries. Why carbonyl-coordinating polymers The dawn of the investigation on SPEs was from the work of

14 n MRS eNeRgy & SUSTAINABIlITy // V O L U M E 7 // e 2 // www.mrs.org/energy-sustainability-journal

Li+-carbonyl group coordination behavior in PEC/LiClO4 SPE systems using Fourier Transform Infrared (FTIR).129 Tominaga et al. extensively studied the lithium ion conduction behavior of PEC-based SPE using various types of lithium salts.46,47,124,130–140 Some lithium salts, e.g., LiBF4, LiFSI, and LiTFSI, have signif-icant plasticizing effects on PEC-based SPE systems at high con-centrations due to the declined intermolecular interactions by highly saturated ion pairs and aggregated ions and suppressed intramolecular interactions between carbonyl and CH2 arising from ion–dipole interaction.133 Distinctly, the PEC/LiCF3SO3 and PEC/LiClO4 SPE systems behaved more similarly to their polyether-based analogues, with respect to the dependencies of ionic conductivity and Tg on lithium salt loadings.130 Besides, the PEC-based SPEs showed uniquely high +Li

t value (>0.5) despite high salt concentration,134 clearly outperforming the polyether-based SPEs, which exhibited a drastically decreased

+Lit at elevated salt concentrations.141 Especially, the ionic con-ductivity of PEC/LiFSI-integrated SPE was estimated to be higher than 10−4 S/cm at room temperature, with a +Li

t value of 0.63.134 Even better electrochemical properties in terms of ionic conductivity (4 × 7 × 10−4 S/cm) and +Li

t (0.88) at 20 °C were reported by Okumura et al. based on the PEC/LiTFSI-in-tegrated SPE systems.48 These unique properties made PEC a promising SPE host, which was well summarized in a focused review by Tominaga.124 Even though high ionic conductivity and +Li

t are achievable at high lithium salt concentrations, the plasticization is accompanied by a notable suppression of Tg and deterioration of mechanical properties, as depicted in Fig. 7.

While the PEC-based SPEs surpassed their PEO-based ana-logues regarding the electrochemical properties, the versatile chemistry of the synthesis of PEC enabled one to explore various functional PEC-derivatives toward higher ionic conductivity.124 Several different side groups functionalized PECs were prepared by Tominaga et al. via the green chemistry approach starting from CO2 and a set of epoxides as electrolyte materials, as sum-marized in Entry 4 of Table 3.124,137,142,143 The side functionali-ties used were methyl, ethyl, n-propyl, methylene phenyl, etc. Depending on the selection of the substituents, the Tg’s of poly-mers were tunable within the temperature range 3–45 °C. The bulky groups (such as methylene phenyl) substituted PEC SPEs showed high Tg (≈45 °C) and low ionic conductivities with a

typical order of 10−8 S/cm at 80 °C, and the flexible side groups (such as ethyl and n-propyl) suppressed Tg (<13 °C) and brought much higher ionic conductivity in an order of 10−5 S/cm at room temperature.124 This is clearly because the bulky side substitutes impede segmental motions, whereas the inclusion of flexible side groups increases the free volume and lowers Tg. Aside from the coordination and conduction effects of lithium ions by carbonyl groups, the embodiment of ether-based side groups has consider-able enhancement on lithium ion transport. Hitherto, the ether-based substituents, e.g., ethyl, i-propyl, n-butyl, t-butyl, phenyl, and oxyethylene groups, have been investigated by Tominaga et al.137 However, the electrochemical stability and +Li

t were found to be deteriorated with the involvement of ether-based side groups due to the lower stability of ethers toward electro-chemical oxidation and stronger chelation of lithium ion with ether oxygen compared with carbonyl groups.

By increasing the number of flexible methylene unit of PEC, the developed PTMC emerged as a promising SPE host candi-date due to its low Tg (≈ −15 °C) and amorphous nature (Entry 5 in Table 3). Smith et al. investigated the influence of various lithium salts, e.g., LiCF3SO3, LiClO4, LiBF4, LiSbF6, and LiPF6, on the ionic conduction of PTMC-based SPEs.64,65,144,145

Figure 7. (a) Tg and storage elastic modulus as a function of LiTFSI molar ratio versus monomer unit of PEC. Reprinted with permission.48 Copyright 2014, Elsevier B.V. (b) Ionic conductivity of PEC/LiTFSI SPE systems as a function of temperature and various lithium salt loadings. Reprinted with permission.124 Copyright 2016, Japan Science and Technology.

entry Polymer host lithium salt σ (S/cm) eSw +Lit References

21 LiTFSI 5.2 × 10−4 (20 °C) 4.6 V 0.75 166

σ: ionic conductivity; ESW: electrochemical stability window; LiCF3SO3: lithium trifluoromethanesulfonate; LiTFSI: lithium bis(trifluoromethanesulfonyl)imide; LiClO4: lithium perchlorate; LiFSI: lithium bis(fluorosulfonyl)amide; LiBF4: lithium tetrafluoroborate; LiPF6: lithium hexafluorophosphate; LiSbF6: lithium hexafluoroantimonate; LiDFOB: lithium difluoro(oxalate) borate.

Table 3. Continued

https://doi.org/10.1557/mre.2020.3Downloaded from https://www.cambridge.org/core. IP address: 54.39.106.173, on 12 Jan 2021 at 17:29:03, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.

Page 15: Carbonyl-coordinating polymers Hongli Xu Jingbing Xie ...€¦ · lithium batteries. Why carbonyl-coordinating polymers The dawn of the investigation on SPEs was from the work of

MRS eNeRgy & SUSTAINABIlITy // V O L U M E 7 // e 2 // www.mrs.org/energy-sustainability-journal n 15

The lithium salt–dependent ionic conductivity (<10−4 S/cm at 80 °C) and electrochemical stability (>4.5 V) are summarized in Entry 5 of Table 3. The PTMC-based SPE systems were also comprehensively studied using LiTFSI as the lithium salt by Brandell et al.63,146–148 The flexible PTMC/LiTFSI-integrated SPE delivered high electrochemical stability (>5 V) and high +Li

t (0.8 at 60 °C), with a moderate ionic conductivity (10−7 S/cm at 60 °C). Interestingly, even though the PTMC had a much lower Tg (−15 °C)63 than PEC (13 °C),124 the ionic conductivity did not outperform PEC. This is because of the lack of lithium salt plasticizing effect in the PTMC-based systems at high lithium salt loadings, which plays a decisive role in the high ionic con-ductivity of PEC/LiFSI or PEC/LiTFSI systems as mentioned above. The PTMC/LiTFSI systems behaved more like PEO/LiTFSI analogues, showing increased Tg with increasing LiTFSI concentration, and a maximum value of ionic conductivity at a moderate LiTFSI loading could be obtained.63 By further increasing the number of methylene units, Mecerreyes et al. reported the synthesis and properties of a series of polycar-bonates containing 4–12 methylene units by polycondensation of aliphatic diols and DMCs, showing low Tg of approximately −40 °C.149 High ionic conductivity (1 × 10−4 S/cm at room temperature) and high electrochemical stability (>4 V) were achieved for poly(dodecamethylene carbonate) at high LiTFSI concentration (80 wt%), as shown in Entry 6 of Table 3. By using the same polycondensation approach using diethylene glycol or triethylene glycol with DMCs, He et al. reported the synthesis of poly(carbonate-ether)s (PEDC and PTEC) incorporating with ether units, as shown in Entry 7 of Table 3.150 Compared with PEC, the polycarbonates with ether units showed significantly superior electrochemical properties, i.e., an optimized ionic con-ductivity of 1.1 × 10−5 S/cm at 25 °C, a decent +Li

t of 0.39, and a broad electrochemical window of approximately 4.5 V. The LiFePO4//SPE//Li cell with PTEC-based SPE was stably cycled at low C rates and at room temperature. The high-voltage cath-ode material of LiFe0.2Mn0.8PO4 (4.35 V) was also used, exhibit-ing decent rate capability and good cycling performance at ambient temperature using PTEC-based SPE. The rate capability and cycling performance of the LiFePO4//PTEC-LiTFSI//Li and LiFe0.2Mn0.8PO4//PTEC-LiTFSI//Li cells at 25 °C are displayed in Fig. 8. Through the in situ polymerization strategy, Liu et al. provided an interpenetrating network IPN-PDEC SPE using acrylate functionalized poly(diethylene glycol carbonate), shown in Entry 8 of Table 3.151 The as-prepared IPN-PDEC-LiTFSI exhibited a decent ionic conductivity of 1.6 × 10−4 S/cm at 25 °C and a high electrochemical oxidation voltage of 4.5 V. The LiFe0.2Mn0.8PO4//IPN-PDEC–LiDFOB//Li cells were capable of being cycled up to 4.35 V and showed a capacity retention of 96% (140 mAh/g) at 0.1 C after 100 cycles. Kim et al. also reported the electrochemical performance of a cross-linked poly(carbonate-ether)-based SPE room temperature for solid- state batteries using an in situ cross-linking strategy (Entry 9 of Table 3).152 The LiNi0.6Co0.2Mn0.2O2//PEEC-based SPE//Li cells delivered an ionic conductivity value of 4 × 10−5 S/cm at 25 °C and a broad electrochemical window (4.9 V), capable of being reversibly cycled at ambient temperature, delivering an

initial discharge capacity of 141.4 mAh/g and a good capacity retention with high Coulombic efficiency.

To further improve the ionic conductivity, polycarbonate–polyether block copolymer was investigated toward the one-component polymer hosts for high-performance SPEs, as summarized in Entries 10–16 of Table 3.138,153–157 Höcker et al. first reported the preparation of SPE based on poly (2,2-dimethyltrimethylene carbonate)-b-poly(ethylene oxide)-b- poly(2,2-dimethyltrimethylene carbonate) (PDTC-b-PEO-PDTC) triblock copolymer and LiCF3SO3, showing a moderate ionic conductivity value at the order of 10−4 S/cm at 60 °C.153 Tominaga et al. also reported the high ionic conductivity (4.8 × 10−4 S/cm at 60 °C) and high +Li

t (0.66) obtained from the ran-dom block copolymer of poly(ethylene carbonate/ethylene oxide)-based SPE.138 By cross-linking, the ionic conductivity of the poly(ethylene carbonate/ethylene oxide) was severely low-ered to an order of magnitude of 10−6 S/cm at 20 °C due to the hindered polymer chain segmental motions.154 FTIR experiment proved that the Li+ had no preference in chelating with carbonyl groups or ether units, which was in sharp contrast with the PEC systems containing oligoether side moieties.137 However, the electrochemical stability of the polycarbonates is most likely compromised from the induction of ether units. Single-ion con-ducting polymer electrolyte was prepared based on block copoly-mer of poly(carbonate-ether) by Deng et al., 157 with a high ionic conductivity of 1.6 × 10−4 S/cm at 80 °C and very high +Li

t (0.86). This work highlighted a facile, efficient, and eco-friendly way of preparing single ion–conducting polymer electrolyte with high ionic conductivity. A similar single ion–conducting polymer elec-trolytes from poly(ethylene oxide carbonate) was reported by David et al., 158 delivering a high ionic conductivity value of 1.2 × 10−4 S/cm at 70 °C and a high +Li

t (0.89).Apart from the polymer backbone design, several ways

have been explored in enhancing the ionic conductivity of polycarbonate-based SPEs. Soaking with liquid electrolytes together with preparing gel polymer electrolytes seems to be the most effective way in improving ionic conductivity, reaching an order of magnitude of 10−4 to 10−2 S/cm at 25 °C. Different polycarbonates (Entries 17–19 of Table 3) have been used as the polymer hosts, including polybutadiene/PPC interpenetrat-ing network,159 poly(vinylidene f luoride)/PPC,160 cellulose- supported poly(propylene carbonate),161 cross-linked poly (propylene carbonate maleate).155 However, such gel polymer electrolytes sacrificed the inherently high security and reliability of SPEs. Ionic liquid has long been regarded as an efficient addi-tive in plasticizing SPEs.162–164 Tominaga et al. reported the PEC-based composite SPE with ionic liquid (N-n-butyl-N- methylpyrrolidinium bis(trifluoromethanesulfonyl)imide, labe-led as Pyr14TFSI) and LiTFSI, showing a low ionic conductivity in the order of 10−7 S/cm at ambient temperature,131,132 which was the reason for its poor electrochemical performance in solid-state batteries. Much better electrochemical perfor-mance of the ionic liquid–containing SPE was reported by Lu et al., using PPC as the matrix, LiClO4 as the lithium salt, and 1-butyl-3-methylimidazolium tetrafluoroborate (BMIM+BF4

−) as the additive, as shown in Entry 20 of Table 3.165

https://doi.org/10.1557/mre.2020.3Downloaded from https://www.cambridge.org/core. IP address: 54.39.106.173, on 12 Jan 2021 at 17:29:03, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.

Page 16: Carbonyl-coordinating polymers Hongli Xu Jingbing Xie ...€¦ · lithium batteries. Why carbonyl-coordinating polymers The dawn of the investigation on SPEs was from the work of

16 n MRS eNeRgy & SUSTAINABIlITy // V O L U M E 7 // e 2 // www.mrs.org/energy-sustainability-journal

Dispersing inorganic fillers, e.g., TiO2,46 SiO2,132 LLZTO,166,167 and POSS,156 in polycarbonate-based SPEs has been regarded as an effective way to improve their electrochem-ical performance, especially the ionic conductivity (Entries 14 and 21 of Table 3). Traditionally, such enhancement is attrib-uted to the reduction of the crystallinity and lower Tg; however, the real case is more complex. A Lewis acid–base model of the interactions between inorganic fillers and polymer hosts was recently proposed, which provided an explanation of the strengthening effect of ionic conductivity from a unique per-spective.37 As an example, the fabrication of PEOEC/POSS hybrid semi-IPN SPEs is shown in Fig. 9.

Polyester-based SPes

Side chain polyester-based SPEs

The SPEs based on side chain–type polyesters with pendant ester moieties attached to the backbone have also been devel-oped, of which the majority are acrylates and methyl acrylates.

To date, most side chain polyester-based SPEs are in the realm of gel polymer electrolytes and are beyond the scope of this review.30,169 Florjańczyk et al. performed a systematic work on the SPEs with polymer matrices of poly(acrylonitrile- co-butyl acrylate), poly(methyl methacrylate), and poly(butyl acrylate).170–172 A considerably high ionic conductivity in the order of 10−4 S/cm at 30 °C was obtained for poly(butyl acrylate) containing 58 wt% LiI, due to its very low Tg (−55 °C) and the plasticizing effect of the concentrated lithium salt. Recently, a grafted copolymer composed of polyrotaxane backbone and PCL grafted side chains was utilized as SPE host,168 as shown in Fig. 10. The polyrotaxanes were prepared via self-assembly of cyclodextrin (CD) host molecules threading onto polyethyl-enoxide (PEO) chains, and the CD severed as the initiator for the ROP of CL to afford pendant PCL side chains with a few repeating units. Extraordinary ionic conductivity (10−3 S/cm at 60 °C and 10−4 S/cm at 25 °C), high +Li

t (0.6 at 60 °C), and superior electrochemical oxidative voltage (up to 4.7 V) were obtained for this SPE.

Figure 8. (a) The charge/discharge curves and (b) rate capability of LiFePO4//PTEC-LiTFSI//Li cells with varied C-rates at 25 °C. (c) The charge/discharge curves and (d) cycling performance of LiFe0.2Mn0.8PO4//PTEC-LiTFSI//Li cells at 25 °C. Reproduced with permission.150 Copyright 2017, Elsevier B.V.

https://doi.org/10.1557/mre.2020.3Downloaded from https://www.cambridge.org/core. IP address: 54.39.106.173, on 12 Jan 2021 at 17:29:03, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.

Page 17: Carbonyl-coordinating polymers Hongli Xu Jingbing Xie ...€¦ · lithium batteries. Why carbonyl-coordinating polymers The dawn of the investigation on SPEs was from the work of

MRS eNeRgy & SUSTAINABIlITy // V O L U M E 7 // e 2 // www.mrs.org/energy-sustainability-journal n 17

Main chain polyesters

Esters were used as electrolyte solvents mainly due to their high electrochemical stability against anodic decomposition at cathode surfaces.36 Sharing the similar skeleton of esters but in the polymeric form, main chain–type polyesters used in SPEs are generally obtained via the ROP of cyclic esters or polycon-densation of a diacid derivative and a diol, as discussed above. Similar to polycarbonates, FTIR confirmed that the coordina-tion and Li+ transport behavior in polyester-based SPEs primar-ily occurred between Li+ and carbonyl groups, rather than alcohol residue ester oxygen.173

Some typical polyester-based SPEs explored by pioneers are summarized in Table 4. To date, the well-studied main chain–type polyester-based SPE system is based on sustainable PCL, with various lithium salts like LiSCN 174,175 and LiClO4.176–178 The ion transport was rather limited due to its semicrystalline

nature, which resulted in low ionic conductivity (10−6 S/cm at room temperature, for both LiSCN and LiClO4) at low lithium salt loadings of 10–15 wt% (Entry 1 in Table 4). Similar systems of PPL/LiClO4 exhibited a moderate ionic conductivity of 3.5 × 10−5 S/cm at 70 °C.179 Surprisingly, quenching of the electro-lyte system increased the ionic conductivity by 1–3 orders of magnitude, achieving 3.7 × 10−4 S/cm, as shown in Entry 2 of Table 4. This is most likely due to the decrease in crystallinity by quenching, because mostly the lithium ion coordination and conduction occur in amorphous regions. The restricted lithium ion conductions were also reported in other semicrystalline polyester-based SPE systems with limited ionic conductivities (Entries 3 and 4 in Table 4), e.g., PEA/LiClO4 (1.06 × 10−5 S/cm at 30 °C),178 PBA/LiClO4 (9 × 10−7 S/cm at 30 °C),178 PHA/LiClO4 (5 × 10−6 S/cm at 30 °C),178 PE-2,4/LiClO4 (10−5 S/cm at 90 °C),180,181 PE-2,4/LiBF4 (3.4 × 10−6 S/cm at 65 °C),182 PE-2,4/LiSCN (4.9 × 10−7 S/cm at 87 °C),181 and PE-2,10/LiSCN (2.2 × 10−8 S/cm at 53 °C).181 It is clear that the ionic conductivity of an SPE is highly dependent on carbonyl groups distributions, Tg, polymer chain rigidity, crystallinity, electron donation strength of carbonyl group, lithium salt, etc. Shriver et al. also reported a moderate ionic conductivity (1.6 × 10−6 S/cm at 25 °C) of a noncrystalline polyester-based SPE system of PEM/LiCF3SO3,183 as shown in Entry 5 of Table 4. All the inves-tigations were performed at low lithium salt concentrations (≈10 wt%). As discussed above, high-concentration imide-type lithium salts, e.g., LiFSI and LiTFSI, have significant plasticiz-ing effect and potentially lower Tg, achieving higher ionic con-ductivity for polycarbonate-based SPE systems.124 Considering the similarity in the skeletons of these two polymers, it is rea-sonable to assume that such plasticizing mechanism is also valid for polyesters with imide-type lithium salt systems.

Better electrochemical performance was reported for two series of systematically studied ether side chain–modified polyester-based SPE systems, as summarized in Table 4. By dop-ing with LiTFSI (Entry 6 in Table 4), the ionic conductivity fell in the range of 1.4 × 10−4 to 3.0 × 10−4 S/cm at 90 °C, except

Figure 9. Fabrication of PEOEC/POSS hybrid semi-IPN SPEs. Reproduced with permission.156 Copyright 2014, Elsevier B.V.

Figure 10. (a) Grafted polyrotaxane structure with cyclodextrin in blue, PEO backbone in green, and PCL side chains in orange. (b) General synthesis scheme for the synthesis of GPR with (1) functionalization of PEO ends, (2) complexation with CD, and (3) end capping and simultaneous grafting. (c) ROP of ε-caprolactone initiated by CD. Reprinted with permission.168 Copyright 2019, Elsevier B.V.

https://doi.org/10.1557/mre.2020.3Downloaded from https://www.cambridge.org/core. IP address: 54.39.106.173, on 12 Jan 2021 at 17:29:03, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.

Page 18: Carbonyl-coordinating polymers Hongli Xu Jingbing Xie ...€¦ · lithium batteries. Why carbonyl-coordinating polymers The dawn of the investigation on SPEs was from the work of

18 n MRS eNeRgy & SUSTAINABIlITy // V O L U M E 7 // e 2 // www.mrs.org/energy-sustainability-journal

Table 4. Molecular structures and electrochemical characteristics of side chain polyester-based SPEs.

entry Polymer host lithium salt σ (S/cm) eSw +Lit References

1 LiSCN 1.0 × 10−6 (25 °C) ≈5 V (25 °C) N/A 174–178

LiClO4 1.2 × 10−6 (25 °C)

2 LiClO4 3.7 × 10−4 (70 °C) N/A 0.82 179

3 LiClO4 PEA: 1.06 × 10−5 (30 °C) N/A N/A 178

PBA: 1.3 × 10−6 (30 °C)

PHA: 9.5 × 10−7 (30 °C)

4 LiClO4 PE-2,4: 10−5 (90 °C) N/A N/A 180–182

LiBF4 PE-2,4: 3.4 × 10−5 (65 °C)

LiSCN PE-2,4: 4.9×10−7 (87 °C), PE-2,10: 2.2×10−8 (53 °C)

5 LiCF3SO3 1.6 × 10−6 (25 °C) N/A N/A 183

6 LiTFSI 1a: 1.5 × 10−4 (90 °C); 2a: 1.4 × 10−4 (90 °C)

N/A N/A 184 and 185

3a: 3.0 × 10−4 (90 °C); 1b: 3.8 × 10−4 (90 °C)

2b: 2.5 × 10−4 (90 °C); 3b: 2.1 × 10−4 (90 °C)

σ: ionic conductivity; ESW: electrochemical stability window; LiSCN: lithium thiocyanate; LiClO4: lithium perchlorate; LiBF4: lithium tetrafluoroborate; LiCF3SO3: lithium trifluoromethanesulfonate; LiTFSI: lithium bis(trifluoromethanesulfonyl)imide.

https://doi.org/10.1557/mre.2020.3Downloaded from https://www.cambridge.org/core. IP address: 54.39.106.173, on 12 Jan 2021 at 17:29:03, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.

Page 19: Carbonyl-coordinating polymers Hongli Xu Jingbing Xie ...€¦ · lithium batteries. Why carbonyl-coordinating polymers The dawn of the investigation on SPEs was from the work of

MRS eNeRgy & SUSTAINABIlITy // V O L U M E 7 // e 2 // www.mrs.org/energy-sustainability-journal n 19

that of 1b polymer was approximately one order of magnitude lower (10−5 S/cm) owing to its higher Tg and slower segmental motion.184,185 Most of the SPEs showed low mechanical proper-ties at room temperature, which was due to their low molecular weight (<10,000 g/mol). By using the long-timescale molecular dynamics (MD) simulations, the lithium ion solvation and diffu-sion mechanisms in the above-mentioned polymers (as well as PEO) were analyzed, showing that the intrachain hopping of Li+ took place in PEO only. The Li+ transport in the polyesters pri-marily depended on the interchain hopping behavior, which was intrinsically slow.

SPes from polyester/polycarbonate copolymersAs a sustainable material, biodegradable PCL has been stud-

ied thoroughly as SPE host for high–energy density solid-state lithium batteries. However, the PCL-based SPEs show restricted lithium ion diffusion (with an ionic conductivity in an order of 10−6 S/cm at room temperature) due to the semicrystalline nature. In polymer science, preparing block copolymers with two distinct repeating units has been regarded as an effective strategy to reduce the chain regular arrangement and thus to suppress crystallinity.186 To this end, the PCL backbone with randomly distributed carbonate repeating units (TMC) was prepared by Mindemark et al. via bulk ROP at a TMC/CL ratio of 2:8.58,148 The as-prepared PTMC/PCL block copolymer-based SPE deliv-ered a high ionic conductivity of 4.1 × 10−5 S/cm at 25 °C and a high +Li

t (>0.6). The corresponding cell could be operated at tem-peratures down to as low as 23 °C, dramatically surpassing their polyether-based SPE analogues. The rate capability of the cells using the PTMC/PCL block copolymer–based SPE is presented in Fig. 11. The enhanced ionic conductivity was attributed to the higher lithium ion mobilities by the random copolymerization with carbonate units in polyesters. FTIR experiments indicated that the Li+ preferentially coordinated with ester carbonyl group rather than with carbonate carbonyl moiety, which was in

consistency with the structure–dynamics model provided by MD simulations. However, by increasing the TMC content to 60%, the highest ionic conductivity achievable was significantly depressed, with a value of 7.9 × 10−7 S/cm at 25 °C, despite the totally amorphous nature.187 The reduced lithium ion mobility further confirmed the FTIR results of the preferential coordina-tion of Li+ with ester carbonyls.148

Summary and perspectiveAs one of the most successful energy storage systems in his-

tory, liquid electrolyte–based lithium ion batteries have been used massively, despite their inherent limitations of high vola-tilities and flammability. To realize the application of lithium ion batteries with high energy density, low cost, and excellent safety, using solid-state lithium batteries with SPEs seems to be promising and probably an inevitable choice. Taking the inher-ent advantages of SPEs, solid-state lithium batteries can allevi-ate the low energy density and safety concerns and endow the energy storage system with improved f lexibility and process-ability. Given the lower donor number and weaker complexa-tion strength of Li+ with carbonyl groups in polycarbonates and polyesters than with ether units in polyethers, the carbonyl- coordinating polymers deliver superior electrochemical proper-ties with respect to ionic conductivity and +Li

t . A unique property regarding the ionic conductivity in carbonyl-coordinating poly-mers is that the maximum value is generally obtained at high lithium salt concentrations due to the plasticizing effect, while the high lithium salt loading in polyethers ordinarily causes inter-/intrachain cross-linking and polymer chain stiffening. Besides, considering the electrochemical oxidation mechanism of SPEs with high-voltage cathode materials, the carbonyl- coordinating polymers exhibit broader ESW compared with their polyether-based analogues.

Carbonyl-coordinating polymer–based SPEs have emerged as a promising class of material for electrochemical applications, in

Figure 11. (a) Rate performance and coulombic efficiencies for a Li/SPE/LiFePO4 cell at room temperature and (b) initial charge/discharge voltage curves cycled at different temperatures and C-rates (right). Reprinted with permission.58 Copyright 2015, Elsevier B.V.

https://doi.org/10.1557/mre.2020.3Downloaded from https://www.cambridge.org/core. IP address: 54.39.106.173, on 12 Jan 2021 at 17:29:03, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.

Page 20: Carbonyl-coordinating polymers Hongli Xu Jingbing Xie ...€¦ · lithium batteries. Why carbonyl-coordinating polymers The dawn of the investigation on SPEs was from the work of

20 n MRS eNeRgy & SUSTAINABIlITy // V O L U M E 7 // e 2 // www.mrs.org/energy-sustainability-journal

the sense that this type of material exhibits several major advan-tages over the polyether-based analogues, including higher ionic conductivity, higher +Li

t and wider electrochemical win-dow. However, such superior electrochemical properties are generally obtained at high lithium salt concentrations or in polymer-in-salt systems, arousing a series of challenges. For example, how can we solve the contradiction between ionic con-ductivity and mechanical strength? Is there any difference in terms of the lithium ion conduction mechanism between salt-in-polymer and polymer-in-salt systems? What is the spe-cial role of polymer dynamics at such high salt concentrations? What is the influence of high lithium salt concentrations on the ESW? How can we further improve the ESW of carbonyl- coordinating SPEs to higher than 4.5 V or even 5 V? What is the electrochemical decomposition mechanism of the carbonyl- coordinating SPEs at high lithium salt concentrations and high voltage (>4.5 V)? How can we improve the interfacial sta-bility of SPE/cathode or SPE/anode at such high lithium salt concentrations and very high voltage? All these questions remain mystery, and much more efforts are still needed to fur-ther study and refine the carbonyl-coordinating polymer–based SPEs for high-voltage electrochemical applications.

In this review, we summarized and highlighted the synthesis, sustainability, and ionic conduction of carbonyl-coordinating polymers and the SPEs thereof. To the best of our knowledge and our understanding on SPEs, we tentatively suggest the fol-lowing challenges and perspective:

(1) Further improvement of ionic conductivity. The trans-

port of lithium ions highly depends on the lithium ion dynamic complexation with the polymer chains and the segmental motions. Given this mechanism, the ionic conductivity of carbonyl-coordinating polymer–based SPEs is possible to be enhanced. Strategies on improv-ing the ionic conductivity of polyether-based SPEs have been explored extensively, including the fabrication of organic/inorganic hybrid solid electrolyte, advanced polymer architecture modification, and incorporation of novel lithium salts, which are seldom reported for car-bonyl-coordinating SPEs.

(2) Balance between mechanical strength and ionic con-duction. As the Li+ transport is highly relied on the pol-ymer chain segmental motions, while the mechanical strength is in the retro-correlation with the segmental movements. To tackle this predicament, preparing block copolymers containing both the hard block and the soft ion conductive block seems to a reasonable strategy. The self-assembled nanostructures composed of the hard block and soft block enable the SPE thereof with both satisfied mechanical strength and unacted high ionic conductivity. As a bonus, the lithium den-drite growth problem is solved by this approach. Taking the advantage of the interaction between polymer and other additives in SPE, e.g., plastic crystal and inor-ganic particles, the ionic conductivity is improved by enhancing the lithium salt dissociation and suppression

of the polymer crystallinity without sacrificing the mechanical strength.

(3) Appropriate evaluation of cathodic stability with high- voltage cathode materials. Despite most of the carbonyl- coordinating polymers show satisfying results from linear sweep and cyclic voltammetry measurements, very few work investigated the long-term cycling and electro(chemical) decomposition of such SPEs in con-tact with high-voltage cathode materials (such as LiMn0.5Fe0.5PO4, lithium nickel cobalt manganese oxide (NCM), or lithium nickel cobalt aluminium oxide (NCA)), not to mention the building up of the practical energy stor-age devices. A technique that can simulate the realistic high-voltage lithium battery internal environment should be developed to evaluate the realistic cathodic stability of the SPEs. It is noteworthy that most of the high-voltage cathode materials, e.g., NCM and NCA, are strong base, which may cause base-catalyzed degradation on the car-bonyl-coordinating polymers. Tailoring the SPE/cathode interface or coating the cathode active materials with high chemical stability layers should be helpful.

(4) Amelioration of the lithium metal compatibility. In theory, a healthy battery requires both the good anodic and the cathodic stabilities with SPE. Due to the extremely high reactivity, the chemical and elec-trochemical stability and compatibility of carbonyl- coordinating polymer–based SPEs with lithium metal anode should be accessed comprehensively. Prepar-ing a dual-layered SPE constructed with both anodic electrolyte (such as polyethers) and cathodic electrolyte with high cathodic stability (such as polycarbonates) should be promising.

(5) Adopting advanced characterization technologies. The interfacial properties of the SPE with electrodes are critical for the healthy long-term cycling of a lithium battery. Using advanced in situ techniques, such as in situ FTIR, Raman, XRD, and cryo-TEM, to study the interfacial layer formation, development and composi-tions, and lithium dendrite growth during long-term cycling is meaningful.

(6) Recycling of lithium battery materials. With the current massive application of lithium batteries in electrical vehicles, millions of tons of lithium battery will be retired in the coming several years, and the recycling of lithium battery material is increasingly urgent. The recycling of lithium battery is seriously challenging due to the versatile battery components and rich chemistry. While most attentions on battery material recycling are focused on valuable anodic and cathodic materials, the liquid electrolytes in current lithium ion batteries con-tain hazardous organic materials and need to be treated very carefully. Developing the alternative solid electro-lytes for lithium batteries from bio-based and biodegrad-able polymers emerges as an important and attractive strategy in terms of environmental benignity, lower cost, and higher operability.

https://doi.org/10.1557/mre.2020.3Downloaded from https://www.cambridge.org/core. IP address: 54.39.106.173, on 12 Jan 2021 at 17:29:03, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.

Page 21: Carbonyl-coordinating polymers Hongli Xu Jingbing Xie ...€¦ · lithium batteries. Why carbonyl-coordinating polymers The dawn of the investigation on SPEs was from the work of

MRS eNeRgy & SUSTAINABIlITy // V O L U M E 7 // e 2 // www.mrs.org/energy-sustainability-journal n 21

AcknowledgmentsThe authors gratefully acknowledge the financial support

from National Key R&D Program of China (2018YFB0104300), Shenzhen Key Laboratory of Solid State Batteries (ZDSYS201802081843465), and Guangdong Provincial Key Laboratory of Energy Materials for Electric Power (2018B030322001).

ReFeReNCeS:

1. Mohanty D., Li J., Nagpure S.C., Wood D.L., and Daniel C.: Understanding the structure and structural degradation mechanisms in high-voltage, lithium-manganese–rich lithium-ion battery cathode oxides: A review of materials diagnostics. MRS Energy Sustainability 2, E15 (2015).

2. US Energy Information Administration (2018). Available at: http://www.eia.gov/ (accessed February 2020).

3. Lewis N.S.: Powering the planet. MRS Bull. 32, 808 (2007). 4. Goodenough J.B. and Kim Y.: Challenges for rechargeable Li batteries.

Chem. Mater. 22, 587 (2010). 5. Bresser D., Hosoi K., Howell D., Li H., Zeisel H., Amine K., and Passerini S.:

Perspectives of automotive battery R&D in China, Germany, Japan, and the USA. J. Power Sources 382, 176 (2018).

6. Tarascon J.M. and Armand M.: Issues and challenges facing rechargeable lithium batteries. Nature 414, 359 (2001).

7. Quartarone E. and Mustarelli P.: Electrolytes for solid-state lithium rechargeable batteries: Recent advances and perspectives. Chem. Soc. Rev. 40, 2525 (2011).

8. Goodenough J.B. and Park K.-S.: The Li-ion rechargeable battery: A perspective. J. Am. Chem. Soc. 135, 1167 (2013).

9. Goodenough J.B. and Singh P.: Review—Solid electrolytes in rechargeable electrochemical cells. J. Electrochem. Soc. 162, A2387 (2015).

10. Goodenough J.B.: How we made the Li-ion rechargeable battery. Nat. Electron. 1, 204 (2018).

11. Manthiram A., Yu X., and Wang S.: Lithium battery chemistries enabled by solid-state electrolytes. Nat. Rev. Mater. 2, 16103 (2017).

12. Osada I., de Vries H., Scrosati B., and Passerini S.: Ionic-liquid-based polymer electrolytes for battery applications. Angew. Chem., Int. Ed. 55, 500 (2016).

13. Sarabi S., Kefsi L., Merdassi A., and Robyns B.: Supervision of plug-in electric vehicles connected to the electric distribution grids. Int. J. Electr. Energy 1, 256 (2013).

14. Wright P.V.: Electrical conductivity in ionic complexes of poly(ethylene oxide). Br. Polym. J. 7, 319 (1975).

15. Armand M.: Polymer solid electrolytes—An overview. Solid State Ionics 9–10, 745 (1983).

16. Wang C., Zhang H., Li J., Chai J., Dong S., and Cui G.: The interfacial evolution between polycarbonate-based polymer electrolyte and Li-metal anode. J. Power Sources 397, 157 (2018).

17. Fish D. and Smid J.: Solvation of lithium ions in mixtures of tetraethylene glycol dimethyl ether and propylene carbonate. Electrochim. Acta 37, 2043 (1992).

18. Zhang C., Ueno K., Yamazaki A., Yoshida K., Moon H., Mandai T., Umebayashi Y., Dokko K., and Watanabe M.: Chelate effects in glyme/lithium bis(trif luoromethanesulfonyl)amide solvate ionic liquids. I. Stability of solvate cations and correlation with electrolyte properties. J. Phys. Chem. B 118, 5144 (2014).

19. Wu J., Rao Z., Cheng Z., Yuan L., Li Z., and Huang Y.: Ultrathin, flexible polymer electrolyte for cost-effective fabrication of all-solid-state lithium metal batteries. Adv. Energy Mater. 9, 1902767 (2019).

20. Xu C., Sun B., Gustafsson T., Edström K., Brandell D., and Hahlin M.: Interface layer formation in solid polymer electrolyte lithium batteries: An XPS study. J. Mater. Chem. A 2, 7256 (2014).

21. Wei Z., Chen S., Wang J., Wang Z., Zhang Z., Yao X., Deng Y., and Xu X.: Superior lithium ion conduction of polymer electrolyte with comb-like structure via solvent-free copolymerization for bipolar all-solid-state lithium battery. J. Mater. Chem. A 6, 13438 (2018).

22. Di Noto V., Lavina S., Giffin G.A., Negro E., and Scrosati B.: Polymer electrolytes: Present, past and future. Electrochim. Acta 57, 4 (2011).

23. Meyer W.H.: Polymer electrolytes for lithium-ion batteries. Adv. Mater. 10, 439 (1998).

24. Agrawal R.C. and Pandey G.P.: Solid polymer electrolytes: Materials designing and all-solid-state battery applications: An overview. J. Phys. D: Appl. Phys. 41, 223001 (2008).

25. Fergus J.W.: Ceramic and polymeric solid electrolytes for lithium-ion batteries. J. Power Sources 195, 4554 (2010).

26. Hallinan D.T. and Balsara N.P.: Polymer electrolytes. Annu. Rev. Mater. Res. 43, 503 (2013).

27. Xue Z., He D., and Xie X.: Poly(ethylene oxide)-based electrolytes for lithium-ion batteries. J. Mater. Chem. A 3, 19218 (2015).

28. Mindemark J., Lacey M.J., Bowden T., and Brandell D.: Beyond PEO—Alternative host materials for Li+-conducting solid polymer electrolytes. Prog. Polym. Sci. 81, 114 (2018).

29. Zhang J., Yang J., Dong T., Zhang M., Chai J., Dong S., Wu T., Zhou X., and Cui G.: Aliphatic polycarbonate-based solid-state polymer electrolytes for advanced lithium batteries: Advances and perspective. Small 14, 1800821 (2018).

30. Manuel Stephan A.: Review on gel polymer electrolytes for lithium batteries. Eur. Polym. J. 42, 21 (2006).

31. Xu K.: Nonaqueous liquid electrolytes for lithium-based rechargeable batteries. Chem. Rev. 104, 4303 (2004).

32. Druger S.D., Nitzan A., and Ratner M.A.: Dynamic bond percolation theory: A microscopic model for diffusion in dynamically disordered systems. I. Definition and one-dimensional case. J. Chem. Phys. 79, 3133 (1983).

33. Webb M.A., Savoie B.M., Wang Z.-G., and Miller T.F. III: Chemically specific dynamic bond percolation model for ion transport in polymer electrolytes. Macromolecules 48, 7346 (2015).

34. Song J.Y., Wang Y.Y., and Wan C.C.: Review of gel-type polymer electrolytes for lithium-ion batteries. J. Power Sources 77, 183 (1999).

35. Wang C., Wang T., Wang L., Hu Z., Cui Z., Li J., Dong S., Zhou X., and Cui G.: Differentiated lithium salt design for multilayered PEO electrolyte enables a high-voltage solid-state lithium metal battery. Adv. Sci. 6, 1901036 (2019).

36. Xu K.: Electrolytes and interphases in Li-ion batteries and beyond. Chem. Rev. 114, 11503 (2014).

37. Zhou Q., Ma J., Dong S., Li X., and Cui G.: Intermolecular chemistry in solid polymer electrolytes for high-energy-density lithium batteries. Adv. Mater. 31, e1902029 (2019).

38. Zaheer M., Xu H., Wang B., Li L., and Deng Y.: An in situ polymerized comb-like PLA/PEG-based solid polymer electrolyte for lithium metal batteries. J. Electrochem. Soc. 167, 070504 (2020).

39. Takahashi Y. and Tadokoro H.: Structural studies of polyethers, (–(CH2)m–O–)n. X. Crystal structure of poly(ethylene oxide). Macromolecules 6, 672 (1973).

40. Gadjourova Z., Andreev Y.G., Tunstall D.P., and Bruce P.G.: Ionic conductivity in crystalline polymer electrolytes. Nature 412, 520 (2001).

41. Cheng S., Smith D.M., and Li C.Y.: How does nanoscale crystalline structure affect ion transport in solid polymer electrolytes? Macromolecules 47, 3978 (2014).

42. Zhou Q., Zhang J., and Cui G.: Rigid–flexible coupling polymer electrolytes toward high-energy lithium batteries. Macromol. Mater. Eng. 303, 1800337 (2018).

43. Matsubara K., Kaneuchi R., and Maekita N.: 13C NMR estimation of preferential solvation of lithium ions in non-aqueous mixed solvents. J. Chem. Soc., Faraday Trans. 94, 3601 (1998).

44. Ong M.T., Verners O., Draeger E.W., van Duin A.C.T., Lordi V., and Pask J.E.: Lithium ion solvation and diffusion in bulk organic electrolytes from first-principles and classical reactive molecular dynamics. J. Phys. Chem. B 119, 1535 (2015).

45. Bogle X., Vazquez R., Greenbaum S., Cresce A.v.W., and Xu K.: Understanding Li+–solvent interaction in nonaqueous carbonate electrolytes with 17O NMR. J. Phys. Chem. Lett. 4, 1664 (2013).

https://doi.org/10.1557/mre.2020.3Downloaded from https://www.cambridge.org/core. IP address: 54.39.106.173, on 12 Jan 2021 at 17:29:03, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.

Page 22: Carbonyl-coordinating polymers Hongli Xu Jingbing Xie ...€¦ · lithium batteries. Why carbonyl-coordinating polymers The dawn of the investigation on SPEs was from the work of

22 n MRS eNeRgy & SUSTAINABIlITy // V O L U M E 7 // e 2 // www.mrs.org/energy-sustainability-journal

46. Tominaga Y. and Yamazaki K.: Fast Li-ion conduction in poly(ethylene carbonate)-based electrolytes and composites filled with TiO2 nanoparticles. Chem. Commun. 50, 4448 (2014).

47. Kimura K., Motomatsu J., and Tominaga Y.: Correlation between solvation structure and ion-conductive behavior of concentrated poly(ethylene carbonate)-based electrolytes. J. Phys. Chem. C 120, 12385 (2016).

48. Okumura T. and Nishimura S.: Lithium ion conductive properties of aliphatic polycarbonate. Solid State Ionics 267, 68 (2014).

49. Doyle M., Fuller T.F., and Newman J.: The importance of the lithium ion transference number in lithium/polymer cells. Electrochim. Acta 39, 2073 (1994).

50. Thomas K.E., Sloop S.E., Kerr J.B., and Newman J.: Comparison of lithium-polymer cell performance with unity and nonunity transference numbers. J. Power Sources 89, 132 (2000).

51. Doyle M. and Newman J.: The use of mathematical modeling in the design of lithium/polymer battery systems. Electrochim. Acta 40, 2191 (1995).

52. Brissot C., Rosso M., Chazalviel J.N., and Lascaud S.: Dendritic growth mechanisms in lithium/polymer cells. J. Power Sources 81–82, 925 (1999).

53. Gorecki W., Jeannin M., Belorizky E., Roux C., and Armand M.: Physical properties of solid polymer electrolyte PEO(LiTFSI) complexes. J. Phys.: Condens. Matter 7, 6823 (1995).

54. Borodin O. and Smith G.D.: Mechanism of ion transport in amorphous poly(ethylene oxide)/LiTFSI from molecular dynamics simulations. Macromolecules 39, 1620 (2006).

55. Mao G., Saboungi M.-L., Price D.L., Armand M.B., and Howells W.S.: Structure of liquid PEO-LiTFSI electrolyte. Phys. Rev. Lett. 84, 5536 (2000).

56. Kim C.S. and Oh S.M.: Importance of donor number in determining solvating ability of polymers and transport properties in gel-type polymer electrolytes. Electrochim. Acta 45, 2101 (2000).

57. Chen L., Venkatram S., Kim C., Batra R., Chandrasekaran A., and Ramprasad R.: Electrochemical stability window of polymeric electrolytes. Chem. Mater. 31, 4598 (2019).

58. Mindemark J., Sun B., Törmä E., and Brandell D.: High-performance solid polymer electrolytes for lithium batteries operational at ambient temperature. J. Power Sources 298, 166 (2015).

59. Manuel Stephan A. and Nahm K.S.: Review on composite polymer electrolytes for lithium batteries. Polymer 47, 5952 (2006).

60. Sun C., Liu J., Gong Y., Wilkinson D.P., and Zhang J.: Recent advances in all-solid-state rechargeable lithium batteries. Nano Energy 33, 363 (2017).

61. Cheng X.B., Hou T.Z., Zhang R., Peng H.J., Zhao C.Z., Huang J.Q., and Zhang Q.: Dendrite-free lithium deposition induced by uniformly distributed lithium ions for efficient lithium metal batteries. Adv. Mater. 28, 2888 (2016).

62. Yang Q., Li W., Dong C., Ma Y., Yin Y., Wu Q., Xu Z., Ma W., Fan C., and Sun K.: PIM-1 as an artificial solid electrolyte interphase for stable lithium metal anode in high-performance batteries. J. Energy Chem. 42, 83 (2020).

63. Sun B., Mindemark J., Edstrom K., and Brandell D.: Polycarbonate-based solid polymer electrolytes for Li-ion batteries. Solid State Ionics 262, 738 (2014).

64. Silva M.M., Barros S.C., Smith M.J., and MacCallum J.R.: Characterization of solid polymer electrolytes based on poly(trimethylenecarbonate) and lithium tetrafluoroborate. Electrochim. Acta 49, 1887 (2004).

65. Barbosa P.C., Rodrigues L.C., Silva M.M., and Smith M.J.: Characterization of pTMCnLiPF6 solid polymer electrolytes. Solid State Ionics 193, 39 (2011).

66. Kobayashi S.: Enzymatic ring-opening polymerization and polycondensation for the green synthesis of polyesters. Polym. Adv. Technol. 26, 677 (2015).

67. Artham T. and Doble M.: Biodegradation of aliphatic and aromatic polycarbonates. Macromol. Biosci. 8, 14 (2008).

68. Cameron D.J.A. and Shaver M.P.: Aliphatic polyester polymer stars: Synthesis, properties and applications in biomedicine and nanotechnology. Chem. Soc. Rev. 40, 1761 (2011).

69. Brannigan R.P. and Dove A.P.: Synthesis, properties and biomedical applications of hydrolytically degradable materials based on aliphatic polyesters and polycarbonates. Biomater. Sci. 5, 9 (2017).

70. Dai Y. and Zhang X.: Recent development of functional aliphatic polycarbonates for the construction of amphiphilic polymers. Polym. Chem. 8, 7429 (2017).

71. Hussain T., Tausif M., and Ashraf M.: A review of progress in the dyeing of eco-friendly aliphatic polyester-based polylactic acid fabrics. J. Clean. Prod. 108, 476 (2015).

72. Yu Y., Wu D., Liu C., Zhao Z., Yang Y., and Li Q.: Lipase/esterase-catalyzed synthesis of aliphatic polyesters via polycondensation: A review. Process Biochem. 47, 1027 (2012).

73. Malmstroem E., Johansson M., and Hult A.: Hyperbranched aliphatic polyesters. Macromolecules 28, 1698 (1995).

74. Undin J., Plikk P., Finne-Wistrand A., and Albertsson A.-C.: Synthesis of amorphous aliphatic polyester-ether homo- and copolymers by radical polymerization of ketene acetals. J. Polym. Sci., Part A: Polym. Chem. 48, 4965 (2010).

75. Mehta R., Kumar V., Bhunia H., and Upadhyay S.N.: Synthesis of poly(lactic acid): A review. J. Macromol. Sci. Part C: Polym. Rev. 45, 325 (2005).

76. Jiang Z.: Lipase-catalyzed synthesis of aliphatic polyesters via copolymerization of lactone, dialkyl diester, and diol. Biomacromolecules 9, 3246 (2008).

77. Varma I.K., Albertsson A.-C., Rajkhowa R., and Srivastava R.K.: Enzyme catalyzed synthesis of polyesters. Prog. Polym. Sci. 30, 949 (2005).

78. Zhang J., Shi H., Wu D., Xing Z., Zhang A., Yang Y., and Li Q.: Recent developments in lipase-catalyzed synthesis of polymeric materials. Process Biochem. 49, 797 (2014).

79. Douka A., Vouyiouka S., Papaspyridi L.-M., and Papaspyrides C.D.: A review on enzymatic polymerization to produce polycondensation polymers: The case of aliphatic polyesters, polyamides and polyesteramides. Prog. Polym. Sci. 79, 1 (2018).

80. Williams C.K.: Synthesis of functionalized biodegradable polyesters. Chem. Soc. Rev. 36, 1573 (2007).

81. Liu Z.-L., Zhou Y., and Zhuo R.-X.: Synthesis and properties of functional aliphatic polycarbonates. J. Polym. Sci., Part A: Polym. Chem. 41, 4001 (2003).

82. Wang X.-L., Zhuo R.-X., Liu L.-J., He F., and Liu G.: Synthesis and characterization of novel aliphatic polycarbonates. J. Polym. Sci., Part A: Polym. Chem. 40, 70 (2002).

83. Tempelaar S., Mespouille L., Coulembier O., Dubois P., and Dove A.P.: Synthesis and post-polymerisation modifications of aliphatic poly(carbonate)s prepared by ring-opening polymerisation. Chem. Soc. Rev. 42, 1312 (2013).

84. Gross R., Kalra B., and Kumar A.: Polyester and polycarbonate synthesis by in vitro enzyme catalysis. Appl. Microbiol. Biotechnol. 55, 655 (2001).

85. Taherimehr M. and Pescarmona P.P.: Green polycarbonates prepared by the copolymerization of CO2 with epoxides. J. Appl. Polym. Sci. 131, 41141 (2014).

86. Tamura M., Ito K., Honda M., Nakagawa Y., Sugimoto H., and Tomishige K.: Direct copolymerization of CO2 and diols. Sci. Rep. 6, 24038 (2016).

87. Carothers W.H., Dorough G.L., and Natta F.J.v.: Studies of polymerization and ring formation. X. The reversible polymerization of six-membered cyclic esters. J. Am. Chem. Soc. 54, 761 (1932).

88. Bendler J.T.: Handbook of Polycarbonate Science and Technology, 1st ed. (CRC Press, New York, 1999).

89. Zhu W., Huang X., Li C., Xiao Y., Zhang D., and Guan G.: High-molecular-weight aliphatic polycarbonates by melt polycondensation of dimethyl carbonate and aliphatic diols: Synthesis and characterization. Polym. Int. 60, 1060 (2011).

90. Park J.H., Jeon J.Y., Lee J.J., Jang Y., Varghese J.K., and Lee B.Y.: Preparation of high-molecular-weight aliphatic polycarbonates by condensation polymerization of diols and dimethyl carbonate. Macromolecules 46, 3301 (2013).

91. Mespouille L., Coulembier O., Kawalec M., Dove A.P., and Dubois P.: Implementation of metal-free ring-opening polymerization in the preparation of aliphatic polycarbonate materials. Prog. Polym. Sci. 39, 1144 (2014).

https://doi.org/10.1557/mre.2020.3Downloaded from https://www.cambridge.org/core. IP address: 54.39.106.173, on 12 Jan 2021 at 17:29:03, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.

Page 23: Carbonyl-coordinating polymers Hongli Xu Jingbing Xie ...€¦ · lithium batteries. Why carbonyl-coordinating polymers The dawn of the investigation on SPEs was from the work of

MRS eNeRgy & SUSTAINABIlITy // V O L U M E 7 // e 2 // www.mrs.org/energy-sustainability-journal n 23

92. Möller M., Hedrick J.L., Degée P., and Dubois P.: Ring opening polymerization. In Encyclopedia of Materials: Science and Technology, Buschow K.H.J., Cahn R.W., Flemings M.C., Ilschner B., Kramer E.J., Mahajan S. and Veyssière P., eds. (Elsevier, Oxford, 2001); p. 8202.

93. Jérôme C. and Lecomte P.: Recent advances in the synthesis of aliphatic polyesters by ring-opening polymerization. Adv. Drug Delivery Rev. 60, 1056 (2008).

94. Paul S., Zhu Y., Romain C., Brooks R., Saini P.K., and Williams C.K.: Ring-opening copolymerization (ROCOP): Synthesis and properties of polyesters and polycarbonates. Chem. Commun. 51, 6459 (2015).

95. Xu J., Feng E., and Song J.: Renaissance of aliphatic polycarbonates: New techniques and biomedical applications. J. Appl. Polym. Sci. 131, 39822 (2014).

96. D’Alessandro D.M., Smit B., and Long J.R.: Carbon dioxide capture: Prospects for new materials. Angew. Chem., Int. Ed. 49, 6058 (2010).

97. Fukuoka S., Kawamura M., Komiya K., Tojo M., Hachiya H., Hasegawa K., Aminaka M., Okamoto H., Fukawa I., and Konno S.: A novel non-phosgene polycarbonate production process using by-product CO2 as starting material. Green Chem. 5, 497 (2003).

98. Darensbourg D.J., Mackiewicz R.M., Phelps A.L., and Billodeaux D.R.: Copolymerization of CO2 and epoxides catalyzed by metal salen complexes. Acc. Chem. Res. 37, 836 (2004).

99. Sugimoto H. and Inoue S.: Copolymerization of carbon dioxide and epoxide. J. Polym. Sci., Part A: Polym. Chem. 42, 5561 (2004).

100. Inoue S., Koinuma H., and Tsuruta T.: Copolymerization of carbon dioxide and epoxide. J. Polym. Sci., Part B: Polym. Lett. 7, 287 (1969).

101. Lu X.-B., Ren W.-M., and Wu G.-P.: CO2 copolymers from epoxides: Catalyst activity, product selectivity, and stereochemistry control. Acc. Chem. Res. 45, 1721 (2012).

102. Coates G.W. and Moore D.R.: Discrete metal-based catalysts for the copolymerization of CO2 and epoxides: Discovery, reactivity, optimization, and mechanism. Angew. Chem., Int. Ed. 43, 6618 (2004).

103. Albertsson A.-C. and Varma I.K.: Aliphatic Polyesters: Synthesis, Properties and Applications, Degradable Aliphatic Polyesters (Springer Berlin Heidelberg, Berlin, Heidelberg, 2002); p. 1.

104. Park E.-S., Cho H.-C., Kim M.-N., and Yoon J.-S.: Chain extension and mechanical properties of unsaturated aliphatic copolyesters based on poly(L-lactic acid). J. Appl. Polym. Sci. 90, 1802 (2003).

105. Eyvazzadeh Kalajahi A., Rezaei M., Abbasi F., and Mir Mohamad Sadeghi G.: The effect of chain extender type on the physical, mechanical, and shape memory properties of poly(ε-caprolactone)-based polyurethane-ureas. Polym. Plast. Technol. Eng. 56, 1977 (2017).

106. Zhao J.-B., Wu X.-F., and Yang W.-T.: Synthesis of aliphatic polyesters by a chain-extending reaction with octamethylcyclotetrasilazane and hexaphenylcyclotrisilazane as chain extenders. J. Appl. Polym. Sci. 92, 3333 (2004).

107. Löfgren A., Albertsson A.-C., Dubois P., and Jérôme R.: Recent advances in ring-opening polymerization of lactones and related compounds. J. Macromol. Sci. Part C: Polym. Rev. 35, 379 (1995).

108. Webb A.R., Yang J., and Ameer G.A.: Biodegradable polyester elastomers in tissue engineering. Expert Opin. Biol. Ther. 4, 801 (2004).

109. Tokiwa Y. and Calabia B.P.: Review degradation of microbial polyesters. Biotechnol. Lett. 26, 1181 (2004).

110. Silvers A.L., Chang C.-C., Parrish B., and Emrick T.: Strategies in Aliphatic Polyester Synthesis for Biomaterial and Drug Delivery Applications, Degradable Polymers and Materials: Principles and Practice, 2nd ed. (American Chemical Society, Washington D.C., 2012); p. 237.

111. Hakkarainen M.: Aliphatic Polyesters: Abiotic and Biotic Degradation and Degradation Products (Degradable Aliphatic Polyesters, Springer, Berlin, Heidelberg, 2002); p. 113.

112. Hilf J. and Frey H.: Propargyl-functional aliphatic polycarbonate obtained from carbon dioxide and glycidyl propargyl ether. Macromol. Rapid Commun. 34, 1395 (2013).

113. Liu F., Yang J., Fan Z., Li S., Kasperczyk J., and Dobrzynski P.: Enzyme-catalyzed degradation of biodegradable polymers derived from trimethylene carbonate and glycolide by lipases from Candida Antarctica and Hog pancreas. J. Biomater. Sci. Polym. Ed. 23, 1355 (2012).

114. Kaplan M.L., Rietman E.A., Cava R.J., Holt L.K., and Chandross E.A.: Crown ether enhancement of ionic conductivity in a polymer-salt system. Solid State Ionics 25, 37 (1987).

115. Wei X. and Shriver D.F.: Highly conductive polymer electrolytes containing rigid polymers. Chem. Mater. 10, 2307 (1998).

116. Matsumoto K., Kakehashi M., Ouchi H., Yuasa M., and Endo T.: Synthesis and properties of polycarbosilanes having 5-membered cyclic carbonate groups as solid polymer electrolytes. Macromolecules 49, 9441 (2016).

117. Chai J., Liu Z., Ma J., Wang J., Liu X., Liu H., Zhang J., Cui G., and Chen L.: In situ generation of poly(vinylene carbonate) based solid electrolyte with interfacial stability for LiCoO2 lithium batteries. Adv. Sci. 4, 1600377 (2017).

118. Xu H., Bijleveld J., Hedge M., and Dingemans T.: Synthesis and characterization of aromatic-PDMS segmented block copolymers and their shape-memory performance. Polym. Chem. 10, 5052 (2019).

119. Soo P.P., Huang B., Jang Y.I., Chiang Y.M., Sadoway D.R., and Mayes A.M.: Rubbery block copolymer electrolytes for solid-state rechargeable lithium batteries. J. Electrochem. Soc. 146, 32 (1999).

120. Mitsuda H., Uno T., Kubo M., and Itoh T.: Solid polymer electrolytes based on poly(1,3-diacetyl-4-imidazolin-2-one). Polym. Bull. 57, 313 (2006).

121. Itoh T., Fujita K., Inoue K., Iwama H., Kondoh K., Uno T., and Kubo M.: Solid polymer electrolytes based on alternating copolymers of vinyl ethers with methoxy oligo(ethyleneoxy)ethyl groups and vinylene carbonate. Electrochim. Acta 112, 221 (2013).

122. Wang P., Chai J., Zhang Z., Zhang H., Ma Y., Xu G., Du H., Liu T., Li G., and Cui G.: An intricately designed poly(vinylene carbonate-acrylonitrile) copolymer electrolyte enables 5 V lithium batteries. J. Mater. Chem. A 7, 5295 (2019).

123. Britz J., Meyer W.H., and Wegner G.: Blends of poly(meth)acrylates with 2-oxo-(1,3)dioxolane side chains and lithium salts as lithium ion conductors. Macromolecules 40, 7558 (2007).

124. Tominaga Y.: Ion-conductive polymer electrolytes based on poly(ethylene carbonate) and its derivatives. Polym. J. 49, 291 (2017).

125. Spiegel E.F., Adamic K.J., Williams B.D., and Sammells A.F.: Solvation of lithium salts within single-phase dimethyl siloxane bisphenol—A carbonate block copolymer. Polymer 41, 3365 (2000).

126. Matsumoto M., Uno T., Kubo M., and Itoh T.: Polymer electrolytes based on polycarbonates and their electrochemical and thermal properties. Ionics 19, 615 (2013).

127. Abdul-Karim R., Hameed A., and Malik M.I.: Ring-opening polymerization of ethylene carbonate: Comprehensive structural elucidation by 1D & 2D-NMR techniques, and selectivity analysis. RSC Adv. 7, 11786 (2017).

128. Lee J.-C. and Litt M.H.: Ring-opening polymerization of ethylene carbonate and depolymerization of poly(ethylene oxide-co-ethylene carbonate). Macromolecules 33, 1618 (2000).

129. Dukhanin G.P., Dumler S.A., Sablin A.N., and Novakov I.A.: Solid polymeric electrolyte based on poly(ethylene carbonate)-lithium perchlorate system. Russ. J. Appl. Chem. 82, 243 (2009).

130. Tominaga Y., Nanthana V., and Tohyama D.: Ionic conduction in poly(ethylene carbonate)-based rubbery electrolytes including lithium salts. Polym. J. 44, 1155 (2012).

131. Kimura K., Hassoun J., Panero S., Scrosati B., and Tominaga Y.: Electrochemical properties of a poly(ethylene carbonate)-LiTFSI electrolyte containing a pyrrolidinium-based ionic liquid. Ionics 21, 895 (2015).

132. Kimura K., Matsumoto H., Hassoun J., Panero S., Scrosati B., and Tominaga Y.: A quaternary poly(ethylene carbonate)-lithium bis(trifluoromethanesulfonyl)imide-ionic liquid-silica fiber composite polymer electrolyte for lithium batteries. Electrochim. Acta 175, 134 (2015).

133. Motomatsu J., Kodama H., Furukawa T., and Tominaga Y.: Dielectric relaxation behavior of a poly(ethylene carbonate)-lithium bis-(trif luoromethanesulfonyl) imide electrolyte. Macromol. Chem. Phys. 216, 1660 (2015).

134. Tominaga Y., Yamazaki K., and Nanthana V.: Effect of anions on lithium ion conduction in poly(ethylene carbonate)-based polymer electrolytes. J. Electrochem. Soc. 162, A3133 (2015).

https://doi.org/10.1557/mre.2020.3Downloaded from https://www.cambridge.org/core. IP address: 54.39.106.173, on 12 Jan 2021 at 17:29:03, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.

Page 24: Carbonyl-coordinating polymers Hongli Xu Jingbing Xie ...€¦ · lithium batteries. Why carbonyl-coordinating polymers The dawn of the investigation on SPEs was from the work of

24 n MRS eNeRgy & SUSTAINABIlITy // V O L U M E 7 // e 2 // www.mrs.org/energy-sustainability-journal

135. Kimura K., Motomatsu J., and Tominaga Y.: Highly concentrated polycarbonate-based solid polymer electrolytes having extraordinary electrochemical stability. J. Polym. Sci., Part B: Polym. Phys. 54, 2442 (2016).

136. Kimura K., Yajima M., and Tominaga Y.: A highly-concentrated poly(ethylene carbonate)-based electrolyte for all-solid-state Li battery working at room temperature. Electrochem. Commun. 66, 46 (2016).

137. Morioka T., Ota K., and Tominaga Y.: Effect of oxyethylene side chains on ion-conductive properties of polycarbonate-based electrolytes. Polymer 84, 21 (2016).

138. Morioka T., Nakano K., and Tominaga Y.: Ion-conductive properties of a polymer electrolyte based on ethylene carbonate/ethylene oxide random copolymer. Macromol. Rapid Commun. 38, 1600652 (2017).

139. Motomatsu J., Kodama H., Furukawa T., and Tominaga Y.: Dielectric relaxation and ionic transport in poly(ethylene carbonate)-based electrolytes. Polym. Adv. Technol. 28, 362 (2017).

140. Kimura K. and Tominaga Y.: Understanding electrochemical stability and lithium ion-dominant transport in concentrated poly(ethylene carbonate) electrolyte. ChemElectroChem 5, 4008 (2018).

141. Munshi M.Z.A., Owens B.B., and Nguyen S.: Measurement of Li+ ion transport numbers in poly(ethylene oxide)–LiX complexes. Polym. J. 20, 597 (1988).

142. Tominaga Y., Shimomura T., and Nakamura M.: Alternating copolymers of carbon dioxide with glycidyl ethers for novel ion-conductive polymer electrolytes. Polymer 51, 4295 (2010).

143. Nakamura M. and Tominaga Y.: Utilization of carbon dioxide for polymer electrolytes [II]: Synthesis of alternating copolymers with glycidyl ethers as novel ion-conductive polymers. Electrochim. Acta 57, 36 (2011).

144. Smith M.J., Silva M.M., Cerqueira S., and MacCallum J.R.: Preparation and characterization of a lithium ion conducting electrolyte based on poly(trimethylene carbonate). Solid State Ionics 140, 345 (2001).

145. Manuela Silva M., Barbosa P., Evans A., and Smith M.J.: Novel solid polymer electrolytes based on poly(trimethylene carbonate) and lithium hexafluoroantimonate. Solid State Sci. 8, 1318 (2006).

146. Sun B., Mindemark J., Edström K., and Brandell D.: Realization of high performance polycarbonate-based Li polymer batteries. Electrochem. Commun. 52, 71 (2015).

147. Sun B., Xu C., Mindemark J., Gustafsson T., Edström K., and Brandell D.: At the polymer electrolyte interfaces: The role of the polymer host in interphase layer formation in Li-batteries. J. Mater. Chem. A 3, 13994 (2015).

148. Sun B., Mindemark J., Morozov E.V., Costa L.T., Bergman M., Johansson P., Fang Y., Furó I., and Brandell D.: Ion transport in polycarbonate based solid polymer electrolytes: Experimental and computational investigations. Phys. Chem. Chem. Phys. 18, 9504 (2016).

149. Meabe L., Lago N., Rubatat L., Li C., Müller A.J., Sardon H., Armand M., and Mecerreyes D.: Polycondensation as a versatile synthetic route to aliphatic polycarbonates for solid polymer electrolytes. Electrochim. Acta 237, 259 (2017).

150. He W., Cui Z., Liu X., Cui Y., Chai J., Zhou X., Liu Z., and Cui G.: Carbonate-linked poly(ethylene oxide) polymer electrolytes towards high performance solid state lithium batteries. Electrochim. Acta 225, 151 (2017).

151. Liu X., Ding G., Zhou X., Li S., He W., Chai J., Pang C., Liu Z., and Cui G.: An interpenetrating network poly(diethylene glycol carbonate)-based polymer electrolyte for solid state lithium batteries. J. Mater. Chem. A 5, 11124 (2017).

152. Jung Y.-C., Park M.-S., Kim D.-H., Ue M., Eftekhari A., and Kim D.-W.: Room-temperature performance of poly(ethylene ether carbonate)-based solid polymer electrolytes for all-solid-state lithium batteries. Sci. Rep. 7, 17482 (2017).

153. Melchiors M., Keul H., and Höcker H.: Preparation and properties of solid electrolytes on the basis of alkali metal salts and poly(2,2-dimethyltrimethylene carbonate)-block-poly(ethylene oxide)-block-poly(2,2-dimethyltrimethylene carbonate). Polymer 37, 1519 (1996).

154. Elmér A.M. and Jannasch P.: Synthesis and characterization of poly(ethylene oxide-co-ethylene carbonate) macromonomers and their use

in the preparation of crosslinked polymer electrolytes. J. Polym. Sci., Part A: Polym. Chem. 44, 2195 (2006).

155. Yu X., Xiao M., Wang S., Han D., and Meng Y.: Fabrication and properties of crosslinked poly(propylene carbonate maleate) gel polymer electrolyte for lithium-ion battery. J. Appl. Polym. Sci. 118, 2078 (2010).

156. Kwon S.-J., Kim D.-G., Shim J., Lee J.H., Baik J.-H., and Lee J.-C.: Preparation of organic/inorganic hybrid semi-interpenetrating network polymer electrolytes based on poly(ethylene oxide-co-ethylene carbonate) for all-solid-state lithium batteries at elevated temperatures. Polymer 55, 2799 (2014).

157. Deng K., Wang S., Ren S., Han D., Xiao M., and Meng Y.: A novel single-ion-conducting polymer electrolyte derived from CO2-based multifunctional polycarbonate. ACS Appl. Mater. Interfaces 8, 33642 (2016).

158. Meabe L., Goujon N., Li C., Armand M., Forsyth M., and Mecerreyes D.: Single-ion conducting poly(ethylene oxide carbonate) as solid polymer electrolyte for lithium batteries. Batteries Supercaps 3, 68 (2020).

159. Huang X., Huang J., Wu J., Yu X., Gao Q., Luo Y., and Hu H.: Fabrication and properties of polybutadiene rubber-interpenetrating cross-linking poly(propylene carbonate) network as gel polymer electrolytes for lithium-ion battery. RSC Adv. 5, 52978 (2015).

160. Huang X., Zeng S., Liu J., He T., Sun L., Xu D., Yu X., Luo Y., Zhou W., and Wu J.: High-performance electrospun poly(vinylidene fluoride)/poly(propylene carbonate) gel polymer electrolyte for lithium-ion batteries. J. Phys. Chem. C 119, 27882 (2015).

161. Zhao J., Zhang J., Hu P., Ma J., Wang X., Yue L., Xu G., Qin B., Liu Z., Zhou X., and Cui G.: A sustainable and rigid-flexible coupling cellulose-supported poly(propylene carbonate) polymer electrolyte towards 5 V high voltage lithium batteries. Electrochim. Acta 188, 23 (2016).

162. Shin J.-H., Henderson W.A., and Passerini S.: Ionic liquids to the rescue? Overcoming the ionic conductivity limitations of polymer electrolytes. Electrochem. Commun. 5, 1016 (2003).

163. Shin J.-H., Henderson W.A., and Passerini S.: PEO-based polymer electrolytes with ionic liquids and their use in lithium metal-polymer electrolyte batteries. J. Electrochem. Soc. 152, A978 (2005).

164. Wu H., Xu Y., Ren X., Liu B., Engelhard M.H., Ding M.S., El-Khoury P.Z., Zhang L., Li Q., Xu K., Wang C., Zhang J.-G., and Xu W.: Polymer-in-“Quasi-ionic liquid” electrolytes for high-voltage lithium metal batteries. Adv. Energy Mater. 9, 1902108 (2019).

165. Zhou D., Zhou R., Chen C., Yee W.-A., Kong J., Ding G., and Lu X.: Non-volatile polymer electrolyte based on poly(propylene carbonate), ionic liquid, and lithium perchlorate for electrochromic devices. J. Phys. Chem. B 117, 7783 (2013).

166. Zhang J., Zang X., Wen H., Dong T., Chai J., Li Y., Chen B., Zhao J., Dong S., Ma J., Yue L., Liu Z., Guo X., Cui G., and Chen L.: High-voltage and free-standing poly(propylene carbonate)/Li6.75La3Zr1.75Ta0.25O12 composite solid electrolyte for wide temperature range and flexible solid lithium ion battery. J. Mater. Chem. A 5, 4940 (2017).

167. He Z., Chen L., Zhang B., Liu Y., and Fan L.: Flexible poly(ethylene carbonate)/garnet composite solid electrolyte reinforced by poly(vinylidene fluoride-hexafluoropropylene) for lithium metal batteries. J. Power Sources 392, 232 (2018).

168. Imholt L., Dörr T.S., Zhang P., Ibing L., Cekic-Laskovic I., Winter M., and Brunklaus G.: Grafted polyrotaxanes as highly conductive electrolytes for lithium metal batteries. J. Power Sources 409, 148 (2019).

169. Zhu M., Wu J., Wang Y., Song M., Long L., Siyal S.H., Yang X., and Sui G.: Recent advances in gel polymer electrolyte for high-performance lithium batteries. J. Energy Chem. 37, 126 (2019).

170. Florjańczyk Z., Zygadło-Monikowska E., Wieczorek W., Ryszawy A., Tomaszewska A., Fredman K., Golodnitsky D., Peled E., and Scrosati B.: Polymer-in-salt electrolytes based on acrylonitrile/butyl acrylate copolymers and lithium salts. J. Phys. Chem. B 108, 14907 (2004).

171. Łasińska A.K., Marzantowicz M., Dygas J.R., Krok F., Florjańczyk Z., Tomaszewska A., Zygadło-Monikowska E., Żukowska Z., and Lafont U.: Study of ageing effects in polymer-in-salt electrolytes based on poly(acrylonitrile-co-butyl acrylate) and lithium salts. Electrochim. Acta 169, 61 (2015).

https://doi.org/10.1557/mre.2020.3Downloaded from https://www.cambridge.org/core. IP address: 54.39.106.173, on 12 Jan 2021 at 17:29:03, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.

Page 25: Carbonyl-coordinating polymers Hongli Xu Jingbing Xie ...€¦ · lithium batteries. Why carbonyl-coordinating polymers The dawn of the investigation on SPEs was from the work of

MRS eNeRgy & SUSTAINABIlITy // V O L U M E 7 // e 2 // www.mrs.org/energy-sustainability-journal n 25

172. Florjańczyk Z., Zygadło-Monikowska E., Affek A., Tomaszewska A., Łasińska A., Marzantowicz M., Dygas J.R., and Krok F.: Polymer electrolytes based on acrylonitrile–butyl acrylate copolymers and lithium bis(trifluoromethanesulfone)imide. Solid State Ionics 176, 2123 (2005).

173. Wu I.D. and Chang F.-C.: Determination of the interaction within polyester-based solid polymer electrolyte using FTIR spectroscopy. Polymer 48, 989 (2007).

174. Ravi M., Song S.-H., Gu K.-M., Tang J.-N., and Zhang Z.-Y.: Effect of lithium thiocyanate addition on the structural and electrical properties of biodegradable poly(ε-caprolactone) polymer films. Ionics 21, 2171 (2015).

175. Ravi M., Song S., Gu K., Tang J., and Zhang Z.: Electrical properties of biodegradable poly(ɛ-caprolactone): Lithium thiocyanate complexed polymer electrolyte films. Mater. Sci. Eng., B 195, 74 (2015).

176. Polo Fonseca C. and Neves S.: Electrochemical properties of a biodegradable polymer electrolyte applied to a rechargeable lithium battery. J. Power Sources 159, 712 (2006).

177. Fonseca C.P., Rosa D.S., Gaboardi F., and Neves S.: Development of a biodegradable polymer electrolyte for rechargeable batteries. J. Power Sources 155, 381 (2006).

178. Lin C.-K. and Wu I.D.: Investigating the effect of interaction behavior on the ionic conductivity of polyester/LiClO4 blend systems. Polymer 52, 4106 (2011).

179. Watanabe M., Togo M., Sanui K., Ogata N., Kobayashi T., and Ohtaki Z.: Ionic conductivity of polymer complexes formed by poly(β-propiolactone) and lithium perchlorate. Macromolecules 17, 2908 (1984).

180. Watanabe M., Rikukawa M., Sanui K., Ogata N., Kato H., Kobayashi T., and Ohtaki Z.: Ionic conductivity of polymer complexes formed by poly(ethylene succinate) and lithium perchlorate. Macromolecules 17, 2902 (1984).

181. Watanabe M., Rikukawa M., Sanui K., and Ogata N.: Effects of polymer structure and incorporated salt species on ionic conductivity of polymer complexes formed by aliphatic polyester and alkali metal thiocyanate. Macromolecules 19, 188 (1986).

182. Dupon R., Papke B.L., Ratner M.A., and Shriver D.F.: Ion transport in the polymer electrolytes formed between poly(ethylene succinate) and lithium tetrafluoroborate. J. Electrochem. Soc. 131, 586 (1984).

183. Lee Y.-C., Ratner M.A., and Shriver D.F.: Ionic conductivity in the poly(ethylene malonate)/lithium triflate system. Solid State Ionics 138, 273 (2001).

184. Pesko D.M., Jung Y., Hasan A.L., Webb M.A., Coates G.W., Miller T.F., and Balsara N.P.: Effect of monomer structure on ionic conductivity in a systematic set of polyester electrolytes. Solid State Ionics 289, 118 (2016).

185. Webb M.A., Jung Y., Pesko D.M., Savoie B.M., Yamamoto U., Coates G.W., Balsara N.P., Wang Z.-G., and Miller T.F.: Systematic computational and experimental investigation of lithium-ion transport mechanisms in polyester-based polymer electrolytes. ACS Cent. Sci. 1, 198 (2015).

186. Van Horn R.M., Steffen M.R., and O’Connor D.: Recent progress in block copolymer crystallization. Polym. Cryst. 1, e10039 (2018).

187. Mindemark J., Törmä E., Sun B., and Brandell D.: Copolymers of trimethylene carbonate and ε-caprolactone as electrolytes for lithium-ion batteries. Polymer 63, 91 (2015).

https://doi.org/10.1557/mre.2020.3Downloaded from https://www.cambridge.org/core. IP address: 54.39.106.173, on 12 Jan 2021 at 17:29:03, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.