7
Contents lists available at ScienceDirect Solid State Ionics journal homepage: www.elsevier.com/locate/ssi Poly(benzyl methacrylate)-poly[(oligo ethylene glycol) methyl ether methacrylate] triblock-copolymers as solid electrolyte for lithium batteries Andreas Bergfelt a, , Laurent Rubatat b , Daniel Brandell a , Tim Bowden a a Department of Chemistry - Ångström Laboratory, Uppsala University, Box 538, SE-751 21 Uppsala, Sweden b CNRS/UNIV Pau & Pays Adour, Institut des Sciences Analytiques et de Physico-Chimie pour l´ Environnement et les Materiaux, UMR5254, 64000 Pau, France ARTICLE INFO Keywords: Solid polymer electrolyte Block copolymer Lithium-ion battery SAXS ATRP ABSTRACT A triblock copolymer of benzyl methacrylate and oligo(ethylene glycol) methyl ether methacrylate was poly- merized to form the general structure PBnMA-POEGMA-PBnMA, using atom transfer radical polymerization (ATRP). The block copolymer (BCP) was blended with lithium bis(triuoro methylsulfonate) (LiTFSI) to form solid polymer electrolytes (SPEs). AC impedance spectroscopy was used to study the ionic conductivity of the SPE series in the temperature interval 30 °C to 90 °C. Small-angle X-ray scattering (SAXS) was used to study the morphology of the electrolytes in the temperature interval 30 °C to 150 °C. By using benzyl methacrylate as a mechanical block it was possible to tune the microphase separation by the addition of LiTFSI, as proven by SAXS. By doing so the ionic conductivity increased to values higher than ones measured on a methyl methacrylate triblock copolymer-based electrolyte in the mixed state, which was investigated in an earlier paper by our group. A Li|SPE|LiFePO 4 half-cell was constructed and cycled at 60 °C. The cell produced a discharge capacity of about 100 mAh g 1 of LiFePO 4 at C/10, and the half-cell cycled for more than 140 cycles. 1. Introduction The electrolytes used today constitute a major threat to Li-battery safety [14]. Replacing ammable and harmful liquid electrolytes with solid polymer electrolytes (SPEs) would drastically improve this situa- tion. SPE materials are today well-known for energy storage applica- tions, but still suer from too low ionic conductivity for most com- mercial products. However, with an increase in operating temperature to 60 °C80 °C, which is not an obstacle for electric vehicles, their ionic conductivity increases drastically [57]. Energy storage materials made from polymers are therefore certainly an interesting option for these applications and the SPE area can be foreseen to experience a re- naissance during the coming decade(s) [810]. Increasing the ionic conductivity and their lithium ion transport number of said SPEs would further boost their use, since the main challenge is to combine mechanical stability with ionic conductivity. The main conduction mechanism in SPEs is the ionic transport through the segmental motion of the polymer main chain, which in this context constitutes a paradox: a more exible polymer can transport ions better, but then fails in terms of rigidity [11]. The popular use of low-T g polyethers, such as poly(ethylene oxide) (PEO), rely on the dissolution of lithium salt though the interaction of ether groups found in the main chain, and increasing the temperature increases the segmental motion and thus the ionic conductivity. One way to further increase the ex- ibility of the functional ether group, and simultaneously reduce the non-conductive crystalline domains, is to lace the ether groups outside the main chain as in a comb polymer [12]. Comb polymers hold the promise of higher ionic conduction since the mobility of the side chain is higher than for the polymer backbone. Here, the main chain could for example be a polyacrylate with oligoether side chains giving higher ionic conductivity due to increased exibility [13,14]. A problem faced by increasing the exibility of the polymer is that the material becomes softer and thus loose its mechanical properties, which also holds true for comb polymers. One way to overcome this problem is to use a block copolymer (BCP) design approach, where the blocks separate into dis- crete phases such as a soft ionic conducting phase and a hard me- chanically stable phase [1517]. BCP electrolytes constitute complex systems, comprising several variables in terms of polymer composition, salt concentration, and morphological organization. Yet, understanding phase behavior in BCPs with added salts, and how these phases behave over a wide temperature range, need to be understood in order to link ionic conductivity and mechanics requirements, since this is fundamental for the development of mechanically robust SPEs with good lithium ion conduction [18,19]. Phase separation becomes an important issue for systems where salt addition has the possibility to induce phase separation. While many https://doi.org/10.1016/j.ssi.2018.04.006 Received 22 February 2018; Received in revised form 3 April 2018; Accepted 4 April 2018 Corresponding author. E-mail address: [email protected] (A. Bergfelt). Solid State Ionics 321 (2018) 55–61 Available online 12 April 2018 0167-2738/ © 2018 The Authors. Published by Elsevier B.V. This is an open access article under the CC BY license (http://creativecommons.org/licenses/BY/4.0/). T

Poly(benzyl methacrylate)-poly[(oligo ethylene glycol ...uu.diva-portal.org/smash/get/diva2:1180064/FULLTEXT01.pdf100mAhg−1 of LiFePO 4 at C/10, and the half-cell cycled for more

  • Upload
    others

  • View
    3

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Poly(benzyl methacrylate)-poly[(oligo ethylene glycol ...uu.diva-portal.org/smash/get/diva2:1180064/FULLTEXT01.pdf100mAhg−1 of LiFePO 4 at C/10, and the half-cell cycled for more

Contents lists available at ScienceDirect

Solid State Ionics

journal homepage: www.elsevier.com/locate/ssi

Poly(benzyl methacrylate)-poly[(oligo ethylene glycol) methyl ethermethacrylate] triblock-copolymers as solid electrolyte for lithium batteries

Andreas Bergfelta,⁎, Laurent Rubatatb, Daniel Brandella, Tim Bowdena

a Department of Chemistry - Ångström Laboratory, Uppsala University, Box 538, SE-751 21 Uppsala, Swedenb CNRS/UNIV Pau & Pays Adour, Institut des Sciences Analytiques et de Physico-Chimie pour l´ Environnement et les Materiaux, UMR5254, 64000 Pau, France

A R T I C L E I N F O

Keywords:Solid polymer electrolyteBlock copolymerLithium-ion batterySAXSATRP

A B S T R A C T

A triblock copolymer of benzyl methacrylate and oligo(ethylene glycol) methyl ether methacrylate was poly-merized to form the general structure PBnMA-POEGMA-PBnMA, using atom transfer radical polymerization(ATRP). The block copolymer (BCP) was blended with lithium bis(trifluoro methylsulfonate) (LiTFSI) to formsolid polymer electrolytes (SPEs). AC impedance spectroscopy was used to study the ionic conductivity of theSPE series in the temperature interval 30 °C to 90 °C. Small-angle X-ray scattering (SAXS) was used to study themorphology of the electrolytes in the temperature interval 30 °C to 150 °C. By using benzyl methacrylate as amechanical block it was possible to tune the microphase separation by the addition of LiTFSI, as proven by SAXS.By doing so the ionic conductivity increased to values higher than ones measured on a methyl methacrylatetriblock copolymer-based electrolyte in the mixed state, which was investigated in an earlier paper by our group.A Li|SPE|LiFePO4 half-cell was constructed and cycled at 60 °C. The cell produced a discharge capacity of about100mAh g−1 of LiFePO4 at C/10, and the half-cell cycled for more than 140 cycles.

1. Introduction

The electrolytes used today constitute a major threat to Li-batterysafety [1–4]. Replacing flammable and harmful liquid electrolytes withsolid polymer electrolytes (SPEs) would drastically improve this situa-tion. SPE materials are today well-known for energy storage applica-tions, but still suffer from too low ionic conductivity for most com-mercial products. However, with an increase in operating temperatureto 60 °C–80 °C, which is not an obstacle for electric vehicles, their ionicconductivity increases drastically [5–7]. Energy storage materials madefrom polymers are therefore certainly an interesting option for theseapplications and the SPE area can be foreseen to experience a re-naissance during the coming decade(s) [8–10].

Increasing the ionic conductivity and their lithium ion transportnumber of said SPEs would further boost their use, since the mainchallenge is to combine mechanical stability with ionic conductivity.The main conduction mechanism in SPEs is the ionic transport throughthe segmental motion of the polymer main chain, which in this contextconstitutes a paradox: a more flexible polymer can transport ions better,but then fails in terms of rigidity [11]. The popular use of low-Tg

polyethers, such as poly(ethylene oxide) (PEO), rely on the dissolutionof lithium salt though the interaction of ether groups found in the mainchain, and increasing the temperature increases the segmental motion

and thus the ionic conductivity. One way to further increase the flex-ibility of the functional ether group, and simultaneously reduce thenon-conductive crystalline domains, is to lace the ether groups outsidethe main chain as in a comb polymer [12]. Comb polymers hold thepromise of higher ionic conduction since the mobility of the side chainis higher than for the polymer backbone. Here, the main chain could forexample be a polyacrylate with oligoether side chains giving higherionic conductivity due to increased flexibility [13,14]. A problem facedby increasing the flexibility of the polymer is that the material becomessofter and thus loose its mechanical properties, which also holds truefor comb polymers. One way to overcome this problem is to use a blockcopolymer (BCP) design approach, where the blocks separate into dis-crete phases such as a soft ionic conducting phase and a hard me-chanically stable phase [15–17].

BCP electrolytes constitute complex systems, comprising severalvariables in terms of polymer composition, salt concentration, andmorphological organization. Yet, understanding phase behavior in BCPswith added salts, and how these phases behave over a wide temperaturerange, need to be understood in order to link ionic conductivity andmechanics requirements, since this is fundamental for the developmentof mechanically robust SPEs with good lithium ion conduction [18,19].Phase separation becomes an important issue for systems where saltaddition has the possibility to induce phase separation. While many

https://doi.org/10.1016/j.ssi.2018.04.006Received 22 February 2018; Received in revised form 3 April 2018; Accepted 4 April 2018

⁎ Corresponding author.E-mail address: [email protected] (A. Bergfelt).

Solid State Ionics 321 (2018) 55–61

Available online 12 April 20180167-2738/ © 2018 The Authors. Published by Elsevier B.V. This is an open access article under the CC BY license (http://creativecommons.org/licenses/BY/4.0/).

T

Page 2: Poly(benzyl methacrylate)-poly[(oligo ethylene glycol ...uu.diva-portal.org/smash/get/diva2:1180064/FULLTEXT01.pdf100mAhg−1 of LiFePO 4 at C/10, and the half-cell cycled for more

different BCP SPE systems have been evaluated throughout the years,the most studied system so far is the polystyrene-polyethylenoxide (PS-PEO) copolymers where ionic conductivity, mechanical properties andmicro phase separation have been studied in depth [20–23].

One drawback with working with PEO is that it's SPEs are generallysemi-crystalline below 60 °C, which has a negative impact on the ionicconductivity [24]. One way to avoid this is to move the PEO func-tionality to a side-chain in the polymer, thus creating a comb copo-lymer. We have previously synthesized and studied two such combcopolymer systems in order to study the effect of random copolymer-ization compared to block copolymerization, where both the ionicconductivity and the miscibility of the blocks were studied [12,25]. Aninteresting finding was that the BAB block copolymer, where B is me-thyl methacrylate and A oligo(ethylene glycol) methyl ether metha-crylate, did not microphase separate. From an application point of viewthis is not desirable, since the hypothesis is that by separating hydro-phobic and hydrophilic constituents on the same polymer chain, a localnano-scale ordering of these segments with different polarity can berealized, promoting both ionic conductivity and mechanical stability.This can, in turn, result in better power performance of the batteries.

To better understand the microphase separation in SPEs and theirelectrochemical and battery performance, we have here synthesized anew type of BAB triblock copolymer based on benzyl methacrylate andoligo(ethylene glycol) methyl ether methacrylate, PBnMA-POEGMA-PBnMA, see Fig. 1. The electrolytes were prepared with bis(trifluoromethylsulfonate) (LiTFSI) to form SPEs. The best performing electrolytewas evaluated with SAXS and TEM to study the morphology, and abattery device was constructed to evaluate the electrochemical per-formance.

2. Experimental section

2.1. Materials

Materials used were oligo(ethylene glycol) methyl ether methacry-late (OEGMA, Mw=500 gmol−1, Sigma), benzyl methacrylate (BnMA,Sigma), dichloromethane (DCM, Fischer Scientific), diethyl ether(Fischer Scientific), cyclohexane (Acros Organics), ethanol (Solveco),CuBr (Sigma), CuBr2 (Sigma), 2,2′-bipyridyl (Sigma), ethylene glycol(Sigma), α-bromoisobutyryl bromide (Sigma), basic Al2O3 (Sigma), drytetrahydrofuran with molecular sieves (THF, Acros Organics), andCDCl3 (Larodan Fine Chemicals). Solvents were used without furtherpurification. Lithium bis(trifluoromethane)sulfonimide (LiTFSI,Purolyte, Ferro Corporation) was dried at 120 °C for 24 h before use.

2.2. Initiator synthesis

The synthesis of the di-functional initiator was adapted from lit-erature [28,29]. 18.6 mL (82.5mmol) α-bromoisobutyryl bromide wasadded to 2mL (35.6mmol) ethylene glycol and 11.5mL (82.5mmol) oftri-ethyl amine in 100mL dry THF using a dropping funnel under argonatmosphere. The reaction was cooled in an ice bath. The reaction wasleft overnight before the salt was filtered off with a Buchner funnel. THFwas evaporated and the solid dissolved in DCM (100mL) and washedthree times with saturated NaHCO3 (100mL). The organic phase wasdried with MgSO4, filtered and evaporated to give a white solid. Theproduct was recrystallized from ethanol and dried in a vacuum oven togive the desired product 2-(2-bromoisobutyryloxy)ethyl methacrylateas white needles, hereafter denoted di-EBiB. 1H NMR (400MHz, CDCl3)δ 4.42–4.44 (4H), 1.95–1.91 (12H).

2.3. Synthesis POEGMA macroinitiator

The OEGMA monomer was passed through a column of basic Al2O3

in order to remove the radical inhibitor. The monomer (OEGMA,31.63mL, 63.3 mol), solvent (ethanol, 30mL), initiator (di-EBiB,123mg, 0.44mmol), CuBr2 (9.2 mg, 0.0041mmol), and ligand (bpy,213mg, 1.36mmol) were added to a 100mL Schlenk flask. The flaskwas sealed with a silicone septum, degassed, backfilled three times withN2, and then left under N2. CuBr (98mg, 0.68mmol) was then added,and the Schlenk flask was placed in an oil bath at 60 °C for 2 h. Thesystem was quenched with acetone, filtered through basic Al2O3, andprecipitated twice in 300mL of a 1:1 mixture of diethyl ether and cy-clohexane. The solvents were removed using rotary evaporation and thefinal product was dried in a vacuum oven. The typical yield is circa30 wt% (POEGMA: Mn, GPC=31,085 gmol−1, PDI= 1.13).

2.4. Polymer synthesis triblock copolymer

BnMA was passed through a column with 10mL basic Al2O3 to re-move the radical inhibitor. The monomer (BnMA, 4.94mL, 46.2 mmol),solvent (ethanol, 6 mL), macroinitiator (POEGMA, 1.54 g), CuBr2 (3 mg,22.3 μmol), and PMDETA (29mg, 0.19mmol) were added to a 50mLSchlenk flask and three freeze-pump-thaw cycles were performed withN2. CuBr (36.8 mg, 0.37mmol) was added before the flask was sealedwith a silicone stopper and placed in an oil bath at 60 °C for 20min. Thereaction was then quenched with acetone, filtered through 20mL ofbasic Al2O3, and precipitated in 300mL of a 1:1 mixture of diethyl etherand cyclohexane. The solvents were removed using rotary evaporationand the final product was dried in a vacuum oven. (PBnMA-POEGMA-PBnMA: Mn, GPC=52,810 gmol−1, PDI= 1.22).

Fig. 1. Synthesis scheme for the triblock copolymer, using ATRP.

A. Bergfelt et al. Solid State Ionics 321 (2018) 55–61

56

Page 3: Poly(benzyl methacrylate)-poly[(oligo ethylene glycol ...uu.diva-portal.org/smash/get/diva2:1180064/FULLTEXT01.pdf100mAhg−1 of LiFePO 4 at C/10, and the half-cell cycled for more

2.5. GPC

An Agilent 1260 Infinity GPC was used to measure the molecularweight and PDI. The GPC was fitted with PolyPore columns and an RIdetector. The mobile phase was DMF with 50mM LiBr (1mLmin−1)operated at 80 °C. PMMA standards were used to calibrate the system.

2.6. Polymer electrolyte preparation

The polymers were mixed with dry THF and LiTFSI (ca 0.15 g ofpolymer and 2mL THF with the corresponding amount of LiTFSI). Thesamples were stirred overnight in an argon-filled glove box. The filmswere cast in Teflon molds, and the solvent was removed via controlledevaporation in a Büchi oven. The pressure was reduced to full vacuumover 20 h before heating at 60 °C at full vacuum for 40 h. The resultingfilms were 20mm in diameter and ca 0.2 mm in thickness.

2.7. DSC

A TA Instruments DSC Q2000 was used. The samples were herme-tically sealed in aluminum pans in a N2 glove box. The samples wereramped down to −90 °C, heated to 150 °C and then ramped down to−90 °C a second time before finally being ramped to 150 °C, with thespeed 10 °Cmin−1 under a flow of N2. The glass transition temperatureswere obtained from the second heating scan.

2.8. SAXS

Small Angle X-ray Scattering (SAXS) data were collected on theBL11-NCD beamline at ALBA (Barcelona, Spain). The X-ray wavelengthused was 0.99 Å, with a sample-to-detector distance of 2.9m for theSAXS. In the present paper, the scattering curves are plotted as afunction of the scattering vector q defined as q=4π/λ.sin(θ/2), with θbeing the scattering angle. Standard data corrections were applied in-cluding background subtraction. The samples were prepared as de-scribed in the casting step and analyzed wrapped in aluminum foil.

2.9. TEM

A JEOL JEM1400Plus transmission electron microscopy equippedwith a Ruby camera operating at 100 kV was used to study the elec-trolyte films. The samples were attached to a silver pin and cooled inliquid nitrogen. Sections (80–100 nm) were prepared at −40 °C and− 80 °C using an ultramicrotome Leica EMFC7 and collected on a for-mvar/carbon coated copper grid. The film sections were stained with1 wt% solution of uranyl acetate in water for 10min.

2.10. AC impedance

The films were placed between two stainless steel electrodes andsealed in a Swagelok cell. Film preparation and cell assembly wereperformed in an argon glove box. The films were annealed at 90 °C for1 h in order to achieve good contact between the sample and thestainless-steel electrodes. Impedance spectroscopy was measured with aSI 1260 Impedance Gain-Phase Analyzer (Schlumberger) at a frequencyrange of 1 Hz to 10MHz with the amplitude set to 10mV. The re-sistance was received with ZWiev using a modified Randles' circuit.

2.11. Cyclic voltammetry

Cyclic voltammetry analysis was carried out on a VMP2 (Bio-Logic).The cells were prepared with a 20mm disc of polymer electrolytesandwiched between a 14mm lithium disc and a 18mm stainless-steeldisc. The stack was sealed in a pouch bag with aluminum and coppercurrent collectors. All work was carried out in an argon filled glove box.The cell was cycled at 60 °C and at 1mV s−1.

2.12. Battery assembly

LiFePO4 electrodes were prepared with a mixture of 75 wt% activematerial, 10 wt% carbon black and 15wt% PVdF on aluminum foil. Theslurry was prepared with NMP as solvent. The electrodes were dried at120 °C for 12 h before cell preparation. The electrolytes (about 100mgpolymer) were casted on the cathodes and vacuum dried for 20 h beforeheating at 60 °C at full vacuum for 40 h, before assembled with lithiumanodes and sealed in pouch cells with aluminum and copper currentcollectors. The cathodes were 14mm in diameter, loaded with about2.5 mg cm−2 of active material.

2.13. Battery cycling

Galvanostatic cycling was carried out on a Digatron MBT batterytest system at 60 °C between 2.7 and 4.2 V versus Li+/Li at C/10. Thecell was annealed at the OCV at 60 °C for 6 h before cycling started.

3. Results and discussion

3.1. Polymer synthesis

The block copolymer was synthesized with atom transfer radicalpolymerization (ATRP). A POEGMA macroinitiator was synthesizedfrom a bifunctional initiator, ethylene bis(2-bromoisobutyrate), andoligo(ethylene glycol) methacrylate (OEGMA). In the second step, afterpurification and a drying step, benzyl methacrylate was used to producethe second block, see Fig. 1. The polymerization time for the POEGMAmacroinitiator was 2 h, and 20min for the benzyl blocks. GPC analysisgave that the macroinitiator molecular weight, Mn, was 31,085 gmol−1

with a PDI of 1.13, and that the final synthesis resulted in a BCP with aPDI value of 1.22 and a Mn of 47,830 gmol−1. The block compositionwas thus 35 wt% PBnMA. The formed BCP was rubbery and trans-parent.

3.2. DSC

All characterized samples in the article were prepared in same way:BCP, LiTFSI and THF were mixed and cast in Teflon molds, and thesolvent was removed via controlled evaporation in a Büchi oven. Thepressure was reduced to full vacuum over 20 h before heating at 60 °C atfull vacuum for 40 h. The resulting films were 20mm in diameter andabout 0.2mm in thickness. The formed electrolytes were transparent,non-sticky and mechanically stable films that were easy to handle.Differential scanning calorimetry (DSC) analysis was run with theramping sequence cooling/heating/cooling/heating in order to in-vestigate both the electrolytes pre-history and the thermal historyproduced by the DSC measurements. The Tg increased with the amountof LiTFSI, as expected, since the addition of LiTFSI slows down the localchain movements, resulting in a hardening of the electrolytes comparedto the salt-free BCP. The salt-free BCP showed only one clear Tg at about−40 °C on both the first and second heating scans, see Fig. 2 andTable 1. The PBnMA homopolymer has a Tg of about 55 °C and thePOEGMA homopolymer about −59 °C, see Table 1. However, if theFlory-Fox is applied the Tg of the blend should be about −29 °C, whichdeviates from the DSC value of −40 °C. Nevertheless, as no second Tg

could be found for the PBnMA block, this strongly indicates a lack of afull phase separation between the blocks, which also is confirmed bySAXS, see supporting information. With the addition of salt, however, aclear second Tg appears, see Fig. 2. The Tg signal becomes more pro-nounced with the addition of salt, which indicates that the phase se-paration is triggered by the amount of added salt. On the secondheating scan, these high temperature Tg:s were smeared out and hardlydetectable. Simultaneously, the lower Tg increases for the three sam-ples, suggesting a mixing of the blocks. This hysteresis points out thecomplexity of the microphase separation and that annealing time and

A. Bergfelt et al. Solid State Ionics 321 (2018) 55–61

57

Page 4: Poly(benzyl methacrylate)-poly[(oligo ethylene glycol ...uu.diva-portal.org/smash/get/diva2:1180064/FULLTEXT01.pdf100mAhg−1 of LiFePO 4 at C/10, and the half-cell cycled for more

temperature are crucial parameters to control.

3.3. Ionic conductivity

The ionic conductivity was evaluated with ac impedance spectro-scopy, see Fig. 3. The SPE conductivities were fitted to the Vogel-Tammann-Fulcher (VTF) equation, which is applicable for most amor-phous solid polymer electrolytes. The VTF relationship assumes that theionic conductivity is coupled to the segmental motion of the polymerchains, and thus coupled to the Tg of the electrolyte. The VTF equationis an empirical relationship:

= ⎡⎣⎢

−−

⎤⎦⎥

−σ A T exp Ek (T T )σ

1/2 σ

B 0 (1)

where Eσ is a pseudo-activation energy, kB is the Boltzmann constant,Aσ is a constant proportional to the number of ion carrier, and To is areference temperature usually associated with the ideal glass transitiontemperature at which the configurational entropy becomes zero(usually said to be ca 50 K below Tg). As expected, the fitting shows thatthe SPEs exhibit VTF behavior, since the SPEs are amorphous modifiedPEO systems. DSC analysis shows that with the addition of LiTFSI the Tg

increases, as a result of slower local chain kinetics. This contradictioncould be explained with that the addition of free charge carriers over-weight the slower local chain kinetics contribution of the final ionicconductivity. It is interesting to notice that with only a slight increasefrom 10:1 to 8:1 in salt concentration the ionic conductivity increasessignificantly.

3.4. SAXS and TEM

Small-angle X-ray scattering (SAXS) was used to study the bulkmorphology and the temperature dependency of the microphase se-paration. The salt-free BCP did not show any scattering pronouncedscattering peaks, indicating that the BCP is in the mixed state (seeSupporting information, SI-1). However, with the addition of saltscattering peaks were observed. Fig. 4 shows the scattering data forsample 8:1, 10:1 and 20:1, collected at 30 °C before and after thetemperature annealing at 150 °C. Before annealing, the spectra showthree broad scattering peaks, and after the annealing the scatteringpeaks become smeared out for sample 10:2 and 20:1, however, sample8:1 becomes well defined. During heating all first order peaks q* aresignificantly shifted to smaller angles, sample 8:1 shifted from 0.160 to0.146 nm−1, sample 10:1 from 0.151 to 0.135 nm−1 and sample 20:1from 0.156 to 0.142 nm−1, meaning that the q*-positions versus tem-perature is not fully reversible (as shown by DSC). On all the spectrathree secondary order peaks are observed, and the peaks at 2q* and√7q* indicates a self-assembly into a hexagonal packing of cylinders,which is in agreement with the expected theoretical equilibrium mor-phology given by the triblock composition [26]. The cylinder-to-cy-linder distance, calculated from d= (4*π)/(√3q*), for the 8:1 sample is45.4 nm at 30 °C before and 49.2 nm at 30 °C after the cooling ramp,48.0 nm and 53.7 nm for sample 10:1, and 46.5 nm and 51.1 nm forsample 20:1. In the temperature range explored no order to disordertemperature (TODT) was reached (see SI-2 to SI-4 in Supplementaryinformation). The temperature annealing at 150 °C thus allowed the

Fig. 2. DSC traces showing the two heating cycles of the DSC program. It isclear that the SPEs are phase separated in the pristine state, but that the DSCannealing smears out the second Tg. The DSC traces are shifted vertically for thesake of clarity. The concentration are given as ethylene oxide:Li+ (EO:Li+)ratios: 8:1, 10:1 and 20:1.

Table 1DSC results for the triblock copolymer electrolyte series. Tg,1 and Tg,2 refers toheat scan one and two for the POEGMA block. Tg,B refers to the PBnMA blocks,which was recorded on heat scan one.

Entry wt% LiTFSI Tg,1 [°C] Tg,2 [°C] Tg,B [°C]

0 0 −43.9 −40.2 –20:1 14.7 −52.5 −45.6 55.510:1 29.5 −48.3 −43.1 55.88:1 36.8 −42.7 −35.7 59.08

Fig. 3. Ionic conductivity as a function temperature, from ac impedance analysis. The dashed lines represent VTF fits. The ratios resemble the EO:Li+ ratio.

A. Bergfelt et al. Solid State Ionics 321 (2018) 55–61

58

Page 5: Poly(benzyl methacrylate)-poly[(oligo ethylene glycol ...uu.diva-portal.org/smash/get/diva2:1180064/FULLTEXT01.pdf100mAhg−1 of LiFePO 4 at C/10, and the half-cell cycled for more

structure to rearrange itself, releasing any kinetically hindered lockingof the structure.

Transmission electron microscopy (TEM) was performed on theelectrolytes before the annealing step. The TEM images show mainlyelongated objects, but with some isotropic objects, see right image in

Fig. 5, red circle. These observations can be interpreted as cylindersforming a limited hexagonal packing, which could support the broadpeaks in the SAXS curves collected at 30 °C before the annealing step.

3.5. Electrochemical performance

Cyclic voltammetry (CV) showed that the electrolyte display in-stabilities over several cycles between 5 to −0.5 V, with a cell config-uration of lithium versus gold, see Fig. 6. The CV showed that the 8:1electrolyte display a rather unsteady current flow over the entire po-tential window, with a rather high current peak at about 1 to 2 V. It isclear that it is more than lithium stripping/plating happening (A inFig. 6) during the CV scans because the onset of the first peak is above0 V. It is also clear that the SPE is not electrochemically stable, since itstarts reducing already below 2 V on the first cycle. The current flow atabout 5 V is expected, since most electrolytes show instability at thesehigh potentials (C in Fig. 6). However, Tominaga and co-workers haveinvestigated ester-based electrolytes that show stability in this potentialregion due to high salt concentrations, which stabilizes the interfacewith the electrodes [30]. The most probable explanation to the shift inthe lithium stripping and plating potential is that the ester functionalityis redox active, which lies close to the Li-plating potential. There is aclear passivation seen in peak A since the current in the Li plating re-gime drops during cycling, indicating that the decomposition productsformed during the repeated cycles is passivating the electrode. How-ever, the main feature is the peak at about 2 V which grows significantlyon cycling, while the peak at about 0 V shrinks during cycling. It isunclear what this current peaks is an outcome of, but the most probableexplanation is that the SPE is degrading at the interfaces during cycling,generating a degradation product which also is redox active (B inFig. 6), but also other type of side products (D in Fig. 6). That comb-based methacrylate electrolytes show this type of electrochemical in-stability have been observed before, thus questioning the use of me-thacrylate based SPEs for lithium half-cell applications [27]. However,it is not clear weather this instability is an outcome of polymerizationtechnique (end group functionality), block type (methyl methacrylate,polystyrene, etc.) or cell design (gold or stainless steel vs. lithium).Battery cycling were investigated in a Li|SPE|LiFePO4 half-cell, usingthe 8:1 composition as electrolyte. The 8:1 electrolyte was solventcasted on the LiFePO4 electrode in order to ensure a good contact be-tween cathode and electrolyte. The half-cell was cycled at 60 °C at C/10, see Fig. 7. The practical capacity for LiFePO4 is usually about150mAh g−1 (theoretical 170mAh g−1), but these values was notrealized in these experiments. This, and the fact that the capacity fadedconstantly during cycling, could be explained by a high cycling over-potential in combination with parasitic side reactions triggered by theoperating temperature of 60 °C. This could be supported by the cyclicvoltammetry experiment which showed that the electrolyte is affectedby cycling over time. Despite showing a constant fade in capacity, thecell still managed to cycle about 140 cycles, proving the capability forthe electrolyte in comparison to similar systems that utilized methyl

Fig. 4. SAXS scattering profiles for sample 8:1, 10:1 and 20:1 before (left) andafter (right) annealing at 150 °C. The scattering profiles are shifted vertically forthe sake of clarity. q* represents the primary peak, pentagon represents 2q*,time-glass √7q* and triangle 4q*.

Fig. 5. TEM images of sample 8:1. The imagesclearly show that the main structure before the an-nealing step is mainly lamella, but with cylindricalregions, see red circle to the right. The left scale baris 500 nm and the right 200 nm. (For interpretationof the references to color in this figure legend, thereader is referred to the web version of this article.)

A. Bergfelt et al. Solid State Ionics 321 (2018) 55–61

59

Page 6: Poly(benzyl methacrylate)-poly[(oligo ethylene glycol ...uu.diva-portal.org/smash/get/diva2:1180064/FULLTEXT01.pdf100mAhg−1 of LiFePO 4 at C/10, and the half-cell cycled for more

methacrylate as the mechanical block, thus indicating that PBnMA is abetter battery device material than methyl methacrylate [12,25,27].The electrolyte did not form a smooth surface, but rather an unevenbubbly surface that may cause a bad contact with the lithium anode andthus be a limiting factor for utilizing the maximum capacity of theLiFePO4-based electrode.

4. Conclusion

A triblock copolymer electrolyte using benzyl methacrylate as me-chanical block and oligo(ethylene glycol) methyl ether methacrylate asionic conductive block was synthesized and evaluated as a SPE in alithium half-cell battery device. Since microphase separated BCP elec-trolytes have the ability to combine both mechanical and electro-chemical properties, they are of interest to study and evaluate bothfrom a morphological and electrochemical point of view. The salt freePBnMA based BCP did not phase separate into clear domains, but withthe addition of LiTFSI, a phase separation was triggered as demon-strated by SAXS and TEM. In the temperature range explored, 30 °C to150 °C, no order to disorder temperature (TODT) was reached; the SAXSprofiles showed secondary order peaks at √4q* and √7q*, indicating a

self-assembly into a hexagonal packing of cylinders with a typical cy-linder to cylinder distance of 45.4 to 53.7 nm. Since sample 8:1 showedthe highest ionic conductivity and clear microphase separation, it wasevaluated as an electrolyte in a Li|SPE|LiFePO4 half-cell. The batterydevice did not perform a steady cycling behavior, and the specific ca-pacity utilization was limited. Cyclic voltammetry indicates that theelectrolyte is electrochemically unstable at 60 °C, which could be aconsequence of that the electrolyte is generating an interface that isthermodynamically and/or kinetically unstable with the lithium metal,thus forming a non-effectively passivating interphase. The limitedelectrochemical stability of methacrylate based SPEs have been re-ported earlier, but the battery cycling behavior seems to be increasedwhen PBnMA rather than PMMA is used as mechanical block[12,25,27]. One way to improve the capacity utilization of LiFePO4 andreducing parasitic side reactions, is to use the electrolyte as a binder[27]. By doing so, the battery cycling performance could be improvedby reducing the mass transport limitations in the cathode which wouldlower the overall cell polarization and resistance. By doing so it shouldbe possible to cycle the battery at lower temperatures, which wouldmitigate any parasitic side reactions.

Fig. 6. Cyclic voltammetry of the 8:1 electrolyte in a cell setup up of lithium versus gold at 60 °C and with a cycling speed of 1 mV s−1.

Fig. 7. Battery cycling of a Li|SPE|LiFePO4 half-cell using the 8:1 electrolyte, operating at 60 °C and at C/10.

A. Bergfelt et al. Solid State Ionics 321 (2018) 55–61

60

Page 7: Poly(benzyl methacrylate)-poly[(oligo ethylene glycol ...uu.diva-portal.org/smash/get/diva2:1180064/FULLTEXT01.pdf100mAhg−1 of LiFePO 4 at C/10, and the half-cell cycled for more

Acknowledgements

We thank ALBA for providing SAXS beam time. We also would liketo thank Michaela Salajkova at Oslo University for the TEM images.This work has been supported by the Swedish Energy Agency (projectTriLi) and STandUP for Energy.

Appendix A. Supplementary data

Supplementary data to this article can be found online at https://doi.org/10.1016/j.ssi.2018.04.006.

References

[1] J.B. Goodenough, Y. Kim, Challenges for rechargeable Li batteries, Chem Mater 22(3) (2010) 587–603.

[2] J.M. Tarascon, M. Armand, Issues and challenges facing rechargeable lithium bat-teries, Nature 414 (6861) (2001) 359–367.

[3] J. Kalhoff, G.G. Eshetu, D. Bresser, S. Passerini, Safer electrolytes for lithium-ionbatteries: state of the art and perspectives, ChemSusChem 8 (13) (2015)2154–2175.

[4] P.G. Balakrishnan, R. Ramesh, T.P. Kumar, Safety mechanisms in lithium-ion bat-teries, J Power Sources 155 (2) (2006) 401–414.

[5] M. Armand, J.M. Tarascon, Building better batteries, Nature 451 (7179) (2008)652–657.

[6] B. Scrosati, J. Hassoun, Y.K. Sun, Lithium-ion batteries. A look into the future,Energy Environ Sci 4 (9) (2011) 3287–3295.

[7] K. Takada, Progress and prospective of solid-state lithium batteries, Acta Mater 61(3) (2013) 759–770.

[8] K. Murata, S. Izuchi, Y. Yoshihisa, An overview of the research and development ofsolid polymer electrolyte batteries, Electrochim Acta 45 (8–9) (2000) 1501–1508.

[9] L. Vandepaer, J. Cloutier, B. Amor, Environmental impacts of lithium metalpolymer and lithium-ion stationary batteries, Renew Sust Energ Rev 78 (2017)46–60.

[10] J. Janek, W.G. Zeier, A solid future for battery development, Nature Energy 1(2016) 4.

[11] Y.Y. Wang, F. Fan, A.L. Agapov, T. Saito, J. Yang, X. Yu, K.L. Hong, J. Mays,A.P. Sokolov, Examination of the fundamental relation between ionic transport andsegmental relaxation in polymer electrolytes, Polymer 55 (16) (2014) 4067–4076.

[12] M. Bergman, A. Bergfelt, B. Sun, T. Bowden, D. Brandell, P. Johansson, Graft co-polymer electrolytes for high temperature Li-battery applications, using poly(me-thyl methacrylate) grafted poly(ethylene glycol)methyl ether methacrylate and li-thium bis(trifluoromethanesulfonimide), Electrochim Acta 175 (2015) 96–103.

[13] W.-S. Young, W.-F. Kuan, T.H. Epps, Block copolymer electrolytes for rechargeablelithium batteries, J Polym Sci B Polym Phys 52 (1) (2014) 1–16.

[14] W.-F. Kuan, R. Remy, M.E. Mackay, I.I.I.T.H. Epps, Controlled ionic conductivity viatapered block polymer electrolytes, RSC Adv 5 (17) (2015) 12597–12604.

[15] G. Jo, H. Ahn, M.J. Park, Simple route for tuning the morphology and conductivity

of polymer electrolytes: one end functional group is enough, ACS Macro Lett 2 (11)(2013) 990–995.

[16] S. Inceoglu, A.A. Rojas, D. Devaux, X.C. Chen, G.M. Stone, N.P. Balsara,Morphology–conductivity relationship of single-ion-conducting block copolymerelectrolytes for lithium batteries, ACS Macro Lett 3 (6) (2014) 510–514.

[17] A.-V.r.G. Ruzette, P.P. Soo, D.R. Sadoway, A.M. Mayes, Melt-formable block co-polymer electrolytes for lithium rechargeable batteries, J Electrochem Soc 148 (6)(2001) A537.

[18] A.A. Teran, N.P. Balsara, Thermodynamics of block copolymers with and withoutsalt, J Phys Chem B 118 (1) (2014) 4–17.

[19] R.A. Farrell, T.G. Fitzgerald, D. Borah, J.D. Holmes, M.A. Morris, Chemical inter-actions and their role in the microphase separation of block copolymer thin films,Int J Mol Sci 10 (9) (2009) 3671–3712.

[20] M. Chintapalli, T.N.P. Le, N.R. Venkatesan, N.G. Mackay, A.A. Rojas, J.L. Thelen,X.C. Chen, D. Devaux, N.P. Balsara, Structure and ionic conductivity of polystyrene-block-poly(ethylene oxide) electrolytes in the high salt concentration limit,Macromolecules 49 (5) (2016) 1770–1780.

[21] K. Timachova, H. Watanabe, N.P. Balsara, Effect of molecular weight and saltconcentration on ion transport and the transference number in polymer electrolytes,Macromolecules 48 (21) (2015) 7882–7888.

[22] J.L. Thelen, A.A. Teran, X. Wang, B.A. Garetz, I. Nakamura, Z.-G. Wang,N.P. Balsara, Phase behavior of a block copolymer/salt mixture through the order-to-disorder transition, Macromolecules 47 (8) (2014) 2666–2673.

[23] M. Chintapalli, X.C. Chen, J.L. Thelen, A.A. Teran, X. Wang, B.A. Garetz,N.P. Balsara, Effect of grain size on the ionic conductivity of a block copolymerelectrolyte, Macromolecules 47 (15) (2014) 5424–5431.

[24] J. Mindemark, M. Lacey, T. Bowden, D. Brandell, Beyond PEO - Alternative hostmaterials for Li+-conducting solid polymer electrolytes, Progress in PolymerScience, https://doi.org/10.1016/j.progpolymsci.2017.12.004

[25] A. Bergfelt, L. Rubatat, R. Mogensen, D. Brandell, T. Bowden, d(8)-Poly(methylmethacrylate)-poly (oligo ethylene glycol) methyl ether methacrylate tri-block-co-polymer electrolytes: morphology, conductivity and battery performance, Polymer131 (2017) 234–242.

[26] Y. Funaki, K. Kumano, T. Nakao, H. Jinnai, H. Yoshida, K. Kimishima, K. Tsutsumi,Y. Hirokawa, T. Hashimoto, Influence of casting solvents on microphase-separatedstructures of poly(2-vinylpyridine)-block-polyisoprene, Polymer 40 (25) (1999)7147–7156.

[27] D. Devaux, D. Glé, T.N.T. Phan, D. Gigmes, E. Giroud, M. Deschamps, R. Denoyel,R. Bouchet, Optimization of block copolymer electrolytes for lithium metal bat-teries, Chem Mater 27 (13) (2015) 4682–4692.

[28] A.A. Kavitha, N.K. Singha, Smart “all acrylate” ABA triblock copolymer bearingreactive functionality via atom transfer radical polymerization (ATRP): demon-stration of a “click reaction” in thermoreversible property, Macromolecules 43 (7)(2010) 3193–3205.

[29] S. Karanam, H. Goossens, B. Klumperman, P. Lemstrat, “Controlled” synthesis andcharacterization of model methyl methacrylate/tert-butyl methacrylate triblockcopolymers via ATRP, Macromolecules 36 (9) (2003) 3051–3060.

[30] Kento Kimura, Joh Motomatsu, Yoichi Tominaga, Highly ConcentratedPolycarbonate-Based Solid Polymer Electrolytes Having ExtraordinaryElectrochemical Stability, Journal of Polymer Science, Polymer Physics (2016),http://dx.doi.org/10.1002/polb.24235.

A. Bergfelt et al. Solid State Ionics 321 (2018) 55–61

61