Solid polymer electrolytes (SPEs) are promising candidates for usage in rechargeable lithium metal batteries (LMBs) as they possess high mechanical, thermal, and chemical stability. However, the poor ionic conductivity of SPEs in comparison to liquid electrolytes hampers the commercialization of SPE-based LMBs. In the case of poly[bis(methoxy-ethoxy-ethoxy-)phosphazene] (MEEP), one explanation for the low ionic conductivity is the trapping of lithium cations in backbone coordination sites, hindering lithium ion movement through the electrolyte membrane. Herein, modelling the ion coordination in MEEP using DFT calculations reveals that, compared to lithium, heavier alkali cations are more likely to be complexed at the backbone coordination sites. With other alkali cations masking these coordination sites, enhanced lithium ion mobility through the SPE is expected. Experimental data proves these expectations: doping MEEP-based LiBOB-containing SPE membranes with small amounts of in-house synthesized potassium bis(oxalato)borate (KBOB) increases the lithium ion transference number from 0.08 to 0.18. Also, the partial lithium ion conductivity of the salt-in-MEEP electrolyte is boosted to outstanding 0.08 mS cm−1, far exceeding state-of-the-art literature values for this material. A cross-check using SPEs based on the structurally similar poly(ethylene oxide) (PEO) validates the proposed cation displacement model. The obtained insights may aid the development of highly effective poly(phosphazene)-based SPEs.
Since the market for electric cars and portable devices has been continuously growing during the last years, the demand for high-energy density batteries with increased lifetime and safety is as high as never before. 1–3 This raised the interest in developing advanced lithium-ion batteries (LIBs) and alternative battery concepts with increased energy density are also actively investigated, especially for use in electric vehicles. 4–8 Next generation batteries like lithium metal batteries (LMBs), including sulfur∣∣lithium and air∣∣lithium, are promising candidates to complement the current systems, with lithium metal offering a very high theoretical specific capacity (3862 mAh g–1) and the lowest standard redox potential of all metallic anode materials (−3.04 V vs SHE). 2,9–12 Unfortunately, the tremendous hurdles associated with the usage of lithium metal as an electrode material made a commercialization of rechargeable LMBs hard to realize so far. The biggest challenge is the formation of inhomogeneous lithium metal deposits during the electrodeposition/electrodissolution processes leading to so-called high surface area lithium (HSAL). 13 Especially dendritic needle-like HSAL morphologies can penetrate the separator and grow towards the cathode, inducing a cell short circuit. This can result in a thermal runaway and cause cell safety issues. 10,12,14,15 Another consequence of HSAL occurrence is a loss of active material due to both the formation of electrochemically isolated dead lithium and continuous side reactions between fresh lithium surfaces and the electrolyte. 16–18 This leads to a poor Coulombic efficiency and cycle life performance of LMBs. 10,12,16,19
To counteract these safety and performance issues, many approaches have been pursued, such as mechanical modification of the lithium metal, which can increase the surface area to reduce the local current density. 13,17,20–22 Mechanical modifications can also be used to smoothen the Li metal surface to reduce defect sites, which is beneficial for battery setups with both liquid or solid electrolytes. 14,23 Another promising approach is the chemical modification of Li metal to generate a robust solid electrolyte interphase (SEI 24 ) as a protective layer, often called artificial SEI (aSEI), that can suppress dendritic growth of lithium metal. 17,19,25–27 Furthermore, the addition of alkali metal ions like cesium to the electrolyte showed a dendrite-suppressing effect as the ions were believed to adsorb but not to deposit at the lithium metal surface building a positively charged electrostatic shield around potential "hot" spots for dendrite formation. 28,29
Moreover, polymer- and ceramic-based solid electrolytes have been proven to be stable barriers against dendrite growth that direct the lithium to be electrodeposited more homogeneously. 30–32 Compared to ceramic electrolytes, especially solid polymer electrolytes (SPEs) feature a high mechanical flexibility and interfacial compatibility coupled with favorable handling characteristics and thin cell processing capabilities. Compared to liquid organic electrolytes, SPEs show an increased safety since they are leakproof and feature a low volatility, which leads to a lower risk of gas pressure increase and leakage of toxic gases. 33
One of the most widely studied polymers to be used in SPEs is poly(ethylene oxide) (PEO). 34–37 For example, SPEs derived from this polymer are already utilized in the form of LiFePO4│SPE│Li batteries in electric vehicles called "Bluecar" which are manufactured by the French company Bolloré. 38,39 Due to the strong donor character of the ether oxygen atoms in the PEO chains, lithium cations are readily dissolved in such SPEs. These oxygen atoms are further known to play a key role in the ionic conduction within the electrolyte. Ions are transported through the polymer by interactions with the ether sites and follow a so-called inter- and intrachain-hopping principle. 40,41 Moreover, PEO has the benefits of high chemical and electrochemical stability against lithium metal and a good chain flexibility, which is also facilitated by its low glass transition temperature. 42–44 Nevertheless, a drawback of PEO-based polymer electrolytes is the partial formation of crystalline domains below about 60 °C leading to a low total conductivity at room temperature. 45
Alongside PEO, SPEs based on poly(siloxane), poly(acrylate), and poly(phosphazene) have also been described, among others. 46–48 One representative of the group of poly(phosphazenes) is poly[bis(2-(2-methoxyethoxy)ethoxy)phosphazene] (MEEP). This polymer consists of the typical inorganic poly(phosphazene) backbone made of alternating phosphorus and nitrogen atoms decorated with PEO-like side chains, yielding a grafted polymer structure. 49 The structural similarity between MEEP and PEO can be observed when comparing their structures shown in Figs 1a and 1b. Compared to carbon-based polymer backbones, the segmental mobility is increased by the low torsional barrier of the N-P-N units along the polymer backbone. 45 Another advantage of MEEP is the low glass transition temperature Tg of ≈–85 °C, which supports fast ion transport. This property leads to MEEP's room temperature ionic conductivity surpassing that of PEO by at least a factor of three. 49 Moreover, MEEP has a good thermal stability and exhibits a high electrochemical stability against oxidation and reduction. 50,51
Theory and Modelling
Theoretical considerations
For achieving higher practical relevance, it is important to further increase the ionic conductivity of solvent-free MEEP-based electrolytes. At room temperature the total ionic conductivity σtotal of solid MEEP with different amounts of dissolved lithium salts can reach around 10–4 S cm–1, but as the lithium ion transference number tLi+ in such systems is typically lower than 0.1, the partial lithium ion conductivity σLi+ is often below 10–5 S cm–1. 52,53 The main reason for this unsatisfactory performance is that the lithium cations are coordinated not only by the oxygen atoms in the side chains but also by the nitrogen atoms in the backbone, which leads to pocket-like coordination structures. The strong coordination to the nitrogen atoms along the poly(phosphazene) backbone traps lithium ions and therefore greatly hinders their propagation through the polymer electrolyte. 52–54
Thus, the trapping of lithium ions in these coordination structures needs to be prevented to increase the lithium ion mobility in MEEP-based SPEs. To gain a better understanding of the lithium ion coordination situation at the MEEP backbone, density functional theory (DFT 55 ) calculations have been performed here. First, a lithium coordination by only ether oxygen atoms in the side chains of MEEP was compared to a coordination directly at the backbone, in which case the lithium is coordinated by a nitrogen atom from the backbone plus additional ether oxygen atoms from the surrounding side chains. This comparison revealed a distinct energetic preference of the backbone-coordination towards lithium. Starting from such a lithium coordination at the MEEP backbone, the lithium cation was replaced by various other alkali and alkaline earth ions and the energy changes for these theoretical "cation exchange reactions" were computed. It was found that all heavier alkali metal ions showed an energetically more favorable coordination to the nitrogen atoms at the MEEP backbone than the lithium cation, with the strongest interaction observed for the potassium ion. This leads to a cation displacement hypothesis, stating that it should be possible to free the lithium ions from the polymer backbone by blocking the nitrogen coordination sites with other alkali metal ions. This would release the lithium ions and allow them a higher mobility along the weakly coordinating polyether side chains, similar to the literature-known lithium ion conduction mechanism in PEO, 40 enabling a higher partial lithium ion conductivity through the MEEP-based SPE.
Quantum chemical calculations
To investigate the proposed cation displacement hypothesis, quantum chemical calculations have been performed with the software Turbomole 7.3 56 using the PBE functional and a def2-TZVP basis set. 57–59 Borodin et al. and Bieker et al. have both compared calculations with the computationally expedient PBE functional with more accurate but computationally expensive G4MP2 calculations, and they reported good agreements for both lithium and magnesium complexes with various popular electrolyte solvents, including ethers. 60,61 Following these findings, the PBE functional has also been used within this work. First, different coordination scenarios regarding lithium in MEEP were investigated using DFT computations. A cluster continuum approach was used, in which the lithium ion and the solvent within the first solvation shell were treated explicitly while the solvent in further distances to the central ion was included via an implicit solvent model. 62,63 The Conductor-like Screening Model (COSMO) implemented in Turbomole was used within this work for the implicit solvent calculations with a dielectric constant ε = 20. 60,61,64 This value was chosen following previous studies by Hall et al., who calculated dielectric constants for a range of different solvent blends typically used in LIBs, and Borodin et al., who showed that the stabilization effect of cationic clusters resulting from the implicit solvent largely saturates for ε > 20. 60,65
Within the performed calculations, the MEEP polymer was approximated by a MEEP oligomer consisting of five monomers in order to make the calculations computationally feasible. No vibrational frequencies were calculated; therefore, all energetic terms correspond to internal potential energies. Figure 2 shows two different coordination situations for lithium in MEEP: a backbone-coordination, where the lithium is coordinated by a backbone nitrogen atom plus three oxygen atoms from surrounding side chains (pocket-like coordination), and a PEO-like side chain coordination, where the lithium is coordinated by only oxygen atoms from two side chains. Comparing the potential energies of these geometry-optimized structures yields a very distinct energetic preference of 10 kcal mol–1 towards the backbone-coordination. This finding strongly supports the earlier-mentioned hypothesis of the lithium ions being captured and locked within the pocket-like coordination sites at the backbone, hampering ion mobility. 54
Our approach to increase lithium ion mobility within MEEP is based upon releasing the lithium ions from those backbone coordination sites and transferring them towards the more mobile and less strongly coordinating side chains. To this end, we decided to introduce further cations other than lithium into the system. Those other ions should feature an even higher energetic preference towards the backbone coordination sites to occupy the backbone slots and force the lithium ions towards the remaining side chain coordination sites. To find suitable cations (X), structures of MEEP with backbone coordination to lithium, sodium, potassium, and cesium, as well as magnesium cations were computed. In addition, the corresponding ethylene carbonate (EC) complexes X(EC)4 were also simulated and were used to compute virtual reaction energies ΔE of the (purely theoretical) cation exchange reactions
The choice of this reference system was motivated by the fact that MEEP can also be used as a gel polymer electrolyte matrix, i.e. including additional liquid electrolyte solvents like EC or DMC. Here, the lithium mobility is similarly hindered by the strong coordination to the MEEP backbone and a release of lithium ions into the liquid electrolyte would be desirable. With the help of this virtual reaction energy, the energetic preferences of the coordination of the different cations to MEEP can be compared. A negative value of indicates an energetic preference of the corresponding cation-MEEP coordination over the lithium-MEEP coordination. Figure 3 displays this comparison and indicates a higher energetic preference of the X-MEEP coordination for all alkali metal ions heavier than lithium (X = Na, K, Rb, Cs). While the value for the sodium ion is at a comparatively small −8.8 kcal mol–1, all other investigated alkali metal ions feature values at ≈−15 kcal mol–1. The highest energetic preference towards the coordination at the MEEP backbone is found for the potassium ion (−16.4 kcal mol–1). In contrast, the magnesium-MEEP coordination displays a positive of 10.8 kcal mol–1 indicating a lower preference than for the lithium-MEEP coordination. These results suggest the possibility of a lithium ion displacement at the nitrogen coordination sites of the polymer backbone by doping the system with additional, heavier alkali metal ions, which might then increase the lithium ion mobility by forcing the lithium ions to occupy the less strongly coordinating sidechain coordination sites in a fashion similar to the one observed in PEO. 40
It should be noted that these calculations ignore possible entropic contributions to the coordination behavior in MEEP. Furthermore, only an approximation of the MEEP polymer in the form of a short oligomer was studied and possible anion effects were ignored by omitting the anions from the calculations. Nonetheless, the energetic preferences shown in Fig. 3a are significant, such that the results presented here can be seen as an applicable indication, keeping in mind that further calculations might be helpful in order to achieve a profound understanding of the ion coordination and mobility in MEEP. To verify the qualitative validity of the presented theoretical findings, experimental studies with MEEP-based SPE systems were conducted. Based on the promising results from the DFT calculations, KBOB was chosen as the representative salt to confirm the assumed cation doping induced lithium ion displacement mechanism.
Materials and methods
Materials
Suppliers
Lithium bis(oxalato)borate (LiBOB, 97.3%) and 1-butyl-1-methylpyrrolidinium bis(oxalato)borate (Pyr14BOB, 99.9%) were obtained from Solvionic. Poly(ethylene oxide) (PEO, Mv = 4 000 000 g mol−1), sodium hydroxide (NaOH, 99.99%), potassium hydroxide (KOH, 99.99%), cesium hydroxide (CsOH, 90%), oxalic acid (C2H2O4, 99%), boric acid (H3BO3, 99%), dimethyl sulfoxide (DMSO, 99.9%, anhydrous), and tetrahydrofuran (THF, 99.9%, anhydrous) were all purchased from Sigma-Aldrich. Benzophenone (C13H10O, 99%) was obtained from Acros. LiBOB, Pyr14BOB, PEO, and benzophenone were dried using a HiCube 80 Eco vacuum pump (Pfeiffer Vacuum, Aßlar, Germany) connected to a Büchi Glass Oven B-585 drying oven (BÜCHI Labortechnik AG, Flawil, Switzerland). All other components were used as received.
Synthesis of poly[bis(methoxy-ethoxy-ethoxy-)phosphazene] (MEEP)
The polymer MEEP was prepared as has been described by Jankowsky et al. 53,66
Synthesis of the alkali metal bis(oxalato)borate salts
The synthesis scheme used to prepare potassium bis(oxalato)borate (KBOB) was adapted from the patent by Wietelmann et al. concerning the synthesis of LiBOB. 67 Oxalic acid (14.41 g, 0.16 mol) and potassium hydroxide (KOH, 4.50 g, 0.08 mol) were suspended in 120 ml of deionized water and heated to 50 °C. Separately, boric acid (H3BO3, 4.95 g, 0.08 mol) was dissolved in 100 ml of deionized water at 50 °C. Over the course of 15 min the H3BO3 solution was added dropwise into the constantly stirred KOH-containing mixture. The solution was heated to boiling and vigorously stirred for ≈ 4 h until the water was totally evaporated. This yielded a white precipitate that was pre-dried at 110 °C for 12 h. Further drying at 80 °C was successively executed at pressures of 510−2 mbar and 10−6 mbar for 48 h each, which yielded the final product. The above procedure was also tested for the synthesis of NaBOB and CsBOB using NaOH (3.20 g, 0.08 mol) or CsOH (13.44 g, 0.08 mol) instead of the KOH.
Preparation of MEEP membranes
In order to fabricate a MEEP SPE membrane, first benzophenone (30 mg, 0.165 mmol, 23.4 mol%), pristine MEEP (200 mg, 0.706 mmol), and variable amounts of KBOB (between 2.5 mol% and 10 mol%, ratio of potassium ions to oxygen atoms in MEEP is between 1:240 and 1:60) were dissolved in THF (200 μl) and DMSO (100 μl) under stirring. LiBOB (20 mg, 0.103 mmol, 14.6 mol%, ratio of lithium ions to oxygen atoms in MEEP is 1:41) was added to the mix and the resulting viscous liquid was dried over night at 10–1 mbar at room temperature before heating it up to 80 °C for 30 min at 10−2 mbar. The now highly viscous mixture was placed between two siliconized poly(ethylene terephthalate) (PET) foils (bo-PET, PPI Adhesive Products Ltd, 100 μm thickness) and manually pressed between two flat metal plates to form a polymer layer with a thickness of 150 μm. For cross-linking, each side of the sample was exposed to UV-A irradiation in a UVACube 100 (Dr. Hönle AG) for 10 min. This resulted in a flexible membrane that was carefully separated from the foil layers using a 100 μm thin razor blade.
Preparation of PEO membranes
SPE membranes based on PEO were fabricated by first mixing the ionic liquid (IL) Pyr14BOB (0.74 g, 2.25 mmol, 19.8 mol%) with benzophenone (0.025 g, 0.14 mmol, 1.2 mol%) and heating the powder blend to 80 °C in order to liquify the IL. In a separate vial, PEO (0.5 g, 11.36 mmol, Mv = 4 000 000 g mol−1) was thoroughly mixed with LiBOB (0.22 g, 1.14 mmol, 10.0 mol%, ratio of lithium ions to oxygen atoms in PEO is 1:10) and variable amounts of KBOB (between 2.5 mol% and 10 mol%, ratio of potassium ions to oxygen atoms in PEO is between 1:40 and 1:10) and the resulting mixture was added to the IL vial. The combined components were stirred at 80 °C which quickly led to the formation of a white gum-like mass that could not be stirred any further. The mass was vacuum sealed and annealed at 100 °C for 48 h. The resulting transparent mixture was pressed at 100 °C between two siliconized bo-PET sheets (PPI Adhesive Products Ltd, 100 μm thickness) using a SERVITEC Polystat 200 T device. The applied pressure started at 10 bar and was increased by another 5 bar every two minutes until the final value of 30 bar was reached. During the pressing procedure, a 100 μm thick spacer was in place to limit the minimum thickness of the sample to that value. For cross-linking, both sides of the obtained homogeneous polymer layer were exposed to UV-A irradiation in a UVACube 100 for 5 min each. This resulted in a stable membrane that could be peeled off from the supporting foil.
All mentioned component amounts expressed in mol% are given relative to the amount of polymer monomers present in the mixture.
Characterization techniques
Nuclear magnetic resonance (NMR) spectroscopy
NMR measurements were performed employing an AVANCE III HD spectrometer (BRUKER, USA) and a broadband probe (PA BBO 400 MHz, BRUKER) at a frequency of 100.61 MHz (13C) or 128.38 MHz (11B). All samples were solvated in DMSO-d6, and all measurements were executed at a temperature of 293.3 K.
X-ray Diffraction (XRD)
A Bruker D8 ADVANCE device (BRUKER, USA) was used to execute powder X-ray diffraction (XRD) measurements. The experiments were performed in step scan mode at a step size of 2θ = 0.02° and a step time of 1 s per step with a range from 2θ = 10° to 80°. A copper target was employed as the source of the Kα radiation with a wavelength of 1.54 Å used for the measurement.
Scanning electron microscopy (SEM)
SEM images were acquired using a Carl Zeiss AURIGA Scanning Electron Microscope (Carl Zeiss Microscopy GmbH, Jena, Germany). Images were recorded using magnification factors between 100 × and 2500 × using an accelerating voltage of 3 kV and a working distance of 3.8 mm. An In-Lens detector was used to acquire the images.
Energy-dispersive X-ray spectroscopy (EDS)
EDS was executed using an Oxford Instruments XMax silicon drift detector (Oxford Instruments plc, Abingdon, UK) that is attached to the above-mentioned Carl Zeiss AURIGA SEM device. The elemental mapping was done at a fixed magnification factor of 250 × using a working distance of 5.0 mm. The applied accelerating voltage was 20 kV.
Thermogravimetric analysis (TGA)
A TA Instruments TGA Q5000 device (TA Instruments, New Castle, USA) was used for thermogravimetric analysis (TGA) of the synthesized BOB salts. The samples were heated from 50 °C to 400 °C at a heating rate of 10 °C per minute. During operation, the samples were flushed with nitrogen gas at a constant gas flow rate of 10 ml min–1.
Electrochemical Measurements
Transference number measurements
To measure the lithium ion transference number tLi+ of the polymer membranes according to the method popularized by Bruce and Vincent, two electrode coin cells (CR2032) were constructed using circular disks of lithium foil with a diameter of 15 mm as both working and counter electrode. 67–71 A 12 mm wide circular disk that was punched from the polymer membranes was placed centrally between the lithium electrodes. A bo-PET-based separator ring with inner and outer diameters of 12 mm and 16 mm, respectively, was placed around the polymer disks to avoid any direct contact between the two lithium electrodes. The cells were then connected to a Bio-Logic VMP3 potentiostat inside a BINDER MK53 climate chamber set to 60 °C and rested for 24 h. After this, the cells were analyzed with electrochemical impedance spectroscopy (EIS) in a frequency range from 1 MHz to 100 mHz. A constant voltage ΔV = 0.01 V was applied to the samples for 24 h and the occurring polarization current was recorded. Finally, another impedance measurement was executed directly after removing the polarization voltage. The tLi+ was calculated from the values acquired in these measurements according to Eq. 1: 69
In this equation, I0 is the initial polarization current after applying the voltage and ISS represents the current measured after a steady state has been reached 24 h later. Similarly, R0 represents the initial electrode resistance before the start of the polarization and RSS is the steady state electrode resistance at the end of the measurement. These resistances were acquired from the recorded impedance spectra. An example of this calculation can be found in the Supporting Information (Fig. S1 (available online at stacks.iop.org/JES/168/070559/mmedia)). Typically, this procedure was executed on four identically composed cells and the acquired results were averaged to obtain the final values presented herein. The same coin cell geometry was also used to determine the chemical stability of the created membranes when in contact with lithium metal. After construction, the respective cells were kept under OCV conditions at a temperature of 60 °C for 90 h. During this time, the impedance values were repeatedly measured in a frequency range between 1 MHz and 100 mHz. All impedance values were determined as R2 + R3 as shown in the equivalent circuit in Fig. S2.
Ionic conductivity measurements
To investigate the ionic conductivity of the polymer samples, 12 mm wide membrane disks were placed between two stainless steel disks with a diameter of 15 mm in coin cells (CR2032). A separator ring was placed around the polymer disks to eliminate unwanted contact between the stainless steel electrodes. After resting for 24 h, the cells were connected to a Novocontrol impedance spectrometer equipped with an AN-alpha analyzer, a POT/GAL 20/11 electrochemical testing station, and a Novocontrol Quatro cryosystem temperature control unit. The samples were first heated from 20 °C to 80 °C and then cooled down again to 20 °C in steps of 10 °C. Three impedance measurements were executed at each temperature step, with resting intervals of 15 min between the individual measurements. From these measurements, the total ionic conductivity (σtotal) values of the polymer electrolyte membranes were calculated. Furthermore, the σLi+ of the samples was calculated by directly multiplying tLi+ with the σtotal obtained for a temperature of 60 °C. As with the transference number measurements, four identical cells were used for each batch and the final results represent averages of the data acquired from each of the four systems.
Results and Discussion
To confirm the proposed cation displacement hypothesis, a MEEP/LiBOB-based SPE was doped with different amounts of the in-house synthesized BOB-salt of potassium to trigger the predicted change of the lithium ion coordination situation. The modified polymer electrolytes were then investigated regarding their lithium ion transference number and ionic conductivity to check for potential increases caused by the cation addition. The same investigations were also applied to polymer blends based on PEO combined with LiBOB and KBOB. PEO serves as a model for an isolated look on the side chains of MEEP as it has the very same structure in its polymer backbone without sharing MEEP's poly(phosphazene) compartment. Therefore, any phenomenon that influences MEEP's side chains should have a comparable effect on PEO and vice versa. Consequently, if doping SPEs based on MEEP and PEO with KBOB improves the electrochemical performance of the MEEP membranes without substantially changing the properties of the treated PEO, it can be concluded that the added potassium ions indeed act mainly on the unique MEEP backbone as predicted by the presented cation displacement model rather than significantly affecting MEEP's side chains.
Salt analytics
All experiments were carried out using alkali metal BOB salts as both conducting agent and additive instead of other typical alkali metal salts due to the reportedly high interfacial compatibility of MEEP polymers with LiBOB content towards lithium metal as well as LiBOB's positive influence on the formation of an effective SEI. 72–76 A known synthesis route for LiBOB was modified to prepare KBOB in this work, which has not been described in literature thus far. 67 The synthesized KBOB was analyzed with 13C and 11B NMR spectroscopy to confirm the successful synthesis; the spectra are depicted in Fig. 4. The 11B NMR spectrum shows only one sharp boron signal which proves that all the boron supplied to the synthesis reaction through the H3BO3 was successfully integrated into the bis(oxalato)borate complexes and no further boron compounds were formed. This is further backed up by the absence of a signal at a chemical shift of δ = 19.4 ppm that would indicate remaining H3BO3 according to literature. 77 Similarly, the depicted 13C NMR spectrum shows only one product signal at δ = 158.4 ppm, with the only other visible feature of the spectrum being the solvent signal of the used DMSO-d6 at δ = 39.5 ppm. The observed NMR shifts are in good agreement with literature data recorded for BOB salts. 78,79 The XRD spectrum in Fig. 4c that was obtained from the synthesized KBOB is typical for this compound. The recorded diffraction pattern matches well with literature reported data, as denoted in the graph. 80 The absence of surplus reflexes in relation to the literature pattern indicates a high purity of the analyzed sample. Furthermore, the TGA analysis of the synthesized KBOB (Fig. S3) shows a decomposition onset at ≈330 °C which is in good agreement with values reported earlier. 79 Judging from the aforementioned analytical results, it can be concluded that the presented synthesis route represents a reliable yet simple way to obtain KBOB. In a similar fashion, it was confirmed that NaBOB and CsBOB are also accessible via the described experimental route; TGA results for these compounds can be found in the Supporting Information. Since potassium ions were found to be most favorable for the desired displacement of lithium ions from the trapping coordination sites at the MEEP backbone via cation doping (see Fig. 3), all further experiments were focused on the usage of KBOB only.
Chemical stability
The chemical stability of the prepared KBOB containing MEEP-based SPE systems towards lithium metal is investigated by monitoring the evolution of the impedance values of symmetric Li│MEEP-SPE│Li cells over an extended period of time. The impedances acquired for cells containing polymer membranes with 0 mol% and 2.5 mol% of KBOB are compared in Fig. 5. As can be seen, the bulk resistance of the MEEP-based SPEs is the same for both compositions, remaining unchanged over 90 h of storing the cell at 60 °C. The KBOB containing cell obviously shows the same impedance evolution pattern as the reference cell, achieving continuously lowered impedance values for 70 h of resting. After 70 h, all equilibration processes inside the membrane systems are finished and the impedance values for both cells remain stable. This confirms that the addition of KBOB to MEEP-based SPE membranes does not impair their stability against lithium metal.
Lithium ion transference numbers
The mobility of lithium ions in the polymer membranes containing various amounts of KBOB alongside LiBOB was investigated by determining the tLi+ of the samples. According to the cation displacement hypothesis proposed herein, one can expect an increase of the lithium ion mobility when KBOB is added to the MEEP-based membranes, which should lead to a distinct increase of the tLi+. This would verify the proposed change in the lithium coordination situation upon addition of further alkali cations to the MEEP polymer. For comparison, the same measurements were also executed with KBOB-doped PEO membrane systems. Since PEO is exclusively composed of an ether structure with carbon and oxygen atoms, the comparison of systems based on MEEP and PEO can help to verify that the observed effects are indeed caused by the typical MEEP backbone structure.
Figure 6 shows the tLi+ values of SPE membranes based on MEEP and PEO that contained between 0 mol% and 10 mol% of KBOB. MEEP SPE samples without KBOB showed a low transference number of ≈0.08, which is in line with literature data. 74 When 2.5 mol% of KBOB were included in the MEEP mixture the tLi+ value drastically increased to 0.18. No other studies have reported comparably high tLi for MEEP-based SPEs to this date, and these values are rather impressive for an all-solid-state MEEP electrolyte system without the addition of plasticizers. However, further enhancements of the KBOB content lead to a steady decline in tLi+ down to about 0.125. This leads to the conclusion that when comparing the composition ratios investigated in this work the best prerequisites for the movement of lithium ions through the MEEP-based SPE material are present at a KBOB content of 2.5 mol%, which equals to a molar potassium-to-lithium ratio of ≈0.17. According to the cation displacement mechanism introduced earlier, it is expected that the potassium ions used for the cation doping are located in the backbone coordination pockets of MEEP. This turns many formerly trapped lithium ions into mobile ions, which can explain the rise in the tLi+ value. Furthermore, the lithium ion mobility is likely also boosted by the effect of KBOB on the inner structure of the MEEP. Both the potassium cation and especially the BOB anion are sterically demanding, which could lead to a widening of the polymer network. This would increase the free volume and allow for an easier movement of the small lithium ions through the polymer matrix, similar to what has been reported for other polymer electrolytes. 81,82 On the other hand, upon increasing the KBOB content in the polymer mixture, the overall share of free lithium ions among all the mobile ion species gets lower until the point where the lithium ion movement through the material could be outweighed by the mobility of the other ions. This way, at higher KBOB concentrations the contribution of the lithium ions to the overall ion mobility would be lowered, which would then be reflected in declining tLi+ values. Also, at high salt concentrations, the ion pairing tendency in the SPE is likely increased, which should lead to a higher rate of neutral lithium salt molecules that do not contribute to tLi+ anymore.
In comparison to the drastic reaction presented for the MEEP membranes, the tLi+ of the PEO-based membranes shows no significant changes upon the addition of KBOB, as displayed in Fig. 6. All investigated PEO-based membrane compositions, including the one without any KBOB, feature tLi+ values of ≈0.03 to 0.04. In relation to MEEP, the different effect of KBOB on tLi+ of the PEO-based membranes is because the structure of PEO does not feature MEEP-like nitrogen-based backbone coordination slots with strong lithium ion trapping properties. Therefore, no lithium cations can be set free upon the addition of KBOB, as has been reasoned for MEEP. The difference between the tLi+ curves of MEEP and PEO caused by the structural difference between the two polymers confirms the validity of the cation displacement model that was the starting point of this work.
Total ionic conductivity
The overall ionic conductivities σtotal of SPE membranes based on MEEP or PEO with KBOB contents between 0 mol% and 10 mol% are displayed in Fig. 7. Aside from an expected change with increasing temperature, an influence of the KBOB content on the σtotal values is discernible. Among the MEEP-based samples investigated herein, the highest σtotal values are recorded for a KBOB content of 2.5 mol%. Further additions of KBOB lead to a continuous decrease in σtotal values with the result that samples containing 10 mol% of the salt additive show a lower σtotal than the membranes with no KBOB present. Similar to what has been discussed for the transference numbers, the initial boost of the overall ionic conductivity is due to the release of free lithium ions from the trapping coordination sites at the MEEP backbone upon the addition of KBOB. The overall increase in mobile charge carriers when KBOB is added also contributes to the σtotal value. Adding to this, the aforementioned widening of the polymer structure in reaction to KBOB content allows for a generally easier ion transport, especially with regard to the low size and weight of the released lithium ions that makes for their particularly swift movement through the widened polymer network. On the other hand, the reduced KBOB dissociation rate at higher salt contents leads to surplus salt mostly residing in the polymer matrix in a neutral and undissociated state. This KBOB is not contributing to the ionic conductivity and even hampers the movement of the other ions by blocking the diffusion pathways in the polymer, which explains the reduced σtotal values at higher KBOB contents. Moreover, with high amounts of KBOB, undissociated salt molecules tend to aggregate and form needle-like crystals with a length of up to 100 μm, as can be seen in the SEM and EDS images in Figs. 8a–8d. The EDS-based visualization of potassium in the sample with 10 mol% of KBOB shows that these crystals are rich in potassium, which confirms that they represent local agglomerations of KBOB. Furthermore, the phosphorus signals in Fig. 8b are diminished at the mentioned locations. Since phosphorus is present only as part of the MEEP backbone in the examined system, its absence at the places of high potassium content implies that the KBOB crystals interrupt the local polymeric structure. Consequently, the ion movement through these areas is blocked, which further contributes to the decrease in the σtotal values. In contrast, Figs. 8e–8h show that a MEEP-based membrane with 2.5 mol% of KBOB is smooth and does not feature any KBOB agglomerations or other irregularities.
While the PEO-based SPE membranes show a stronger temperature influence on their σtotal, they seem to perform best when used without KBOB. Any addition of the salt to the polymer formulation leads to reduced σtotal values of the produced membranes. As discussed before, this is largely due to the absence of trapped lithium ions that could be set free via the stated ion displacement mechanism to facilitate an increase in ion mobility. Instead, the increased salt loading with more and more KBOB being added leads to a higher ion pairing tendency which effectively reduces the general ion mobility through the samples, as has been reasoned earlier. Also, the share of potassium ions among the mobile charge carriers gradually increases with higher KBOB contents, which leads to a decline in overall ion mobility since potassium ions are significantly heavier and thus less mobile compared to lithium ions.
Partial lithium ion conductivity
By multiplying the recorded tLi+ and σtotal values, σLi+ of the analyzed SPE samples is obtained. The σLi+ curves displayed in Fig. 9 illustrate the effective transportability of lithium ions through the polymer membranes. Without added KBOB, the samples based on MEEP show a σLi+ of ≈0.02 mS cm–1, which is just above the values exhibited by the PEO-based membranes. However, while the membranes based on PEO feature an unchanged or even slightly diminished performance with different amounts of KBOB added, the σLi+ value of the MEEP system is increased by a factor of four upon the incorporation of 2.5 mol% of KBOB. The measured σLi+ of ≈0.08 mS cm–1 is almost three times as high as literature values of 0.028 mS cm–1 measured at the same temperature, which represent the state-of-the-art for MEEP SPEs with LiBOB as the conducting salt. 74 On the other hand, increasing the KBOB content of the MEEP-based membranes to 5 mol%, 7.5 mol%, and 10 mol% leads to a gradual performance decline down to 0.024 mS cm–1 at 10 mol% of KBOB. Since σLi+ is calculated directly from tLi+ and σtotal, the obtained changes of the values are due to the same phenomena as already discussed before. It is concluded that the σLi+ of MEEP-based SPE membranes can indeed be effectively boosted by adding small amounts of KBOB, which might lead to an increased performance in a battery application scenario, which will be investigated in future research. By comparing the observed performance increase to the fundamentally different response of PEO-based membranes upon the addition of KBOB, the existence of the proposed cation displacement mechanism induced by alkali ion doping is verified.
Conclusions
A cation doping method was introduced in this work to overcome the comparably low lithium ion mobility within MEEP-based SPE membranes. Based on the known occurrence of two distinct lithium coordination sites with differing coordination strength in the polymer system, we proposed to displace the lithium cations from the strongly coordinating sites along the polymeric backbone toward the less mobility restraining side chain coordination sites. Using DFT calculations, we determined that the best way to achieve this goal is to dope the SPE mixture with potassium ions that have a high preference to be coordinated at the backbone sites and could thus expel the lithium ions from these places. This approach was realized by adding various amounts of in-house synthesized KBOB to mixtures of MEEP and LiBOB. The lithium ion transference number tLi+ as well as the total ionic conductivity σtotal and the partial lithium ion conductivity σLi+ of the SPE membranes were found to benefit greatly from this addition of KBOB. Mixtures with 2.5 mol% of KBOB showed a σLi+ of 0.08 mS cm–1, which is superior to all other similar MEEP-based SPE formulations known to date. As expected, a cross-check using PEO-based LiBOB-containing electrolyte membranes revealed no such positive response to KBOB doping, which is due to the absence of a poly(phosphazene) backbone. Highlighting the usefulness of the obtained structural understanding, this cross-check confirmed the validity of the proposed cation displacement model in the investigated systems. The obtained results emphasize the effectiveness of the described modifications in enhancing the lithium ion transportation properties of poly(phosphazene)-based SPE membranes. A more detailed analysis of MEEP-based SPEs with different KBOB contents between 0 and 5 mol% could be a very interesting starting point for further studies aiming to find the optimal composition to realize the best possible lithium ion mobility in such polymer electrolytes.
Acknowledgments
Financial support provided by the German Federal Ministry of Education and Research (BMBF) within the research projects "MEET Hi-EnD II" (03XP0084A), "AMaLiS" (03XP0125D), and "FestBatt" (03XP0174B) is gratefully acknowledged. Furthermore, the authors wish to thank Aleksei Kolesnikov for assistance with the XRD and SEM measurements, Annika Buchheit for support regarding the conductivity measurements, Rayan Guerdelli and Mariano Grünebaum for aiding in the MEEP synthesis and optimizing the synthesis procedure, Debbie Berghus for the TGA measurements, and Johannes Thienenkamp for acquiring the NMR measurements.