Elucidating the Solvation Structure and Dynamics of...

5
Elucidating the Solvation Structure and Dynamics of Lithium Polysuldes Resulting from Competitive Salt and Solvent Interactions Nav Nidhi Rajput,* ,,Vijayakumar Murugesan, ,Yongwoo Shin, Kee Sung Han, ,Kah Chun Lau, §,Junzheng Chen, Jun Liu, ,Larry A. Curtiss, §,Karl T. Mueller, ,and Kristin A. Persson* ,,Lawrence Berkeley National Laboratory, Berkeley, California 94720, United States Pacic Northwest National Laboratory, Richland, Washington 99352, United States § Materials Science Division, Argonne National Laboratory, Argonne, Illinois 60439, United States Department of Physics and Astronomy, California State University, Northridge, California 91330, United States Joint Center for Energy Storage Research (JCESR), Lemont, Illinois 60439, United States * S Supporting Information D esigning optimal electrolytes is key to enhancing the performance of energy storage devices,especially relating to cycle life, eciency, and stability. 1 Specically, for future beyond-Li ion energy storage, there is still ample room for electrolyte improvements. Among the candidates for higher gravimetric energy storage, the LiS battery is considered quite promising, owing to its theoretical specic energy density (2600 Wh/kg) and specic capacity (1675 mAh/g) and signicantly lower cost as compared to state-of-art lithium-ion batteries. 24 However, despite these attractive attributes, successful commercialization of LiS batteries is currently hindered by poor cycling performance and capacity retention that is primarily caused by the parasitic reactions between the Li metal anode and dissolved polysulde (PS) species from the cathode during the cycling process. 3,5 Most of the eorts to overcome this degradation mechanism has focused on suppressing the dissolution of PS species and/or protecting the negative electrode using connement strategies or protective layers. However, these additional components not only fail to block completely the PS species but also restrict the volumetric energy density. 2,6 In contrast, less attention has been given to designing optimal electrolytes with reduced PS solubility and improved electrochemical stability. Traditional Li salts used in Li-ion batteries (eg., LiPF 6 , LiBF 4 , LiBOB, LiBF 2 C 2 O 4 ) and solvents (eg., ester, carbonates, phosphates) are unsuitable for LiS battery applications due to their parasitic reactions with PSs. 2 Hence, rational selection or design of the electrolyte is critical in controlling the deleterious shuttle reactions and protecting the electrode surface. So far, 1 M lithium bis(triuoromethanesulfonyl)imide (Li-TFSI) in a binary solvent mixture of 1,3-dioxolane and 1,2-dimethoxy- ethane (DOL:DME) is considered as one of the most suitable electrolytes for LiS cells. 2,5 However, despite its wide usage, the DOL:DME solvent system provides signicant PS solubility, which enables the shuttle process and subsequent parasitic reactions. 2 Therefore, to develop electrolytes with low solubility, high chemical stability, and low viscosity, it is important to improve our understanding of the solvation structure and dynamics of the intermediate PS species formed during discharge. Ion solvation in electrolytes is composed of highly correlated ionion and ionsolvent interactions spanning wide spatial and temporal ranges. Currently, there is limited understanding of the solvation structure of various types of PS species (Li 2 S x ; x = 1 to 8) formed during the discharge process. Experimental eorts based on spectroscopic techni- ques have mostly focused on specic constituents of the electrolyte and do not comprehensively report the solvation structure of PS species. 7,8 Ab initio based computational methods have provided valuable insights about the dispropor- tionation and intermolecular association of PSs in LiS electrolyte systems. 912 Nevertheless, comprehensive under- standing regarding the evolution of solvation phenomena with respect to the PS chain length and solvent system remains elusive. Classical molecular dynamics (MD) simulations are well suited to obtain the needed molecular level understanding of nonreactive interactions and dynamics of multicomponent systems covering larger temporal and lateral scales. As per our knowledge, there is no previous work on understanding bulk structural and dynamical properties of PSs using classical molecular dynamics (MD) simulations mainly due to lack of eective force eld parameters. Various reaction mechanisms have been proposed during the discharge cycle using DFT calculations, 7,13 and spectroelec- trochemical techniques, 7,14 reporting a variety of possible PS reaction intermediates (S 8 2,S 7 2,S 6 2,S 5 2,S 4 2,S 3 2,S 2 2). Although these dierent PS species are postulated to coexist in the solution through chemical equilibrium, 2 it is nevertheless likely that their interactions with the salt and solvent are to a rst-order approximation independent of each other. Hence, in this work we focus on understanding the properties of individual PS species S x ; x = 2 to 8, cognizant of the fact that they may all be present in the solution at some state-of- charge (SOC). As the rst step in this direction, we study the solvation structure and diusion coecient of 0.25 M Li 2 S x (x Received: January 6, 2017 Revised: March 29, 2017 Published: March 31, 2017 Communication pubs.acs.org/cm © 2017 American Chemical Society 3375 DOI: 10.1021/acs.chemmater.7b00068 Chem. Mater. 2017, 29, 33753379

Transcript of Elucidating the Solvation Structure and Dynamics of...

Page 1: Elucidating the Solvation Structure and Dynamics of ...perssongroup.lbl.gov/papers/rajput-2017-solvation-structure.pdf · Elucidating the Solvation Structure and Dynamics of Lithium

Elucidating the Solvation Structure and Dynamics of LithiumPolysulfides Resulting from Competitive Salt and SolventInteractionsNav Nidhi Rajput,*,†,⊥ Vijayakumar Murugesan,‡,⊥ Yongwoo Shin,† Kee Sung Han,‡,⊥

Kah Chun Lau,§,∥ Junzheng Chen,‡ Jun Liu,‡,⊥ Larry A. Curtiss,§,⊥ Karl T. Mueller,‡,⊥

and Kristin A. Persson*,†,⊥

†Lawrence Berkeley National Laboratory, Berkeley, California 94720, United States‡Pacific Northwest National Laboratory, Richland, Washington 99352, United States§Materials Science Division, Argonne National Laboratory, Argonne, Illinois 60439, United States∥Department of Physics and Astronomy, California State University, Northridge, California 91330, United States⊥Joint Center for Energy Storage Research (JCESR), Lemont, Illinois 60439, United States

*S Supporting Information

Designing optimal electrolytes is key to enhancing theperformance of energy storage devices,especially relating

to cycle life, efficiency, and stability.1 Specifically, for futurebeyond-Li ion energy storage, there is still ample room forelectrolyte improvements. Among the candidates for highergravimetric energy storage, the Li−S battery is considered quitepromising, owing to its theoretical specific energy density(2600 Wh/kg) and specific capacity (1675 mAh/g) andsignificantly lower cost as compared to state-of-art lithium-ionbatteries.2−4 However, despite these attractive attributes,successful commercialization of Li−S batteries is currentlyhindered by poor cycling performance and capacity retentionthat is primarily caused by the parasitic reactions between theLi metal anode and dissolved polysulfide (PS) species from thecathode during the cycling process.3,5 Most of the efforts toovercome this degradation mechanism has focused onsuppressing the dissolution of PS species and/or protectingthe negative electrode using confinement strategies orprotective layers. However, these additional components notonly fail to block completely the PS species but also restrict thevolumetric energy density.2,6 In contrast, less attention has beengiven to designing optimal electrolytes with reduced PSsolubility and improved electrochemical stability. TraditionalLi salts used in Li-ion batteries (eg., LiPF6, LiBF4, LiBOB,LiBF2C2O4) and solvents (eg., ester, carbonates, phosphates)are unsuitable for Li−S battery applications due to theirparasitic reactions with PSs.2 Hence, rational selection or designof the electrolyte is critical in controlling the deleterious shuttlereactions and protecting the electrode surface. So far, 1 Mlithium bis(trifluoromethanesulfonyl)imide (Li-TFSI) in abinary solvent mixture of 1,3-dioxolane and 1,2-dimethoxy-ethane (DOL:DME) is considered as one of the most suitableelectrolytes for Li−S cells.2,5 However, despite its wide usage,the DOL:DME solvent system provides significant PSsolubility, which enables the shuttle process and subsequentparasitic reactions.2 Therefore, to develop electrolytes with lowsolubility, high chemical stability, and low viscosity, it isimportant to improve our understanding of the solvationstructure and dynamics of the intermediate PS species formed

during discharge. Ion solvation in electrolytes is composed ofhighly correlated ion−ion and ion−solvent interactionsspanning wide spatial and temporal ranges. Currently, there islimited understanding of the solvation structure of various typesof PS species (Li2Sx; x = 1 to 8) formed during the dischargeprocess. Experimental efforts based on spectroscopic techni-ques have mostly focused on specific constituents of theelectrolyte and do not comprehensively report the solvationstructure of PS species.7,8 Ab initio based computationalmethods have provided valuable insights about the dispropor-tionation and intermolecular association of PSs in Li−Selectrolyte systems.9−12 Nevertheless, comprehensive under-standing regarding the evolution of solvation phenomena withrespect to the PS chain length and solvent system remainselusive. Classical molecular dynamics (MD) simulations arewell suited to obtain the needed molecular level understandingof nonreactive interactions and dynamics of multicomponentsystems covering larger temporal and lateral scales. As per ourknowledge, there is no previous work on understanding bulkstructural and dynamical properties of PSs using classicalmolecular dynamics (MD) simulations mainly due to lack ofeffective force field parameters.Various reaction mechanisms have been proposed during the

discharge cycle using DFT calculations,7,13 and spectroelec-trochemical techniques,7,14 reporting a variety of possible PSreaction intermediates (S8

2−, S72−, S6

2−, S52−, S4

2−, S32−, S2

2−).Although these different PS species are postulated to coexist inthe solution through chemical equilibrium,2 it is neverthelesslikely that their interactions with the salt and solvent are to afirst-order approximation independent of each other. Hence, inthis work we focus on understanding the properties ofindividual PS species Sx; x = 2 to 8, cognizant of the factthat they may all be present in the solution at some state-of-charge (SOC). As the first step in this direction, we study thesolvation structure and diffusion coefficient of 0.25 M Li2Sx (x

Received: January 6, 2017Revised: March 29, 2017Published: March 31, 2017

Communication

pubs.acs.org/cm

© 2017 American Chemical Society 3375 DOI: 10.1021/acs.chemmater.7b00068Chem. Mater. 2017, 29, 3375−3379

Page 2: Elucidating the Solvation Structure and Dynamics of ...perssongroup.lbl.gov/papers/rajput-2017-solvation-structure.pdf · Elucidating the Solvation Structure and Dynamics of Lithium

= 2 to 8) in DOL:DME (1:1) and 0.25 M Li2Sx (x = 2 to 8)with 1 M Li(TFSI) dissolved in a DOL:DME solvent mixtureusing newly developed and well-benchmarked force fieldparameters. Details of the simulation procedure, force field,and experimental procedure are provided in the SupportingInformation (SI). Note that although previous studies havereported the formation of sulfur radical anions during discharge,in this work, we focus on the dianions.7

The present work focuses on three key scientific studiespertaining to Li−S electrolytes: (1) how Li-PS intermediatesinteract with the solvent molecules and the Li-salt; (2) theeffect of salt addition on solvation structure and dynamics; and(3) the effect of PS chain length on the structure and dynamics.We begin our analysis by elucidating the local environment ofLi2Sx in DOL:DME and Li2Sx in Li(TFSI)/DOL:DME usingradial distribution functions (RDF) and coordination numbers(CN) of Li+−Li+, Li+−Sx2−, Li+−TFSI−, Li+−DOL, and Li+−DME. Previous studies have reported that the terminal S atomplays an important role in the PS interactions with Li+.7,15

Hence, the terminal and inner S atom interactions within theLi+−Sx2− RDF for all PSs (except for Li2S2) are clearlydistinguished (Figure S2). Figure 1a shows the RDF of Li2S2 in

an equimolar solvent mixture of DOL:DME. We observed verystrong Li+−S22− interactions as indicated by a Li+−S22− RDFfirst peak at 2.1 Å followed by a shoulder at 4.2 Å. In contrast,the Li−solvent interaction is quite weak due to (1) the lowdielectric constants of both DOL (ε = 7.0) and DME (ε = 7.2)and (2) the strong interaction between Li+ and S2

2−. The RDFsignifying the Li+−Li+ correlation exhibits two sharp peaksfollowed by a shoulder at ∼5.4 Å illustrating highly orderedlong-range (>5 Å) structures where the PSs acts as a bridgebetween the Li+ ions through Li+···S···Li+ ionic chains. Thesnapshot in Figure S3a shows that Li2S2 is not well solvated andpreferentially forms aggregates in the solution. Such agglom-eration of Li2S2 is likely to result in an extremely low solubilityof Li2S2 in DOL:DME. Previous experimental studies16 havealso reported weak Li−solvent interaction even in high donornumber solvents such as DMSO (ε = 46.7) resulting ingenerally low solubility of Li2S2.

16 Hence, strong Li+−S22−bonding is presumed the primary cause for low solubility ofLi2S2. Furthermore, it is likely that Li salts play an importantrole in controlling the solubility of PS intermediates.16 Hence,to understand how the interaction strength between the Li+ andSx

2− changes with the addition of Li salt, we computed the RDFof Li2S2 in a solution with Li(TFSI) in DOL:DME solvent(Figure 1b). In this solution, although strong Li+−S22−interactions are still observed, a significant change is observedin the solvation structure around Li+. Instead of two sharppeaks in the Li+−Li+ RDF, we observe a much smaller peak at

2.9 Å followed by a sharp peak at 3.6 Å and a broader shoulderat 5.4 Å. Similarly, for the Li+−S22− RDF the electrolyte exhibitsa single peak at 2.1 Å with low intensity and a much broadershoulder at larger distances. The Li+−TFSI− RDF shows twopeaks centered at 3.6 and 4.2 Å indicating a cis-conformationwith bidentate and a trans-conformation with monodentatearrangements of the Li+−TFSI− interactions, respectively.17

The changes in the RDF of Li+−Li+ and Li+−S22− stronglyindicates that TFSI− anions are competing for interaction withLi+ and thereby weaken the Li+−S22− interactions leading to adisruption of the Li+−S22− cluster network. Decrease in theLi+−Li+ and Li+−S22− coordination numbers and an increase inLi+−DME coordination number reveals the changes in the localenvironment around Li+ with the addition of Li(TFSI) salt inLi2S2 (Figure S4). As the weakening of the Li+−S22− bond byTFSI− interactions allows more solvent molecules to coordinatewith Li+ we expect the solubility of Li2S2 to increase. We notethat this effect is in contrast to common ion effect theory,6

which predicts the solubility to decrease with the addition ofsalt. Furthermore, a closer look at the Li−solvent interactionsreveals that the linearly structured ether DME has strongerinteraction as compared to the cyclic ether DOL owing tohigher oxygen donor denticity and structural flexibility.18 Suchpreferential solvation of DME over DOL for Li+ has also beenreported by NMR experimental techniques. Despite theweakened Li+−S22− interaction and increased Li+−solventinteractions due to the presence of TFSI−, the enduring Li2S2clustering network is likely to limit the solubility of the PS(Figure S3b).Another, key parameter which affects the solubility is the PS

chain length; hence we endeavor to understand how thecoordination environment of Li+ and Sx

2− changes as a functionof PS chain length. Figure 2a and S5a presents the coordinationbetween Li+−Li+ as a function of PS chain length in a neatsolvent system (i.e., without Li(TFSI) salt). Well-defined sharppeaks at shorter distances (<5 Å) and broader peaks at longerdistances (>5 Å) indicate an ordered polymeric structurepresent for lower polysulfide chains (i.e., Sx with x ≤ 4). In fact,a well-defined peak at ∼7 Å (Figure S6a) suggests theformation of a long-range structured Li2S4 cluster with a size of∼14 Å radius. This finding is in good agreement with theeffective hydrodynamics radius determined by pulse fieldgradient NMR (see Figure S7).5

Recent DFT and AIMD results by Partovi-Azar et al.8 alsosuggest the preferential tendency of Li2S4 toward forming largeionic clusters by analyzing cluster size up to (Li2S4)8 due tocomputational limitation. As MD simulations are capable ofstudying larger length and time scales, we found the largestcluster formed is (Li2S4)15 even at 0.25 M concentration (seeFigure S8a). For higher order PSs (Sx

2− with x > 4), the Li+−Li+ RDF peaks shifts toward larger distances and for Li2S8 onlya single main peak is observed at 4.1 Å, indicating the absenceof long-ranged clustering (see Figure 2a). By analyzing theevolution in bonding of Li+−Sx2−, Li+−DOL, and Li+−DME forvarious PS species, it is clear that the Li+−Sx2− interactionweakens with an increase in PS chain length (see Figure S5). Inparticular, the Li+ interaction with the inner S atoms changesand shifts toward larger distances (see Figure S2) showcasingthe flexibility of higher order PSs in accommodating Li+ in theirfirst solvation shell. Nevertheless, the Li+−solvent interaction(with preferential interaction with DME over DOL) is alsoincreased at the cost of the Li+−Sx2− interaction strength. Suchpronounced Li+−solvent interaction correlates with the

Figure 1. (a) Radial distribution function of Li+−Li+, Li+−S22−, Li+−DME, Li+−DOL in Li2S2/DOL:DME. (b) Radial distribution functionof Li+−Li+, Li+−S22−, Li+−DME, Li+−DOL and Li+−TFSI− in(Li2S2+LiTFSI)/DOL:DME.

Chemistry of Materials Communication

DOI: 10.1021/acs.chemmater.7b00068Chem. Mater. 2017, 29, 3375−3379

3376

Page 3: Elucidating the Solvation Structure and Dynamics of ...perssongroup.lbl.gov/papers/rajput-2017-solvation-structure.pdf · Elucidating the Solvation Structure and Dynamics of Lithium

observed higher solubility of longer PS chains in a DOL:DMEmixture.16 Conversely, the extended (∼few nm) polymericstructure formation can possibly lower the solubility of theshorter PS chains. This finding is in good agreement with ourprevious results that clustering (through polymerization) oflower order PSs lower their solubility.11

The effect of PS chain length on the local environmentaround Li+ is observed for the case when a Li salt is added tothe solution. Figures 2b and S5b show the RDF analyses ofLi+−Li+ and Li+−TFSI− for Li2Sx (x = 2 to 8) in DOL:DME inthe presence of the Li(TFSI) salt. The solvation structure of Li+

as a function of PS species shows a similar trend as thatobserved in the neat solvent system discussed earlier. Multiplesharp peaks in Li+−Li+ RDF at shorter distances are observedfor shorter PSs (Sx with x ≤ 4), indicating clustering withextended-range PS structure. On the other hand, the absence offeatures at larger distances (>5 Å) in longer PSs represents thedecrease in long-range structures such as aggregates and Li-richdomains. Hence, we can expect the solubility to increase withan increase in PS length even in the presence of the salt. Theeffect of the addition of Li salt is found to be minimal in smallerPSs; hence large clusters are still observed for (Li2S4 +LiTFSI)/DOL:DME solution (Figures S8b and S9). As one ofthe dominant PS species during the discharge process,7,13 thelarge cluster observed for Li2S4 is likely to increase theelectrolyte viscosity. This increase in viscosity, in turn, maycontribute to the sharp voltage decline observed in the secondregion of the discharge profile, where liquid−liquid single phasereduction from dissolved Li2S8 to low order PSs is observed. Asdiscussed earlier, the two peaks around 3.5 and 4.2 Å observedin the RDF of Li+−TFSI− are assigned to bidentate and

monodentate conformational arrangements, respectively (Fig-ure 2c). With the increase in PS chain length, the probability ofthe bidentate conformation increases at the cost of themonodentate conformation. The bidentate conformation mayweaken the Li+−Sx2− interaction in longer PS chains by twomechanisms (1) delithiation of PS anions that will enhance thesolvent interaction with the Li+ ions, and (2) increased oxygencoordination and steric hindrance from the TFSI− anions.Figure 2d shows the coordination number of Li+−Li+, Li+−TFSI−, Li+−Sx2−, Li+−DOL, Li+−DME interactions for theLi2Sx (x = 2 to 8) + Li(TFSI) in DOL:DME solution. Theterminal S atom is used to compute the coordination numbersof Li+−Sx2− and the RDF minima at ∼5 Å, which defines thefirst coordination shell of the anion including both mono-dentate and bidentate orientations for Li+−TFSI−. Thecoordination number of Li+−Li+ in Li2S2 indicates that eachLi+ has 1.8 (∼2) other Li+ located within a radius of 4.4 Å,whereas Li2S8 has 0.76 (∼1) other Li+ ions located within 4.6 Å.It is observed that Li+−Li+ and Li+−Sx2− coordination numbersdecrease with an increase in PS chain length indicating asignificant change in Li+ environment. Higher orders PSs aresufficient to provide the desired charge to Li+ ions and thesteric hindrance of PSs result in a lower coordination numbersfor Li+−Sx2− and higher coordination numbers between Li+ andthe solvent. This in turn is expected to increase the solubility oflarger PSs. In particular, the increase in coordination of Li+ withDME is more significant as compared to Li+ with DOL as afunction of PS chain length due to the preferential coordinationof DME with Li+ over DOL as discussed earlier. We find thathigher order PSs exist as monomers and show increasedinteraction with solvent molecules, which may result in fasterreaction kinetics. In contrast, smaller PSs exist as dimers orclusters resulting in less interaction with the solvent molecules,which is likely to slow down the reaction kinetics. As a primaryfunction of Li−S electrolytes is the efficient transport of ionsbetween anode and cathode, while suppressing the dissolutionof PS species, it is essential to understand the effect of solvationstructure on the diffusion coefficient of ionic species, which inturns affects the conductivity. The translational dynamics isstudied by extracting the self-diffusion coefficient of all ionicspecies using MD simulations as well as pulsed-field gradient(PFG) NMR (Figure 3).Figure S10 shows the self-diffusion coefficients of Li+, TFSI−

and Sx2− computed from MD simulations. It is observed that

the diffusion coefficients of Li+, TFSI− and Sx2− decrease with

increased PS chain length, which indicates sluggish dynamics ofionic species and possibly higher viscosity for solutionscontaining the larger PSs. Figure 3 shows a comparisonbetween the self-diffusion coefficients obtained from MDsimulations and PFG-NMR. The trends in self-diffusioncoefficient are in good agreement; however, the simulationdiffusion coefficients are consistently lower than the exper-imental counterparts owing to the well-known limitation ofnonpolarizable force fields.19 A significant decrease in diffusioncoefficients for Li+, TFSI−, DOL as well as DME in Li(TFSI)/DOL:DME is observed as the concentration of Li(TFSI)increases for 0.25 to 3 M indicating that high viscosity and lowionic conductivity can be expected in highly concentratedelectrolytes. For both Li2Sx/DOL:DME and [Li2Sx+Li(TFSI)]/DOL:DME the self-diffusion coefficient of the solutes as well assolvents decrease with an increase in PS chain length (Figure 3)again pointing to the high viscosity and low ionic conductivityof higher order PS solutions. As TFSI− is weakly coordinated to

Figure 2. (a) Radial distribution function of Li+−Li+ in Li2Sx (x = 2 to8) in DOL:DME. (b) Radial distribution function of Li+−Li+ and (c)Li+−TFSI− in Li2Sx (x = 2 to 8) + Li(TFSI) in DOL:DME. (d)Coordination number of Li+−Li+, Li+−TFSI−, Li+−Sx2− (terminal),Li+−DME, Li+−DOL in Li2Sx (x = 2 to 8) + Li(TFSI) in DOL:DME.

Chemistry of Materials Communication

DOI: 10.1021/acs.chemmater.7b00068Chem. Mater. 2017, 29, 3375−3379

3377

Page 4: Elucidating the Solvation Structure and Dynamics of ...perssongroup.lbl.gov/papers/rajput-2017-solvation-structure.pdf · Elucidating the Solvation Structure and Dynamics of Lithium

Li+ compared to Sx2−, its diffusion coefficient is higher than the

diffusion coefficient of Li+ and Sx2− irrespective of the PS chain

length. We note that in all systems considered in this work theself-diffusion coefficients follow the order Sx

2− < Li+ < TFSI−.Furthermore, both NMR and simulation results indicate thatthe self-diffusion coefficients of Li+, PS anions and solvents aredecreased by the addition of salt as well as an increase in PSchain length. Although enhanced ionic transport whilemaintaining low PS solubility is desirable for high performanceLi−S battery operations, one of the most suitable electrolytesystems (Li(TFSI) in DOL:DME) exhibits both highersolubility of larger PS and slower dynamics for larger PSspecies. Thus, the recently proposed high concentration (lean)electrolytes20 may in the case of Li(TFSI): DOL/DMEconsequently suffer from higher PS solubility and slowerionic diffusion of larger PSs. However, we also note that a moredetailed understanding of the diffusion mechanism is requiredto understand if an increase in viscosity is fundamentallyproblematic in all Li−S electrolytes or if high transferencenumbers can balance out these effects.Optimizing the lithium solvation environment in electrolyte

solutions is key to controlling dissolution of polysulfide species.This work provides insights into the interactions betweensolute and solvent molecules in Li−S electrolytes and theirdynamical properties as a function of polysulfide species andlithium salt present in the solution. Our results indicate largecluster formation in lower order polysulfides, whereas highersolubility is observed with an increase in polysulfide chainlength. Stronger Li+−Sx2− bonding is found to be the primarycause for the low solubility of lower order polysulfides.Furthermore, longer polysulfide chain lengths correlate withan increase in Li polysulfide−solvent interactions that allows forhigher solubility, which may result in faster reaction kinetics.The addition of salt (Li−TFSI) weakens the strong Li+−Sx2−networks due to competing interactions between TFSI− andpolysulfide dianions with Li+. This competition results in higherLi+−solvent interactions and increased solubility, which iscontrary to common ion effect theory. However, highlyconcentrated electrolytes might decrease the solubility ofpolysulfides. A high salt concentration also reduces the mobility

of ionic species, which may negatively impact the ionicconductivity of the electrolyte. Overall, an improved under-standing of the intramolecular interactions in the electrolytewill aid in designing solutions with limited solubility ofdetrimental species, without limiting functional propertiessuch as ionic conductivity.

■ ASSOCIATED CONTENT*S Supporting InformationThe Supporting Information is available free of charge on theACS Publications website at DOI: 10.1021/acs.chemma-ter.7b00068.

Details of computational and experimental methodologytogether with additional figures and tables (PDF)

■ AUTHOR INFORMATIONCorresponding Authors*(N.N.R.) E-mail: [email protected]. Tel.: +1-510-486-5389.*(K.A.P.) E-mail: [email protected]. Tel.: +1-510-486-7218.ORCIDNav Nidhi Rajput: 0000-0003-4740-8217Vijayakumar Murugesan: 0000-0001-6149-1702Larry A. Curtiss: 0000-0001-8855-8006Karl T. Mueller: 0000-0001-9609-9516NotesThe authors declare no competing financial interest.

■ ACKNOWLEDGMENTSThis research was intellectually led by the Joint Center forEnergy Storage Research (JCESR), an Energy Innovation Hubfunded by the U.S. Department of Energy (DOE), Office ofScience, Basic Energy Sciences (BES), under Contract No.DEAC02-06CH11357. The calculations were performed usingthe computational resources of the National Energy ResearchScientific Computing Center (NERSC), which is supported bythe Office of Science of the U.S. Department of Energy underContract No. DE-AC02-05CH11231. The NMR measure-ments were performed at the Environmental MolecularSciences Laboratory (EMSL), a national scientific user facilitysponsored by the DOE’s Office of Biological and Environ-mental Research and located at Pacific Northwest NationalLaboratory (PNNL). K. C. Lau acknowledges the grants ofcomputer time through the LCRC Blues Cluster and ALCFVesta Cluster at Argonne National Laboratory.

■ REFERENCES(1) Scheers, J.; Fantini, S.; Johansson, P. A review of electrolytes forlithium−sulphur batteries. J. Power Sources 2014, 255, 204−218.(2) Zhang, S. S. Liquid electrolyte lithium/sulfur battery:fundamental chemistry, problems, and solutions. J. Power Sources2013, 231, 153−162.(3) Bruce, P. G.; Freunberger, S. A.; Hardwick, L. J.; Tarascon, J.-M.Li-O2 and Li-S batteries with high energy storage. Nat. Mater. 2012,11, 19−29.(4) Manthiram, A.; Fu, Y.; Su, Y.-S. Challenges and prospects oflithium−sulfur batteries. Acc. Chem. Res. 2012, 46, 1125−1134.(5) Chen, J.; Han, K. S.; Henderson, W. A.; Lau, K. C.; Vijayakumar,M.; Dzwiniel, T.; Pan, H.; Curtiss, L. A.; Xiao, J.; Mueller, K. T.; Shao,Y.; Liu, J. Restricting the Solubility of Polysulfides in Li-S Batteries ViaElectrolyte Salt Selection. Adv. Energy Mater. 2016, 6, 1600160.(6) Shin, E. S.; Kim, K.; Oh, S. H.; Cho, W. I. Polysulfide dissolutioncontrol: the common ion effect. Chem. Commun. 2013, 49, 2004−2006.

Figure 3. Self-diffusion coefficients of Li+, TFSI−, DOL, DME and PSsin 0.25 M, 1 and 3 M Li(TFSI), 0.25 M Li2Sx (x = 4, 6, 8) inDOL:DME and 0.25 M Li2Sx (x = 4, 6, 8) + 1 M Li(TFSI) inDOL:DME computed from MD simulations and PFG-NMR.

Chemistry of Materials Communication

DOI: 10.1021/acs.chemmater.7b00068Chem. Mater. 2017, 29, 3375−3379

3378

Page 5: Elucidating the Solvation Structure and Dynamics of ...perssongroup.lbl.gov/papers/rajput-2017-solvation-structure.pdf · Elucidating the Solvation Structure and Dynamics of Lithium

(7) Wujcik, K. H.; Pascal, T. A.; Pemmaraju, C.; Devaux, D.; Stolte,W. C.; Balsara, N. P.; Prendergast, D. Characterization of PolysulfideRadicals Present in an Ether-Based Electrolyte of a Lithium−SulfurBattery During Initial Discharge Using In Situ X-Ray AbsorptionSpectroscopy Experiments and First-Principles Calculations. Adv.Energy Mater. 2015, 5, 1500285.(8) Partovi-Azar, P.; Kuhne, T. D.; Kaghazchi, P. Evidence for theexistence of Li 2 S 2 clusters in lithium−sulfur batteries: ab initioRaman spectroscopy simulation. Phys. Chem. Chem. Phys. 2015, 17,22009−22014.(9) Kamphaus, E. P.; Balbuena, P. B. Long-Chain PolysulfideRetention at the Cathode of Li−S Batteries. J. Phys. Chem. C 2016,120, 4296−4305.(10) Assary, R. S.; Curtiss, L. A.; Moore, J. S. Toward a molecularunderstanding of energetics in Li−S batteries using nonaqueouselectrolytes: a high-level quantum chemical study. J. Phys. Chem. C2014, 118, 11545−11558.(11) Vijayakumar, M.; Govind, N.; Walter, E.; Burton, S. D.; Shukla,A.; Devaraj, A.; Xiao, J.; Liu, J.; Wang, C.; Karim, A.; Thevuthasan, S.Molecular structure and stability of dissolved lithium polysulfidespecies. Phys. Chem. Chem. Phys. 2014, 16, 10923−10932.(12) Pascal, T. A.; Pemmaraju, C.; Prendergast, D. X-ray spectros-copy as a probe for lithium polysulfide radicals. Phys. Chem. Chem.Phys. 2015, 17, 7743−7753.(13) Wang, L.; Zhang, T.; Yang, S.; Cheng, F.; Liang, J.; Chen, J. Aquantum-chemical study on the discharge reaction mechanism oflithium-sulfur batteries. J. Energy Chem. 2013, 22, 72−77.(14) Hagen, M.; Schiffels, P.; Hammer, M.; Dorfler, S.; Tubke, J.;Hoffmann, M.; Althues, H.; Kaskel, S. In-situ Raman investigation ofpolysulfide formation in Li-S cells. J. Electrochem. Soc. 2013, 160,A1205−A1214.(15) Pascal, T. A.; Wujcik, K. H.; Velasco-Velez, J.; Wu, C.; Teran, A.A.; Kapilashrami, M.; Cabana, J.; Guo, J.; Salmeron, M.; Balsara, N.;Prendergast, D. X-ray absorption spectra of dissolved polysulfides inlithium−sulfur batteries from first-principles. J. Phys. Chem. Lett. 2014,5, 1547−1551.(16) Pan, H.; Wei, X.; Henderson, W. A.; Shao, Y.; Chen, J.;Bhattacharya, P.; Xiao, J.; Liu, J. On the Way Toward UnderstandingSolution Chemistry of Lithium Polysulfides for High Energy Li−SRedox Flow Batteries. Adv. Energy Mater. 2015, 5, 1500113.(17) Li, Z.; Borodin, O.; Smith, G. D.; Bedrov, D. Effect of OrganicSolvents on Li+ Ion Solvation and Transport in Ionic LiquidElectrolytes: A Molecular Dynamics Simulation Study. J. Phys. Chem.B 2015, 119, 3085−3096.(18) Shao, Y.; Rajput, N. N.; Hu, J.; Hu, M.; Liu, T.; Wei, Z.; Gu, M.;Deng, X.; Xu, S.; Han, K. S. Nanocomposite polymer electrolyte forrechargeable magnesium batteries. Nano Energy 2014, 12, 750−759.(19) Han, K. S.; Rajput, N. N.; Wei, X.; Wang, W.; Hu, J. Z.; Persson,K. A.; Mueller, K. T. Diffusional motion of redox centers in carbonateelectrolytes. J. Chem. Phys. 2014, 141, 104509.(20) Cheng, L.; Curtiss, L. A.; Zavadil, K. R.; Gewirth, A. A.; Shao, Y.;Gallagher, K. G. Sparingly Solvating Electrolytes for High EnergyDensity Lithium−Sulfur Batteries. ACS Energy Lett. 2016, 1, 503−509.

Chemistry of Materials Communication

DOI: 10.1021/acs.chemmater.7b00068Chem. Mater. 2017, 29, 3375−3379

3379